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/0303607 | 1 | 0303 | 2003-03-27T16:41:33 | Near-IR spectra of ISOGAL sources in the Inner Galactic Bulge | [
"astro-ph"
] | In this work we present near-IR spectra (HK-band) of a sample of 107 sources with mid-IR excesses at 7 and 15 $\rm \mu$m detected during the ISOGAL survey. Making use of the DENIS interstellar extinction map from Schultheis et al. (1999) we derive luminosities and find that the $\rm M_{bol}$ vs.~$\rm ^{12}CO$ and $\rm M_{bol} vs. H_{2}O$ diagrams are powerful tools for identifying supergiants, AGB stars, giants and young stellar objects. The majority of our sample are AGB stars (~ 80%) while we find four good supergiant candidates, nine young stellar objects and 12 RGB candidates. We have used the most recent $\rm K_{0}-[15]$ relation by Jeong et al. (2002) based on recent theoretical modeling of dust formation of AGB stars to determine mass-loss rates. However, the uncertainties in the mass-loss rates are rather large. The mass-loss rates of the supergiants are comparable with those in the solar neighbourhood while the long-period Variables cover a mass-loss range from $\rm -5 < log \dot{\it{M}} < -7$. The red giant candidateslie at the lower end of the mass-loss rate range between $\rm -6.5 < log \dot{{\it{M}}} < -9$. We used the equivalent width of the CO bandhead at 2.3 $\rm \mu m$, the NaI doublet and the CaI triplet to estimate metallicities using the relation by Ram\'{\i}rez et al. (\cite{Ramirez2000}). The metallicity distribution of the ISOGAL objects shows a mean [Fe/H] $\sim$ -0.25 dex with a dispersion of $\rm \pm 0.40 dex$ which is in agreement with the values of Ram\'{i}rez et al. (\cite{Ramirez2000}) for Galactic Bulge fields between $\rm b = -4^{o}$ and $\rm b = -1.3^{o}$. A comparison with the solar neighbourhood sample of Lan\c{c}on & Wood (LW) shows that our sample is ~ 0.5 dex more metal-rich on average. | astro-ph | astro-ph | Astronomy & Astrophysics manuscript no. SOFIaccepted
(DOI: will be inserted by hand later)
November 7, 2018
3
0
0
2
r
a
M
7
2
1
v
7
0
6
3
0
3
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
Near-IR spectra of ISOGAL sources in the Inner Galactic Bulge ⋆
M. Schultheis1, A. Lan¸con2, A. Omont1, F. Schuller1, and D. K. Ojha3
1 Institut d'Astrophysique de Paris, CNRS, 98bis Bd Arago, 75014 Paris, France
2 Observatoire de Strasbourg, UMR 7550, 11 rue de l'Universite, 6700 Strasbourg, France
3 Tata Institute of Fundamental Research, Homi Bhabha Road, Colaba, Mumbai, 400 005, India
Received .... / Accepted .. .........
In this work we present near-IR spectra (HK-band) of a sample of 107 sources with mid-IR excesses
Abstract.
at 7 and 15 µm detected during the ISOGAL survey. Making use of the DENIS interstellar extinction map from
Schultheis et al. (1999) we derive luminosities and find that the Mbol vs. 12CO and Mbol vs. H2O diagrams are
powerful tools for identifying supergiants, AGB stars, giants and young stellar objects. The majority of our sample
are AGB stars (∼ 80%) while we find four good supergiant candidates, nine young stellar objects and 12 RGB
candidates. We have used the most recent K0 − [15] relation by Jeong et al. (2002) based on recent theoretical
modeling of dust formation of AGB stars to determine mass-loss rates. The mass-loss rates of the supergiants
are comparable with those in the solar neighbourhood while the long-period Variables cover a mass-loss range
from −5 < log M < −7. The red giant candidates lie at the lower end of the mass-loss rate range between
−6.5 < log M < −9. We used the equivalent width of the CO bandhead at 2.3 µm, the NaI doublet and the CaI
triplet to estimate metallicities using the relation by Ram´ırez et al. (2000b). The metallicity distribution of the
ISOGAL objects shows a mean [Fe/H] ∼ -0.25 dex with a dispersion of ±0.40 dex which is in agreement with the
values of Ram´irez et al. (2000b) for Galactic Bulge fields between b = −4o and b = −1.3o. A comparison with
the solar neighbourhood sample of Lan¸con & Wood (2000) shows that our sample is ∼ 0.5 dex more metal-rich
on average.
Key words. stars: spectroscopy: infrared, extinction-ISM, stars - Galaxy: Bulge, stars: AGB
1. Introduction
The inner galactic Bulge sometimes referred to as the
central stellar cluster (see e. g. Serabyn & Morris 1996),
or as the nuclear Bulge (Mezger et al. 1996), presents
quite extreme conditions (see also Philipp et al. 1999,
Figer 2002). Extending only ∼ 200-300 pc in the galac-
tic plane and ∼ 30-50 pc perpendicular to it,
it con-
tains a mass ∼ 4x109 M⊙, with mean stellar and inter-
stellar densities ∼ 500 times larger than in the galac-
tic disk. The galactic Bulge provides a wide metallicity
range with −1 < [Fe/H] < 1 which makes it an ideal
place for studying stellar evolution. While in the past, sev-
eral studies claimed that we deal in the Bulge with a su-
persolar metal-rich stellar population (see e. g. Whitford
1978, Frogel & Whitford 1987, Frogel 1988, Rich 1988),
chemical abundances there have recently been revised (see
e. g. McWilliam & Rich 1994, Frogel et al. 1999) and at
present the iron relative abundance is believed to peak at
∼ -- 0.3 dex (Freeman & Bland-Hawthorn 2003, van Loon
Send offprint requests to: [email protected]
⋆ Based on observations collected at the European Southern
Observatory, La Silla Chile
et al. 2003, Ibata & Gilmore 1995, Minniti et al. 1995,
Houdashelt 1996), with a wide dispersion. The metallicity
distribution is an important ingredient for the different
scenarios of galaxy formation such as dissipational col-
lapse or accretion of matter. We want to refer to Freeman
& Hawthorn (2003) for the most recent review about the
formation and evolution of our Galaxy.
In most parts of the galactic Bulge, the study of the
stellar population is hampered by its high interstellar ab-
sorption (Frogel et al. 1999, Schultheis et al. 1999); studies
in the Infrared are therefore crucial.
Surveys with the ISO satellite, especially with the
ISOCAM instrument (C´esarsky et al. 1996), whose sen-
sitivity is several orders of magnitude greater than IRAS
and whose angular resolution is ten times better, have
led to new possibilities. The ISOGAL 7 and 15µm survey
(Omont et al. 2003) in particular has observed ∼ 16 deg2
of the central obscured regions of the Galaxy. The total
number of stars detected (∼ 105) is comparable to the
number detected by IRAS in the whole Galaxy. The main
goals of the ISOGAL survey are to quantify the spatial
distributions of the various stellar populations and their
properties in the inner Galaxy, together with the proper-
2
M. Schultheis et al.: Near-IR spectroscopy of ISOGAL sources
ties of the warm interstellar medium. The combination of
ISOGAL and DENIS (or 2MASS) near-infrared data is a
powerful means for determining the nature of sources even
in regions of high extinction (AV up to 20 -- 30). The var-
ious colour-colour and colour-magnitude diagrams avail-
able with the five ISOGAL-DENIS bands provide rich in-
formation on extinctions, distances, intrinsic colours and
absolute magnitudes. Among the M giants, a large pro-
portion of those detected at 15 µm are AGB-LPVs with
mass-loss, well traced by their 15 µm excesses (Omont et
al. 1999, Glass et al. 1999, Alard et al. 2001). Ojha et
al. (2003, hereafter referred to as OOS) studied ISOGAL
sources in the outer Bulge (b & 1◦) and discussed their
nature as well as their mass-loss rates. AGB stars con-
tribute more than 70% towards the replenishment of the
ISM in the solar neighbourhood (Sedlmayr 1994). Thus,
it is important to study the mass-loss of these objects in
different galactic environments.
However, detailed characterization of the sources faces
various difficulties: very large interstellar extinction, un-
certainty in the mid-IR extinction law, photometry un-
certainty in the case of the weakest ISOGAL sources, etc..
Spectroscopic follow-up observations have therefore been
deemed essential. For example, Schultheis et al. (2002) ob-
tained visible spectra of nearby sources with mid-IR ex-
cesses and could identify interesting objects such as Ae/Be
stars, possibly post-AGB stars and stars with red excess.
Optical spectroscopy in the Inner Galactic Bulge fails as
the sources become invisible due to the high interstellar
extinction (AV > 20 mag). Therefore, only spectroscopy
in the near-IR, where interstellar absorption is about ten
times smaller (at ∼ 2.2 µm), can reveal us the nature of
the source.
Up to now the study of stellar populations in the in-
ner Galaxy have mostly been restricted to low extinc-
tion fields, such as Baade's windows (see e. g. Rich 1988,
Terndrup 1988). Recently, however, Ram´ırez et al. (2000b,
hereafter referred to as RSFD) studied the M giant popu-
lation in the inner Bulge using K-band spectra and deter-
mined a number of metallicities (see also Frogel et al. 1999,
Ram´ırez et al. 2000a). Wood et al. (1998), Blommaert et
al. (1998) and Ortiz et al. (2002) studied OH/IR stars
which are known to be the most extreme mass-losing stars
around the Galactic Center while Glass et al. (2001) per-
formed a monitoring program of large-amplitude variables
in the Galactic Centre.
In this work, we perform a spectroscopic follow-up
study (H and K-bands) of 107 ISOGAL sources with IR
excess at 7 and 15 µm in the Inner Galactic Bulge to
study their nature, mainly to identify young stellar objects
and to establish criteria to distinguish between different
classes of objects. We discuss the possibility of separat-
ing the different stellar populations such as AGB stars,
M giants, supergiants and young stellar objects in highly
obscured regions (AV ∼ 20 − 30 mag), combining near-IR
spectroscopy, near and mid-IR photometry and interstel-
lar extinction data further. We discuss the mass-loss of
AGB stars as well as of supergiants and red giant stars.
Using the spectral features of CO, NaI and CaI, we esti-
mate the metallicity and make a comparison to the solar
neighbourhood sample of Lan¸con & Wood (2000, here-
after referred to as LW). The derived metallicity distribu-
tion will be compared to the dissipative collapse model of
Molla et al. (2000).
2. Observations and data reduction
2.1. Near-IR observations
The near-IR spectra were obtained between 16-19 July
2000 with the NTT telescope at ESO, La Silla, Chile, using
the red grism of the SOFI spectrograph. The spectra were
taken under photometric conditions through a 1′′ slit pro-
viding a resolving power of R ∼ 1000, and were recorded
on a Hawai HgCdTe 1024x1024 array. Before each spec-
trum, an image in the K band was taken in order to check
the identification of the source.
For optimal sky subtraction, "ABBA" observing se-
quences were used. The star was moved 10′′ along the slit
between the A and B exposures. The exposure time was
30 s in each position, repeated 2-10 times, depending on
the brightness of the object.
B, A, late F and early G spectrophotometric standard
stars were observed during the night (typically 6-8 stars
per night) to correct for telluric absorption features.
2.2. Data reduction
The data were reduced using MIDAS, the standard ESO
reduction package. After removal of cosmic ray events,
subtraction of the bias level and the dark, all frames were
divided by a normalized flat field. The traces of stars at
the two positions along the slit were used to subtract the
sky. After extraction and co-adding of the spectra, a wave-
length calibration was performed using the Xe-Ne lamp
which gives an accuracy better than 2 A. The spectra were
rebinned to a linear scale, to obtain a dispersion of ∼
10 A/pixel and a resulting wavelength range from 1.5 µm
to 2.5 µm.
We used B, A, F and G standard stars to correct for
the instrumental and atmospheric transmissions. For each
of the standard stars, a suitable model spectrum from
Kurucz (1993) was selected. Significant stellar absorption
features were removed by interpolation both in the mod-
els and the standard star observations (Brγ in B, A stars;
the strongest metal lines in F and G stars). The model
spectra were then divided into the standard star spectra
to provide the combined instrumental and atmospheric
response. The choice of the standard used for the correc-
tion of a particular program star depended on proximity
in time and airmass. We did not apply a second order
correction for airmass, as the residual effects of the tel-
luric bands on the features we were interested in are small
compared to intrinsic variations. For a more detailed dis-
cussion on spectroscopic data reduction in general, we re-
fer the reader to LW. The final spectra are normalized
10
01 00
00 30
00
-00 30
-01 00
-04
M. Schultheis et al.: Near-IR spectroscopy of ISOGAL sources
3
20
30
40
-03
-02
-01
00
GALACTIC LONG.
01
02
03
04
.
I
T
A
L
C
T
C
A
L
A
G
Fig. 1. Observed near-IR spectra (indicated by plus signs) superimposed on the extinction map by Schultheis et
al. (1999). The greyscale on the top shows the range of AV. Galactic longitude and latitude are given in units of
degrees. Note the high interstellar extinction with AV > 20 mag. The comparison fields of RSFD are located at (l = 0,
b = −1.3, −1.8, −2.3, −2.8o) and (b = −1.3o, l = 1.0, 2.0, 3.0, 4.0o)
Table 1. Definition of bandpasses for continuum and fea-
tures
Feature
NaI feature
NaI continuum #1
NaI continuum #2
CaI feature
CaI continumm #1
CaI continuum #2
12CO(2, 0) bandhead
12CO(2, 0) continuum#1
12CO(2, 0) continuum#2
H2O continuum
H2O absorption wing 1
H2O absorption wing 2
band passes (µm)
2.204-2.211
2.191-2.197
2.213-2.217
2.258-2.269
2.245-2.256
2.270-2.272
2.289-2.302
2.242-2.258
2.284-2.291
1.629-1.720
1.515-1.525
1.770-1.780
reference
RDF
RDF
RDF
RDF
RDF
RDF
LW
LW
LW
at 2.28 µm and are dereddened (see Sect. 3.1) using the
extinction values of Schultheis et al. (1999). They are dis-
played in Appendix B. Their classification is discussed in
Sect. 4. The spectra are available electronically at CDS
together with the corresponding finding charts.
2.3. Equivalent widths
Prominent spectral features in our data are the NaI, CaI
and CO(2,0) bands which have also been discussed by
Ram´ırez et al. (1997, hereafter referred to as RDF), and
the CO(6,3) and SiI lines (see Origlia et al. 1993 for
details). In addition, the OH radical has many groups
of prominent lines scattered over the whole H window
(Origlia et al. 1993, LW)
In our analysis we will use the equivalent widths of
the 12CO(2, 0) bands at 2.3 µm (EW(CO)) and the equiv-
alent widths of the NaI (EW(Na)) and CaI (EW(Ca))
lines. Additionally, the water absorption (EW(H2O)) at ∼
1.6 µm has been measured. The adopted index measures
the curvature of the spectrum in the H window due to the
wings of the water bands centered at 1.4 and 1.9 µm. The
index compares the flux in a central passband that is only
weakly contaminated by H2O, to the fluxes in passbands
on either sides. For consistency with the units of the other
feature measurements, its formal expression is that of an
equivalent width (but it takes negative values when H2O
absorption is present). The measured features are identi-
fied in Fig. 2. The exact bandpasses for all measurements
are provided in Table 1.
2.4. The ISOGAL catalogue
The final product of the ISOGAL catalogues (ISOGAL
PSC) at present gives magnitudes, I, J, KS, [7], [15], at
five wavelengths (0.8, 1.25, 2.15, 7 & 15 µm) with DENIS
providing I,J,KS associations when available. We refer for
a detailed description of data reduction and the cross-
identification method to Schuller et al. (2003). Note that
here and elsewhere we use [λ] to denote the ISOGAL mag-
nitude at wavelength λ µm ([7], [15])
3. Sample
Figure 1 shows the location of the ISOGAL sources ob-
served with SOFI superimposed on the extinction map of
Schultheis et al. (1999). The sources were selected from the
ISOGAL fields located at l = −0.27o, b = −0.03o (here-
4
M. Schultheis et al.: Near-IR spectroscopy of ISOGAL sources
hood sample of oxygen-rich stars (M giants, semi-regular
Variables and Mira Variables) of LW.
3.1. Interstellar reddening
The high interstellar extinction in the Galactic Bulge ham-
pers the study of the stellar populations in the inner
parts of our Galaxy. As in most parts of the Galactic
Bulge, interstellar absorption is not homogeneous but oc-
curs in clumps (Glass et al. 1987). Catchpole et al. (1990)
mapped the interstellar extinction around the Galactic
Centre (∼ 2 deg2) using the red giant branch of 47Tuc
as a reference while Frogel et al. (1999) determined the
interstellar reddening for a few fields in the inner Galactic
Bulge using the red giant branch of Baade's window.
Schultheis et al. (1999) constructed a high resolution map
of the whole inner part using DENIS data together with
isochrones calculated for the RGB and AGB phases. We
used their extinction table to deredden our objects (see
Table A1) according to the interstellar extinction law of
Glass (1999) with AV : AJ : AK = 1 : 0.256 : 0.089. For
7 and 15 µm photometry we used A7/AV = 0.020 and
A15/AV = 0.025 (Hennebelle et al. 2001). However, the
extinction curve particularly at 7 and 15 µm is rather un-
certain. One has also to keep in mind that such extinction
values correspond to the peak value of AV on the line of
sight; but the actual value for each individual star may
differ by several magnitudes. In addition, the extinction
values in the very high extincted regions (AV > 25) are
only lower limits. Nevertheless there seems to be on aver-
age a reasonable agreement with the AV values derived by
Wood et al. (1998) for OH/IR stars as discussed by Ortiz
et al. (2002).
3.2. Bolometric magnitudes
Bolometric magnitudes for our ISOGAL objects were ob-
tained by using the multi-band bolometric correction for
AGB stars (Loup et al. 2003). It employs multi-band pho-
tometry with passbands as accurate as possible, taking
into account a mean atmospheric absorption at the site
for ground-based observations. The bolometric corrections
are derived numerically using 69 models for O-rich stars
(see Loup et al. 2003 for a detailed description). The mod-
els are based on Groenewegen's radiative transfer code,
spanning a large range of dust opacities. The advantage
of multi-band bolometric corrections is that the deter-
mination of Mbol is more accurate than with traditional
one-band bolometric corrections. We used for the input
parameter the DENIS J and KS counterparts, the LW2
and LW3 magnitudes of our objects and the AV values
of Schultheis et al. (1999). The main errors in the bolo-
metric corrections result from the interstellar extinction
values giving an error in mbol of at least ∼ 0.2 − 0.3 mag
and the intrinsic depth of the "Bulge" on the line of sight
(Alard et at al. 2001, OOS). Variability of the stars will
introduce additional errors in the determination of Mbol
Fig. 2. Spectrum of an AGB star superimposed by the
most prominent features of SiI, CO(6-3), NaI, CaI, CO(2-
0) and the H2O band. The shaded area indicates the region
of very strong telluric absorption
after referred to as field A), l = −0.62o, b = −0.03o (field
B), l = −0.90o, b = −0.03o (field C), l = −1.21o, b =
−0.03o (field D),
l = −0.44o, b = −0.18o (field E),
l = −1.12o, b = +0.30o (field F), l = −1.12o, b = −0.33o
(field G), l = −5.76o, b = +0.17o (field H). Precise details
of the fields can be found in the Explanatory Supplement
of the ISOGAL catalog (Schuller et al. 2003). We selected
sources with [7] -- [15] > 1.4 and [15] < 8.0 (except for 4
cases, see Fig. 5). These colour criteria were adopted to
favour the detection of dusty young objects with mid in-
frared excesses close to the galactic centre, but they are
also satisfied by a number of evolved stars, as discussed
in Sect. 4. We further restricted the sample to sources
brighter than ∼ 11 mag in KS in order to avoid spurious
associations (Schuller et al. 2003). Such a value is slightly
brighter than the KS completeness limit of DENIS in these
regions (see also Schultheis & Glass 2001). From the 1130
ISOGAL targets obeying the above criteria in these fields,
the 107 sources actually observed with SOFI were selected
approximately at random, with the majority of the sources
(65) in field A.
After the observations, we cross-identified our sample
again with the latest version of the ISOGAL Point Source
Catalogue (Version 1) described in Schuller et al. (2003) in
order to get the final DENIS and ISO photometry. 4 & 5;
see Schuller et al. 2003 for details). 5% of our sample now
show a more doubtful association between ISOGAL and
DENIS (quality flag < 3) and have been dropped. Table
A1 lists the coordinates and the corresponding ISOGAL
and DENIS magnitudes as well as the measured equivalent
widths and the type of each source.
As a comparison sample, we used the low resolution
(R ∼ 1300) K-band spectra of M giants of RSFD which
are located at (l = 0, b = −1.3, −1.8, −2.3, −2.8o) and at
(b = −1.3o, l = 1.0, 2.0, 3.0, 4.0o) and the solar neighbour-
M. Schultheis et al.: Near-IR spectroscopy of ISOGAL sources
5
because of non-simultaneity of the photometry. In Table
2 we list the derived Mbol using a distance modulus of
∆m = 14.5 (or a distance of ∼ 8 kpc).
3.3. Cross-identification with SIMBAD
All the objects have been searched for in the SIMBAD
data base using a search radius of 3′′. As a result, fifteen
objects have been identified as OH/IR stars from the sam-
ple of Wood et al. (1998, see also Ortiz et al. 2002), and
twelve as Long Period variables from Glass et al. (2001).
These subsamples are shown, respectively, in Fig. B3 and
Fig. B4. Four objects (A33, A40, D6, D7) are known ra-
dio sources with IRAS fluxes characteristic of young stellar
objects(see also Sect. 4.3).
Fig. 3. Mbol vs. [7]-[15] diagram for the ISOGAL+SOFI
sample. Known OH/IR stars (Wood et al. 1998) are indi-
cated by filled triangles, LPVs (Glass et al. 2001) by open
squares, while the remaining ISOGAL objects of our sam-
ple are shown by crosses. The two lines indicate the tip of
the RGB and AGB.
4. Classification
4.1. Supergiants
As shown in Fig. 4, we find six very luminous objects
(A17, B14, B15, B19, E5, E13), besides one OH/IR star,
with Mbol . −7.0 (or & 50, 000 L⊙), assuming a distance
of 8 kpc to the Galactic Center), which is approximately
the AGB tip luminosity (Vassiliadis & Wood 1993). They
show very strong CO bands at 2.3 µm with equivalent
widths > 20 A (see Sect. 2.3 for the band passes). In ad-
dition, four of them (A17,B14,B15,B19) do not show any
significant H2O absorption (EW(H2O) < 100 A) which is
an indication of supergiants (Bessell et al. 1991, Lan¸con &
Rocca-Volmerange 1992). The objects E5 and E13 show
very blue colours in K0 −[15] and [7]-[15] and could be late
K or early M-type supergiant candidates. Schuller (2003)
found from a systematic search of very luminous ISOGAL
sources (Mbol < −6) several blue supergiant candidates
with K0 − [15] < 1.0.
In the [15] vs [7] -- [15] diagram (see Fig. 4a), the su-
pergiants (stars) are rather luminous at 15 µm and show
rather blue K0 − [15] colours compared to the AGB
Variables (see Fig. 4b).
4.2. Long Period variables
The majority of our sources we find to be Long-Period
Variables (LPVs) of the AGB, mainly based on their very
deep water absorption features (LW and Fig. 3a). Strong
H2O bands are associated with large period variability
and depend on the atmospheric structure of the star (see
Bessell et al. 1996). Their luminosities are in the range
−4 < Mbol < −7 and their strong CO bands, characteris-
tic of stars on the AGB, also imply that they are likely to
be Miras and semi-regular variables (SRVs). Fig. 6 shows
a comparison between the water bands and the CO bands
of the ISOGAL sample and a solar neighbourhood sample
of semi-regular variables and Miras (LW). The samples
agree surprisingly well within the broad range of the band
strength values of both bands. We identify the ISOGAL
sources having strong H2O absorption as AGB variables
(semi-regular variables or Miras). Greene & Lada (1996)
and G´omez & Mardones (2003) showed that YSOs of
class III can show rather strong H2O and CO bands too.
However, they have rather low luminosities with Mbol > 2
(see G´omez & Mardones 2003), much too faint to be on
the AGB (Mbol < −3.5) Therefore, only additional infor-
mation about the luminosity give us the real confirmation
that our spectra of luminous stars with strong water ab-
sorption are indeed AGB star candidates. (see Fig. 4).
Figure 3 shows the bolometric magnitudes vs. the [7] --
[15] colour ([7] and [15] denote the magnitude at 7 and
15 µm), which is nearly insensitive to interstellar extinc-
tion. Indicated also are the known OH/IR stars and LPVs
in our sample (see Sect. 4.2). It is obvious from Fig. 3 that
photometric information alone is not sufficient to separate
the different categories: additional spectroscopic observa-
tions are necessary.
In Fig. 4a one can see, except for supergiants and the
other very bright sources (Mbol . −6.0), some correlation
between the strength of the water bands and the bolo-
metric magnitude in the sense that more luminous objects
show deeper water absorption. The [15] vs. K0 − [15] dia-
gram (Fig. 5) shows that the LPVs have a wide K0 − [15]
colour range (0 < K0 − [15] < 7) indicating a large range
of mass-loss rates (see discussion below).
6
M. Schultheis et al.: Near-IR spectroscopy of ISOGAL sources
Fig. 4. Equivalent width of H2O vs. Mbol (a) and K0 vs equivalent width of CO (b). YSOs are indicated by filled
circles, OH/IR stars (Wood et al. 1998) by filled triangles, LPVs (Glass et al. 2001) by open squares, candidates of
red giants by open circles, supergiant candidates by stars and AGB Variables by crosses. The dotted line indicated
the tip of the RGB. The two AGB stars with Mbol ≤ −7 are probably blue supergiant candidates (see text).
Twelve LPVs from our sample and five OH/IR stars
(see Table A1) were observed by Glass et al. (2001) and
we have therefore additional information about their pe-
riods and their K amplitudes. The period range is rather
narrow, starting from ∼ 400d up to 800d. The sources
within the sample follow a period-luminosity relation but
we do not find any relation between the CO or water band
strengths and the amplitude.
4.3. Young stellar objects
A33, A40, D6 and D7 are already known to be young
stellar objects. They have been detected by IRAS with
very red colours and also show radio emission. They have
more-or-less featureless spectra (see Fig. B.1) with no CO
absorption at 2.3 µm but in some cases possess hydrogen
absorption lines (Brackett series). The objects B23, B27,
B35, B37, C19 are newly found young stellar objects show-
ing the same spectral features as the four known ones.
Greene & Lada (1996) presented the first system-
atic spectroscopic survey of YSOs, comprising a sample
from the Ophiucus molecular cloud. They found that the
strengths of atomic and CO absorption features are closely
related to the evolutionary state. The line strengths de-
crease from the Class III phase through Class II to the
self-embedded Class I, where the absorption features are
absent. All nine YSOs of our sample show more-or-less
featureless spectra with no CO absorption lines and no
Brγ emission. Thus, we associate them with young stellar
objects of Class I.
However, Fig. 5 shows that YSOs cannot be sepa-
rated unambiguously from other stellar populations by
using ISOGAL colours alone. Felli et al. (2000) give a cri-
terion for selecting YSO candidates (see Fig. 4) in the
ISOGAL [15] vs. [7] -- [15] diagram and Felli et al. (2002)
present a catalog of YSO candidates with [15] < 4.5 and
[7] − [15] > 1.8. The criteria defined by Felli et al. (2002)
are satisfied by seven of the nine identified YSOs in this
region (see Fig. 4a). However, as shown in Fig. 5, there
are also 32 AGB variables, eight OH/IR stars and four
supergiants in the same region. Felli et al. (2002) point
out also that [7] − [15] ≥ 2.5 is a more conservative cri-
terion for identifying YSOs. Six of the YSOs meet this
criterion but it is still satisfied by seven objects which
are probably AGB stars. Thus, an unambiguous separa-
tion between YSOs and evolved stars can only be made
by using additional spectroscopic information.
4.4. M giants
In contrast to the typical LPV, M giants show nearly no
water vapor bands. Using an additional luminosity crite-
rion, namely Mbol fainter than -- 3.5 (see Fig. 3), which is
approximately the tip of the RGB (see Tiede et al. 1996,
Omont et al. 1999), we find 12 objects. Figure 5 shows
that indeed the M giant candidates populate the lower
end in both the [15] vs [7] -- [15] and [15] vs. K0 − [15] dia-
gram. However, as seen in Fig. 6, the separation between
M giants and variable AGB stars in the CO vs. H2O plane
is ambiguous. Some of the stars classified as giants might
M. Schultheis et al.: Near-IR spectroscopy of ISOGAL sources
7
Fig. 5. [15] vs. [7] -- [15] diagram (left panel) and [15] vs K0 − [15] diagram (right panel). Symbols are the same as in
Fig. 4. The dotted line shows the region searched by Felli et al. (2000) for YSO candidates (see text). The long-dashed
line indicates a more conservative criterion for identifying YSOs (Felli et al. 2002 )
be AGB stars that happen to be caught at a phase of pul-
sation where H2O is not very strong (see LW). Additional
multi-epoch observations are necessary, especially because
it has recently been shown that most M giants with late
spectral types are variable (Alcock et al. 2000, Glass &
Schultheis 2002).
5. Discussion
Figure 5 shows that near and mid-IR photometry alone
are not sufficient for distinguishing the different stellar
populations; additional near-IR spectroscopy is necessary.
Using the Mbol vs H2O and the 12CO vs. K0 diagram (see
Fig. 4) we can identify supergiants, AGB Variables, red
giants and young stellar objects. In addition, mid-IR data
at 7 and 15 µm enables us to estimate mass-loss rates of
the stellar population while the equivalent widths of the
CO bandhead, the NaI doublet and the CaI triplet give
us estimates of the metallicity.
5.1. Mass loss
One of the most promising tools for determining mass-
loss rates is the combination of near-IR and mid-IR colour
such as the IRAS K0 -- [12] or the ISOGAL K0 -- [15] colours
(see e. g. Whitelock et al. 1994, Habing 1996, Le Bertre
& Winters 1998, Omont et al. 1999, Jeong et al. 2002,
OOS etc.). We will use the most recent colour-mass loss
relation for oxygen-rich AGB stars by Jeong et al. (2002)
which is based on a consistent time dependent treatment
of hydrodynamics, thermodynamics, equilibrium chem-
Fig. 6. Equivalent width of H2O vs. CO, using the sam-
ple of oxygen-rich LPVs of LW for comparison. YSOs are
indicated by filled circles, OH/IR stars by filled triangles,
LPVs by open squares, candidates of giants by open cir-
cles, supergiant candidates by stars and AGB variables by
crosses. Filled squares indicate the comparison sample of
semi-regular variables and Miras of LW.
8
M. Schultheis et al.: Near-IR spectroscopy of ISOGAL sources
From this relation, the OH/IR stars in our sample
cover mass-loss rates ranging from −6.0 < log M <
−4.5 while LPVs range between −6.5 < log M < −5.
Considering the uncertainties (see above) these values lie
within the ranges determined by Ortiz et al. (2002) for
OH/IR stars and Alard et al. (2001) for LPVs. Our red
giant candidates show mass-loss rates in the range be-
tween −8 < log M < −6. Origlia & Ferraro (2002) de-
termined mass-loss rates of red giants in globular clusters
using ISOCAM observations. They find mass-loss rates
in the range of −7 < log M < −6 assuming a gas-to-dust
mass ratio of 200 and using the DUSTY code. Their mass-
loss range is much narrower than we find. However, taking
the color-mass loss relation of Jeong et al. (2002) would
result in at least 10 times smaller mass-loss rates for the
sample of Origlia & Ferraro (2002). It is obvious that red
giants certainly contribute significantly to the integrated
mass-loss (see also OOS). We want to stress that this is
the first attempt to quantify mass-loss rates of red giants
in the central part of our Galaxy. A systematic study of
these objects in the inner Bulge is certainly needed to de-
rive accurate number densities and their contribution to
the integrated mass-loss.
Our 4 supergiant candidates show mass-loss rates be-
tween −6 < log M < −5.5 assuming that the relation by
Jeong et al. (2002) is valid for supergiants. However, as
pointed out by Josselin et al. (2000) the gas-to-dust ratio
for these kind of objects is rather uncertain. Therefore a
detailed comparison with the solar neighbourhood sample
of Josselin et al. (2000) is at this stage rather difficult.
5.2. Spectral determination of luminosity class
Do spectra in the H and K bands offer the possibility of
determining the luminosity class? Ram´ırez et al. (1997)
used the quantity log [EW(CO)/(EW(Na) + EW(Ca))] to
distinguish between giants and dwarfs over the effective
temperature range between 3400 and 4600 K (see their
Fig. 11). They argued that this quantity might be a pow-
erful luminosity diagnostic. Figure 8 shows that the M
giants of RSFD (indicated by the two straight lines) and
our sample agree quite well. For comparison, we calculated
also the values for the LW sample. We find that there
is no clear separation between supergiants and M giants
or LPVs. However, as seen in Fig. 8, the dispersion in
log [EW(CO)/(EW(Na) + EW(Ca))] is much narrower for
M giants than for variable AGB stars. Pulsation is respon-
sible for the extended atmospheres of LPVs (Feuchtinger
et al. 1993) which is seen in the large dispersion of
log [EW(CO)/(EW(Na) + EW(Ca))] of the LPVs and
OH/IR stars. We get similar results using the LW sample.
While the quantity log [EW(CO)/(EW(Na) + EW(Ca))]
is a good discriminator between dwarfs and giants, it can
not be adopted to separate LPVs from supergiants and
red giants.
Fig. 7. Mass-loss rate vs. 15 µm magnitude vs. for the
present sample. OH/IR stars are indicated by filled trian-
gles, LPVs by open squares, red giant candidates by open
circles and supergiants by stars.
istry and dust formation. They give an explicit relation
between M and the (K − [15])0 colour.
log M = −6.83/(K − [15])0 − 3.78
[M⊙/yr]
(1)
One has to be aware that using this relation between
M and K0 − [15] for our sample includes uncertainties in
the determination of the mass-loss rate arising from the
following causes: (1) Our de-reddend (K − [15])0 colours
are strongly affected by the uncertainty in the determi-
nation of interstellar extinction, in particularly in these
highly extincted regions (see Schultheis et al. 1999). This
could lead easily to errors in M of a factor of 2 -- 3; (2) The
relation was derived for oxygen-rich AGB stars in the solar
neighbourhood. Habing et al. (1994) argued that metal-
licity affects the dust to gas ratio and the outflow velocity
from evolved stars which is directly related to the mass-
loss and that, therefore, the color- M relation might dif-
fer in different galactic environments such as between the
Galactic Bulge and the Magellanic Clouds; (3) It is im-
portant to emphasize that the KS magnitudes of DENIS
are single epoch measurements. The average K amplitude
of our sources associated with known LPVs (Glass et al.
2001) is ∼ 1.0 mag, which gives a factor of ∼ 2 -- 5 uncer-
M for −7 < log M < −5.
tainty in the determination of
We want to emphasize that the colour mass-loss relation
by Jeong et al. (2002) is calculated for variable AGB stars.
Its adaption to non-variable red giants and supergiants has
to be questioned.
M. Schultheis et al.: Near-IR spectroscopy of ISOGAL sources
9
Fig. 8. The luminosity indicator of Ram´ırez et al. (1997)
vs. CO band strengths. The two straight lines show the
location of the M giants of Ram´ırez et al. (1997). The
symbols are the same as in Fig. 3
5.3. Metallicity
Recently, RSFD and Frogel et al. (2001) have obtained
a new metallicity scale for luminous red giants based on
equivalent width measurements of the CO bandhead, the
NaI doublet and the CaI triplet (see Table 1). Their cali-
bration is based on giants in globular clusters for −1.8 <
[Fe/H] < −0.1 . We have used the following relation (so-
lution 1, RSFD)
[Fe/H]index
RSFD = -1.782 + 0.352 EW(Na) - 0.0231 EW(Na)2
-0.0267 EW(Ca) + 0.0129 EW(Ca)2
+ 0.0472 EW(CO) - 0.00109 EW(CO)2
(2)
where EW(Na), EW(Ca) and EW(CO) are the equiv-
alent widths of NaI, CaI and 12CO(2, 0) (see Table 1). As
described in RA97, the typical errors in the determination
of [Fe/H] are of the order of ∼ 0.1 dex. RSFD provide a
second expression for [Fe/H]index
RSFD, that reduces the scat-
ter in their data with correction terms based on (J − K)0.
As most of our program stars do not have measurable J
magnitudes due to heavy extinction, this second relation
could not be used.
In addition, we determined this index from Eq. (2) for
the sample of oxygen-rich stars (M giants, semi-regular
variables and Miras) in the solar neighbourhood of LW
using the same bandpasses. We used repeated observations
of LW for semi-regular and Mira variables, such as BD
Hya, R Cha or KV Car. For the same star the metallicity
index of RSFD can vary by ∼ 0.5 dex showing that it does
Fig. 9. Distribution of metallicity. The solid lines is our
ISOGAL sample located in the inner Bulge; the dotted
line is the solar neighbourhood sample of oxygen-rich AGB
stars and M giants by LW.
not actually measure metallicity in individual spectra of
strongly variable stars.
Fig. 9 shows the distribution of the RSFD metallic-
ity index of our ISOGAL objects superimposed on the
solar neighbourhood sample of LW. If we use only the
non-varying static stars of LW and our sample of red gi-
ant candidates, the mean [Fe/H]index
RSFD does not change, al-
though the dispersion gets smaller. This shows that, even
if the metallicity index of RSFD is not strictly valid for
single observations of a variable star, it can be used on av-
erage over a whole pulsation cycle. This is important for
stellar population studies, as our results show that very
time consuming multi-epoch observations of LPVs are not
necessary in order to obtain average metallicities. We will
only use average values in the remaining discussion.
The mean [Fe/H]index
RSFD is −0.25 dex, with a disper-
sion of ±0.40 dex, which is in agreement with the val-
ues obtained for the static sample of RSFD. Their mean
[Fe/H]index
RSFD is −0.21 dex with a dispersion of ±0.30 dex.
This suggests that the ISOGAL sample and the sample
of RSFD have similar metallicities despite their spatial
separation.
Superimposed is the solar neighbourhood sample of
RSFD of ∼ −0.6 dex (see Fig. 9).
LW with an average [Fe/H]index
10
M. Schultheis et al.: Near-IR spectroscopy of ISOGAL sources
It is known that there is a large spread in the solar
neighbourhood metallicity distribution (see e. g. Haywood
2001) with −0.3 < [Fe/H]index
RSFD < 0; an offset of ∼ 0.3 dex
with respect to the solar neighbourhood sample of LW is
apparent. The relation of RSFD is based on the calibra-
tion of globular clusters and thus is biased towards metal-
poor stars. At higher metallicities the CO band reaches a
plateau and becomes insensitive to changes in [Fe/H]index
RSFD.
Since solution 1 has a strong dependence on EW(CO), it
will therefore underestimate [Fe/H]index
RSFD for larger metal-
licities. However, Fig. 9 shows that our ISOGAL sample
as well as the sample of RSFD is ∼ 0.4 dex metal richer
than the solar neighbourhood sample of LW. Haywood
(2001) gives a modelled and an observed age-metallicity
relation (see their Fig. 14a). According to that, a [Fe/H] of
−0.6 dex corresponds to ages of about 8 -- 9 Gyr and thus
for AGB stars an initial mass of the order of one solar
mass. So, according to the age-metallicity relation, the
solar neighbourhood sample of LW have initial masses
around 1 M⊙, in agreement with mass estimates derived
from pulsation periods (see Table 8 of LW). The differ-
ence in the metallicity distribution in the solar neighbour-
hood and in the Bulge results probably from both different
mean stellar ages and different mean age-metallicity rela-
tions due to different star formation histories and chemi-
cal evolution histories. However, the validity of such age-
metallicity relations is still on debate (see e.g. the recent
review of Freeman & Bland-Hawthorn 2003)
The accurate determination of mean stellar metallici-
ties is essential for constraining models of star formation
and chemical evolution in the Bulge. Our mean [Fe/H]
value is consistent with previous chemical abundance stud-
ies of the galactic Bulge (see e. g. McWilliam & Rich 1994,
Minniti et al. 1995, Houdashelt 1996, Frogel et al. 1999,
RSFD) where the metallicity peaks ∼ −0.3 dex. Moreover,
Molla et al. (2000) model the evolution of the Galactic
Bulge which assumes a dissipative collapse of the gas from
a protogalaxy or halo to form the Bulge and the disk. They
predict a mean stellar Bulge metallicity [Fe/H] ∼ −0.2 dex
with a dispersion of ±0.40 dex.
A chemical abundance gradient in the Bulge is one
characteristics of a Bulge formed by a dissipative col-
lapse. The determination of the metallicity gradient in
the central regions of our Galaxy is important for test-
ing models of Galaxy formation. Up to now, only a few
studies of the metallicity gradient exist in the inner Bulge
(b < 30). Frogel et al. (1999) found a metallicity gradi-
ent based on near-IR photometric data while RSFD could
not find any evidence. The dissipative collapse model of
Molla et al. (2000) predict a steep metallicity gradient
(−0.8 dex kpc−1) in the inner Bulge (R < 500 pc) which is
not supported by the data of RSFD. Our sample of near-
IR spectra presented here, which agrees with the metal-
licity distribution of RSFD, seems to confirm the absence
of a metallicity gradient in the Bulge.
We want to stress that our data set is not sufficient to
do a large statistical analysis of chemical abundances in
the galactic Bulge. Therefore a more detailed discussion
of the implications of the results is beyond the scope of
this paper; a large observational program is necessary.
It is obvious from the discussion above that only realis-
tic models of M giants including the effects of metallicity
can enable us to make a quantitative determination of
their metallicity. First tests of synthetic spectra based on
hydrostatic MARCS model atmospheres (in collaboration
with B. Aringer) and including a complete atomic line list
already show some promising results (Aringer et al. 2002).
However, for AGB stars (the majority of our stars) hydro-
dynamical models would be more appropriate (see e. g.
Hofner 1999, Aringer et al. 2000).
6. Conclusions
We have studied a sample of 107 near-IR spectra of
ISOGAL sources with excesses at 7 and 15 µm in the in-
nermost parts of the galactic Bulge where the interstellar
extinction is high and clumpy. We have shown that us-
ing the molecular bands of CO and H2O, together with
the bolometric magnitudes and the interstellar extinction
values, one can reasonably well separate AGB stars, red
giants, supergiants and young stellar objects. We have
found four supergiant candidates, twelve red giant can-
didates and nine YSOs, while the rest are probably vari-
able AGB stars. We have used the most recent K0 − [15]
M relation by Jeong et al. (2002) which is based on
vs.
a self-consistent time dependent model of dust formation
in AGB stars to determine mass-loss rates. We empha-
size that the color- M relation has been determined from
a model of AGB/LPVs and gives only an indication of
the mass-loss rate. From our sample, OH/IR stars cover
mass-loss rates ranging from −6 < log M < −4.5 while
LPVs range between −6.5 < log M < −5. Our red gi-
ant candidates show mass-loss rates in the range between
−8 < log M < −6. However, this selection is biased in fa-
vor of large M . A comparison with mass-loss rates of red
giants in globular clusters (Origlia & Ferrero 2002) shows
that our mass-loss range is much broader towards larger
M . However, the uncertainty in the determina-
values of
tion of
M is rather large.
While the quantity log [EW(CO)/(EW(Na) + EW(Ca))]
of RDF is a good discriminator between giants and dwarfs,
it cannot be used to separate supergiants and LPVs from
red giants. However, the dispersion is much narrower for
M giants than for variable AGB stars.
We have used the metallicity index [Fe/H]index
RSFD of
RSFD, determined from the line strengths of CO, NaI
and CaI, to estimate the metallicity of the stellar popula-
tion. Our mean [Fe/H]index
RSFD is −0.25 dex with a dispersion
of ±0.40 dex which is in agreement with the values ob-
tained by RSFD and supports the argument of RSFD that
there is no metallicity gradient in the Bulge. Our mean
[Fe/H]index
RSFD is consistent with previous abundance studies
of the galactic Bulge and with the multiphase evolution
models of Molla et al. (2000) which assume a dissipative
collapse of a protogalaxy to form the Bulge and the galac-
tic disk. Our results confirm that even if the metallicity
M. Schultheis et al.: Near-IR spectroscopy of ISOGAL sources
11
index of RSFD is only valid for M giants, it can be used
to estimate an average metallicity valid over a whole pul-
sation cycle. This is an important result for future stellar
population studies using multi-fiber spectrographs (such
as NIRMOS, GIRAFFE,etc.). However, there is a press-
ing need for a grid of realistic models of M giants with
different [Fe/H] to facilitate a quantitative determination
of metallicities, as well as further modeling of LPV/AGB
stars.
Acknowledgements:
We want to thank I. S. Glass, J. Blommaert, J. van
Loon for their fruitful comments and discussions. MS
is supported by the Fonds zur Forderung der wis-
senschaftlichen Forschung (FWF), Austria, under the
project number J1971-PHY.
This research is supported by the project 1910-1
of Indo-French Center for the Promotion of Advanced
Research (CEFIPRA).
is supported,
The DENIS project
in France by
the Institut National des Sciences de l'Univers, the
Education Ministry and the Centre National de la
Recherche Scientifique, in Germany by the State of Baden-
Wurtemberg, in Spain by the DGICYT, in Italy by the
Consiglio Nazionale delle Ricerche,
in Austria by the
Fonds zur Forderung der wissenschaftlichen Forschung
und Bundesministerium fur Wissenschaft und Forschung.
This research has made use of the Simbad database,
operated at CDS, Strasbourg, France.
References
Alard, C., Blommaert, J. A. D. L., Cesarsky, C., 2001, ApJ
552, 289
Alcock, C., Allsman, R. A., Alves, D. R., et al., 2000, AJ 119,
2194
Aringer, B., Kerschbaum, F., Hron, J., 2000, in " The 2nd
ISO workshop on analytical spectroscopy", eds. A. Salama,
M. F. Kessler, K. Lech & B. Schulz, ESA-SP 456
Aringer B., Jørgensen U. G., Schultheis M., et al., 2002,
in "Determining Stellar parameters for red giants", IAU
symp. 210, Uppsala, Sweden
Bessell, M. S., Brett, J. M., Scholz, M., Wood, P. R., 1991,
A&ASS 89, 335
Bessell, M. S., Scholz, M., Wood, P. R., 1996, A&A 307, 481
Blommaert, J. A. D. L., van der Veen, W. E. C. J., van
Langevelde, et al., 1998, A&A 329, 991
Catchpole, R. M., Whitelock, P. A., Glass, I. S., 1990, MNRAS
247, 479
Cesarsky, C., Abergel, A., Agnese, P. et al. 1996 A&A, 315L,
32
Felli, M., Comoretto, G., Testi, L., et al., 2000, A&A 362, 199
Felli, M., Testi, L., Schuller, F.,et al., 2002, A&A 392, 971
Feuchtinger, M. U., Dorfi, E. A., Hofner, S., 1993, A&A 273,
513
Figer D. F., 2002, in "A Massive Star Odysee, from Main
Sequence to Supernova", IAU symp. 212, eds. K. A. van
der Hucht % C. Esteban (astro-ph/0207300)
Frogel J.A., Whitford A. E., 1987, ApJ 320, 199
Frogel J. A., 1988, ARA&A 26, 51
Frogel, J. A., Tiede, G. P., Kuchinski, L. E., 1999, AJ, 117,
2296
Frogel, J. A., Stephens, A., Ram´ırez, S. V., et al., 2001, AJ
122, 1896
Glass, I. S., Catchpole, R. M., Whitelock, P. A., 1987, MNRAS
227, 373
Glass, I. S., 1999, Handbook of infrared astronomy, Cambridge
University Press
Glass, I.S., Ganesh, S., Alard, C. et al., 1999, MNRAS 308, 127
Glass I. S., Matsumoto S., Carter B. S., Sekiguchi K., 2001,
MNRAS 321, 77
Glass, I. S., Schultheis, M., 2002, MNRAS 337, 519
G´omez M., D. Mardones,
submitted
2003,
to AJ,
(astro-ph/0301081)
Greene, T. P., Lada, C. J., 1996, AJ 112, 2184
Habing, H. J., Tignon, J., Tielens, A. G. G. M., 1994, A&A
286, 523
Habing, H.J., 1996, A&AR 7, 97
Haywood, M., 2001, MNRAS 325, 1365
Hennebelle, P., P´erault, M., Teyssier, D., et al., 2001, A&A
365, 598
Hofner, S., 1999, A&A 346, L9
Houdashelt, M., 1996, PASP 108, 828
Ibata R. A., Gilmore G. F., 1995, MNRAS 275, 591
Jeong K. S., Winters J. M., Le Bertre T., et al., 2002, in "Mass-
losing Pulsating stars and their Circumstellar Matter", eds.
Y. Nakasa, M. Honma and M. Sekiin
Josselin, E., Blommaert, J. A. D. L., Groenewegen, M. A. T.,
Li F. L., 2000, A&A 357, 225
Kurucz, 1993, CDrom Number 13 (provided by the author)
Lan¸con, A., Rocca-Volmerange, B., 1992, A&AS 96, 593
Lan¸con, A., Wood, P.R., 2000, A&AS, 146, 217 (LW)
Le Bertre, T., Winters, J. M., 1998, A&A 334, 173
Loup, C., Groenewegen, M. A. T., Cioni M. R., et al., 2003,
submitted to A&A
Mathis J. S., 1990, ARA&A 28, 37
Mezger P. G., Dusch W. J., Zylka R., 1996, A&AR 7, 289
McWilliam, A., Rich, R. M., 1994, ApJS 91, 749
Minniti, D., Olszewski, E. W., Liebert, J., et al., 1995, MNRAS
277, 1293
Molla, M., Ferrini, F., Gozzi, G., 2000, MNRAS 316, 345
Ojha, D. K., Omont, A., Schuller, F., et al., 2003, A&A, sub-
mitted to A&A (OOS)
Omont, A., Ganesh, S., Alard, C. et al., 1999, A&A 348, 755
Omont, A., Gilmore, G., Alard, C., et al., 2003, A&A, in press
Origlia, L., Moorwood, A. F. M., Oliva, E., 1993, A&A 280,
536
Origlia L., Ferraro F. R., 2002, ApJ 571, 458
Ortiz, R., Blommaert, J. A. D. L., Copet, E., et al., 2002, A&A
388, 279
Philipp S., Zylka R., Mezger P. G., Duschl W. J., Herbst T.,
Tuffs R. J., 1999, A&A 348, 768
Ram´ırez, S. V., DePoy, D. L., Frogel, J. A. et al., 1997, AJ 113,
1411 (RDF)
Ram´ırez, S. V., Sellgren, K., Carr, J. S., et al., 2000a, ApJ 537,
205
Ram´ırez, S. V., Stephens, A. W., Frogel, J. A., et al., 2000b,
AJ 120, 833 (RSFD)
Fluks, M. A., Plez, B., Th´e, P. S., et al.. 1994, A&AS 141, 313
Freeman K., Bland-Hawthorn J., 2003, submitted to ARA&A,
Rich, R. M., 1988, AJ 95, 828
Schuller, F., Ganesh, S., Messineo, M., et al., 2003, A&A in
(astro-ph/0208106)
press
12
M. Schultheis et al.: Near-IR spectroscopy of ISOGAL sources
Schuller, F., PhD thesis
Schultheis, M., Ganesh, S., Simon, et al., et al., 1999, A&A
349, L69
Schultheis, M., Glass, I. S., 2001, MNRAS 327, 1193
Schultheis, M., Parthasarathy, M., Omont, A., et al., 2002,
A&A 386, 899
Sedlmayr, E., 1994, in "Molecules in the Stellar Environment",
U G. Jorgensen (ed.), Springer, Berlin, p.163
Serabyn, E., Morris, M., 1996, Nature 382, 602
Terndrup, D. M., 1988, AJ 96, 884
Tiede, G. P., Frogel, J. A., Terndrup, D. M., 1996, AJ 110,
2788
van Loon, J. Th., Gilmore, G. F., Omont, A., et al., 2003,
MNRAS 338, 857
Vassiliadis, E., Wood, P. R., 1993, ApJ 413, 641
Whitelock, P. A., Menzies, J. W., Feast M. W., et al., 1994,
MNRAS 267, 711
Whitford A. E., 1978, ApJ 226, 777
Wood, P. R., Habing, H. J., McGregor, P. J., 1998, A&A 336,
925
Appendix A: Tables
Appendix B: Near-IR spectra
The spectra were dereddened using the extinction curve
by Mathis et al. (1990) and the DENIS extinction map
(Schultheis et al. 1999), see Section 3.1. The flux densi-
ties per unit wavelength are normalized at 2.28 µm. The
spectra are electronically available at CDS. Strong telluric
bands are seen between 1.8 and 1.9 µm.
Table A1: Coordinates (J2000), magnitudes (DENIS,ISOGAL),AV values, equivalent widths (in A) of 12CO(2, 0),NaI,CaI,H2O,Mbol,log M
name
right ascension
A3
A4
A5
A6
A7
A8
A9
A10
A12
A13
A14
A16
A17
A18
A19
A20
A21
A22
A23
A24
A25
A26
A27
A28
A29
A30
A31
A33
A34
A35
A36
A38
A39
A40
A41
A43
A45
A47
A49
A50
A51
A52
17 44 23.8
17 44 30.4
17 44 31.7
17 44 31.8
17 44 34.4
17 44 35.7
17 44 36.4
17 44 39.7
17 44 44.5
17 44 44.6
17 44 48.0
17 44 48.6
17 44 49.1
17 44 49.5
17 44 50.3
17 44 52.7
17 44 52.9
17 44 53.1
17 44 54.2
17 44 54.7
17 44 55.5
17 44 56.8
17 44 57.0
17 44 57.1
17 44 57.8
17 44 57.9
17 44 58.9
17 44 59.5
17 44 59.6
17 45 1.0
17 45 1.7
17 45 4.8
17 45 4.8
17 45 4.9
17 45 5.3
17 45 7.0
17 45 9.8
17 45 10.8
17 45 12.2
17 45 13.5
17 45 14.0
17 45 14.3
declination
-29 8 55.3
-29 7 15.5
-29 6 20.6
-29 17 10.5
-29 10 38.5
-29 13 5.4
-29 9 25.2
-29 16 45.9
-29 5 38.3
-29 7 33.6
-29 6 49.8
-29 0 13.2
-29 19 54.2
-29 3 15.9
-29 19 21.5
-29 14 11.1
-29 7 6.7
-28 59 46.5
-29 13 44.9
-29 3 59.0
-29 1 43.4
-29 10 20.3
-29 5 57.3
-29 15 24.7
-29 20 42.5
-29 4 6.5
-29 9 10.8
-29 16 4.6
-29 11 15.4
-29 1 14.9
-29 2 50.0
-29 1 24.9
-29 5 48.5
-29 11 46.8
-29 1 35.8
-29 3 34.2
-29 5 17.8
-29 7 12.2
-29 4 25.2
-29 5 26.7
-29 15 27.4
-29 7 20.8
J
14.57
15.14
13.80
--
13.41
15.20
--
--
--
15.84
15.24
15.62
11.08
15.08
14.58
--
--
--
14.84
--
14.34
15.77
15.25
--
--
14.64
--
--
13.43
--
13.83
13.94
13.69
15.24
13.89
13.94
--
15.59
--
13.17
15.93
--
KS
10.03
9.23
8.50
9.58
8.04
8.68
9.67
10.99
11.18
9.26
9.34
9.89
6.17
9.42
9.01
9.61
9.30
10.78
8.58
10.70
8.53
9.05
10.11
8.84
10.65
9.15
8.19
10.78
7.95
9.43
8.23
8.52
7.85
9.01
8.46
8.56
10.87
9.61
8.72
7.62
10.31
10.60
[7]
8.74
4.77
5.08
5.58
4.05
4.83
4.20
4.50
3.57
5.90
5.03
8.16
3.41
6.92
6.92
7.62
5.38
8.76
4.89
5.13
5.48
6.11
4.48
5.30
4.66
7.14
4.73
4.08
5.50
5.90
4.91
5.52
4.86
4.27
5.47
4.86
7.82
7.69
4.85
4.56
3.47
4.14
[15]
6.43
3.93
2.92
4.04
1.76
3.15
2.62
2.35
2.05
4.35
2.85
99.99
1.31
5.21
5.34
5.13
3.73
6.82
3.23
3.29
3.91
4.41
1.65
3.51
2.45
5.46
2.99
0.50
3.75
4.27
3.26
3.86
3.31
1.97
3.75
2.73
6.26
4.82
3.28
2.98
1.59
2.40
AV
19.60
19.10
19.70
24.50
21.60
25.10
22.30
25.00
20.50
19.40
20.50
21.50
25.40
21.60
24.80
24.50
28.50
19.80
24.50
22.10
19.80
25.30
25.30
20.30
23.40
22.30
24.40
20.30
23.80
19.80
18.50
23.10
23.10
23.50
20.60
24.80
22.80
22.80
22.30
23.90
24.00
22.90
12CO(2, 0)
19.48
9.83
21.04
16.88
21.93
13.75
20.74
19.09
12.99
8.38
20.94
21.04
20.58
29.17
19.80
20.76
18.84
18.94
22.34
7.52
20.48
19.99
19.20
17.13
8.98
23.18
17.21
-2.28
22.46
12.08
20.32
7.78
22.72
-1.61
17.11
18.78
17.07
18.75
21.14
16.59
15.75
8.78
NaI
4.93
2.44
2.26
3.94
4.10
2.17
4.88
5.25
1.29
2.30
4.57
6.26
0.78
6.16
5.21
5.47
4.21
5.42
3.70
0.19
3.54
3.76
4.29
1.49
-0.05
5.60
1.34
0.42
3.96
1.27
2.64
0.65
3.34
0.58
2.32
-0.14
3.82
4.46
3.20
1.52
3.32
1.01
CaI
1.87
-0.50
-0.74
0.23
-0.11
-0.51
2.41
3.61
-2.11
-1.37
1.94
1.35
-0.24
1.31
1.41
3.40
1.63
2.91
--
-0.95
-0.76
-1.08
3.11
-1.44
-1.75
2.32
-0.77
-1.24
-1.06
--
-0.16
0.69
0.20
-0.18
0.85
0.36
1.12
2.02
-0.38
-0.06
2.54
0.12
H2O
-35.78
-261.11
-666.39
-157.16
-386.72
-259.69
-23.84
-13.92
-181.29
-434.36
-120.23
-54.97
-52.70
-336.94
-184.69
-27.37
-144.06
-19.26
-543.70
-277.07
-623.25
-526.70
-8.01
-473.84
129.52
-112.56
-332.12
Mbol
-3.25
-4.32
-4.83
-4.52
-5.71
-5.04
-5.08
-4.45
-6.03
-3.74
-4.36
-3.14
-8.06
-3.76
-4.65
-4.01
-4.98
-2.48
-5.08
-3.80
-4.52
-4.36
-5.07
-4.67
-4.32
-4.13
-5.63
1.78
2504.72
-505.14
-430.36
-694.69
-498.84
-347.86
77.81
-536.87
-993.36
-86.09
-52.85
-643.49
-629.52
-54.14
-318.86
-5.64
-4.18
-4.83
-5.06
-5.56
-5.14
-4.77
-5.46
-2.82
-3.61
-5.19
-6.03
2504.72
-4.94
log M
-6.69
-5.46
-5.36
-5.50
-5.17
-5.52
-4.99
-4.75
-4.65
-5.64
-5.10
--
-5.89
-6.20
-7.06
-6.13
-5.60
-6.32
-5.59
-4.92
-5.82
-6.04
-4.78
-5.47
-4.80
-6.80
-5.66
--
-6.33
-5.53
-5.58
-5.93
-6.01
--
-5.79
-5.39
-5.95
-5.83
-5.48
-5.98
-4.73
-4.79
type
RGB
AGB
AGB
AGB
OH/IR
AGB
AGB
OH/IR
OH/IR
AGB
OH/IR
RGB
supergiant
LPV
AGB
AGB
AGB
RGB
OH/IR
AGB
AGB
LPV
AGB
AGB
OH/IR
AGB
LPV
YSO
LPV
LPV
OH/IR
OH/IR
LPV
YSO
LPV
OH/IR
RGB
AGB
LPV
LPV
OH/IR
OH/IR
M
.
S
c
h
u
l
t
h
e
i
s
e
t
a
l
.
:
N
e
a
r
-
I
R
s
p
e
c
t
r
o
s
c
o
p
y
o
f
I
S
O
G
A
L
s
o
u
r
c
e
s
1
3
A54
A55
A56
A57
A58
A59
A60
A61
A62
A64
A66
A67
A68
A69
A70
A71
A72
A74
A75
A76
A77
A78
A79
B7
B13
B14
B15
B17
B18
B19
B20
B22
B23
B27
B28
B31
B32
B35
B37
B40
C3
C6
C10
C14
17 45 16.7
17 45 17.5
17 45 18.0
17 45 18.8
17 45 19.4
17 45 20.5
17 45 20.9
17 45 21.5
17 45 21.9
17 45 26.4
17 45 27.5
17 45 28.4
17 45 28.5
17 45 29.5
17 45 30.5
17 45 34.4
17 45 36.0
17 45 38.2
17 45 40.0
17 45 40.7
17 45 40.7
17 45 41.4
17 45 43.4
17 43 48.6
17 44 4.7
17 44 7.5
17 44 8.1
17 44 11.5
17 44 12.7
17 44 12.8
17 44 16.5
17 44 17.6
17 44 17.8
17 44 21.7
17 44 22.8
17 44 27.4
17 44 27.5
17 44 29.9
17 44 33.8
17 44 44.5
17 42 58.8
17 43 3.4
17 43 11.2
17 43 14.7
-29 7 34.3
-29 10 18.9
-29 8 42.9
-29 5 5.4
-29 14 5.9
-29 7 19.5
-29 13 42.4
-29 6 36.5
-29 13 44.3
-29 8 4.2
-29 4 39.9
-29 4 22.9
-29 18 5.0
-29 9 16.6
-29 15 50.9
-29 12 54.2
-29 4 38.1
-29 17 20.1
-29 16 2.8
-29 4 27.7
-29 14 54.7
-29 13 43.6
-29 13 29.1
-29 27 18.8
-29 25 18.7
-29 27 37.6
-29 32 22.5
-29 31 31.1
-29 31 24.6
-29 26 55.5
-29 26 7.7
-29 25 58.9
-29 27 13.0
-29 27 36.2
-29 34 55.0
-29 38 5.8
-29 28 52.8
-29 28 57.3
-29 23 57.8
-29 31 36.8
-29 49 59.9
-29 38 1.5
-29 51 29.8
-29 37 49.9
--
14.18
--
--
--
--
--
--
13.75
--
16.14
15.60
--
--
14.76
--
--
13.69
14.22
15.27
--
15.99
15.16
--
--
13.94
13.99
15.16
15.91
13.22
--
--
14.07
--
--
15.80
--
--
15.80
--
15.57
--
13.31
15.32
10.80
9.19
9.39
10.05
8.57
10.54
8.80
10.13
8.62
8.67
9.97
9.27
10.30
10.35
9.54
9.90
10.36
9.05
8.11
9.59
10.12
10.08
10.42
11.54
10.51
6.64
6.86
9.30
9.30
6.72
10.56
10.99
7.80
10.35
10.46
8.70
11.27
10.38
9.75
10.41
9.60
9.36
7.14
9.26
8.56
7.19
6.79
7.69
4.08
6.29
6.68
6.59
6.61
4.63
7.91
6.88
8.35
4.87
7.16
4.30
7.42
6.79
4.62
7.76
4.31
8.36
9.18
8.16
6.58
2.40
2.67
5.41
6.98
3.02
8.37
99.99
3.99
6.27
7.36
4.47
8.14
7.84
6.73
8.00
5.91
4.02
4.52
6.62
Table A1 (cont.)
6.77
5.59
5.22
6.08
2.02
4.29
5.18
4.99
4.87
2.69
6.06
3.90
99.99
3.06
5.65
2.71
4.94
4.80
2.82
6.17
2.40
6.93
7.26
99.99
2.91
0.59
0.85
3.20
5.25
0.98
6.49
6.34
1.28
2.61
5.47
2.34
6.11
5.47
4.64
5.23
3.59
1.58
1.49
4.49
22.90
23.50
22.80
25.90
21.00
23.90
23.90
24.00
21.90
23.60
24.50
25.00
19.30
22.50
24.50
24.50
24.80
20.40
22.50
24.10
23.20
20.50
21.30
27.20
28.20
28.60
28.50
28.50
29.40
29.00
29.00
28.80
29.00
29.60
27.00
27.30
34.90
32.30
29.80
27.00
18.40
22.00
20.30
22.80
15.51
22.39
21.33
20.37
18.28
20.85
19.69
18.46
23.40
19.96
22.98
22.45
19.83
7.59
21.55
14.15
20.18
20.94
11.02
20.53
14.51
23.56
16.47
11.25
17.49
23.84
23.36
19.93
23.61
27.10
21.52
17.41
-1.21
0.92
24.17
13.17
23.83
-0.75
-0.73
22.41
25.42
13.47
18.34
19.37
2.97
6.42
5.60
4.82
2.32
3.16
5.00
2.97
5.88
5.88
5.62
5.22
4.46
0.23
6.38
1.43
5.78
5.41
2.48
5.19
2.26
5.46
2.48
0.95
4.07
4.72
4.88
5.49
4.07
3.79
3.85
3.79
-1.06
1.26
5.40
1.78
5.62
0.09
0.69
5.60
5.21
0.84
0.49
2.59
4.24
2.94
2.88
3.82
-0.75
-0.45
1.04
-0.32
4.14
1.98
3.06
2.80
4.13
-0.91
2.22
-0.63
2.50
2.68
0.32
2.98
0.94
3.39
2.19
1.69
4.14
3.82
3.35
2.68
5.65
3.68
4.81
4.86
-0.15
0.52
5.22
2.24
4.40
0.76
1.10
4.47
3.22
0.23
0.94
1.37
-54.35
-173.11
-89.14
-27.65
-806.41
-252.72
-276.51
-573.02
-129.72
-416.59
-88.49
-171.28
-16.33
236.30
-127.51
-620.11
-129.24
-95.58
-341.28
-80.12
-52.42
-88.19
-78.17
-70.05
-48.80
-36.04
9.05
-115.19
-229.09
-45.72
-57.55
-3.47
-0.34
-2.38
-97.87
-317.56
37.23
-2.22
14.83
0.04
-279.57
-541.95
-744.63
-529.80
-2.75
-4.56
-4.19
-3.71
-5.25
-3.37
-4.80
-3.77
-4.77
-5.17
-3.39
-4.21
2504.72
-4.10
-4.29
-5.03
-3.33
-4.38
-5.30
-3.88
-4.62
-2.76
-3.05
2504.72
-3.75
-7.41
-7.19
-5.10
-4.64
-7.43
-3.50
2504.72
-6.53
-4.06
-3.49
-5.31
-3.33
-3.94
-4.62
-3.41
-3.63
-5.11
-5.92
-4.00
-6.44
-7.04
-6.30
-6.73
-5.09
-5.23
-7.05
-5.68
-6.69
-5.31
-6.70
-5.59
-3.70
-4.95
-6.72
-4.99
-5.56
-6.10
-5.55
-7.42
-4.88
-7.50
-7.58
-3.70
-4.96
-5.40
-5.41
-5.38
-6.93
-5.54
-6.86
-6.21
--
--
-5.87
-5.26
-6.11
--
--
-5.76
-5.19
-4.85
-5.35
-5.84
RGB
AGB
AGB
AGB
OH/IR
LPV
AGB
LPV
AGB
OH/IR
RGB
AGB
AGB
LPV
AGB
OH/IR
AGB
AGB
AGB
AGB
AGB
RGB
RGB
AGB
AGB
supergiant
supergiant
AGB
AGB
supergiant
RGB
AGB
YSO
YSO
RGB
AGB
RGB
YSO
YSO
RGB
AGB
AGB
AGB
AGB
1
4
M
.
S
c
h
u
l
t
h
e
i
s
e
t
a
l
.
:
N
e
a
r
-
I
R
s
p
e
c
t
r
o
s
c
o
p
y
o
f
I
S
O
G
A
L
s
o
u
r
c
e
s
C19
C23
C25
C45
D5
D6
D7
D11
D13
D22
E4
E5
E13
E20
E51
F13
F14
F23
G35
G46
H01
17 43 19.3
17 43 24.6
17 43 25.0
17 43 44.2
17 42 27.3
17 42 28.0
17 42 29.9
17 42 44.4
17 42 47.7
17 43 29.8
17 43 13.0
17 43 15.5
17 43 28.2
17 43 39.2
17 44 35.0
17 42 28.0
17 42 39.5
17 40 36.3
17 43 16.4
17 44 7.5
17 30 53.2
-29 50 28.9
-29 53 12.4
-29 41 23.0
-29 38 26.1
-29 54 17.8
-29 56 14.6
-30 1 15.9
-29 58 36.1
-29 56 25.5
-30 1 27.3
-29 21 1.4
-29 24 58.7
-29 17 41.7
-29 22 44.5
-29 4 35.5
-29 37 46.4
-29 43 27.3
-29 49 13.7
-30 13 10.7
-30 7 41.3
-33 40 21.7
13.92
14.33
--
15.61
13.28
13.90
--
14.03
--
12.17
--
11.51
11.33
--
--
15.21
15.36
11.51
14.14
14.09
15.72
10.98
8.56
9.32
9.75
8.04
10.55
10.97
7.94
9.56
7.89
7.22
6.58
6.93
10.61
9.09
9.87
9.91
8.18
10.98
8.63
10.53
6.53
6.89
5.37
7.50
6.50
3.87
3.36
4.60
6.74
5.73
5.32
4.87
5.21
6.61
1.30
5.23
4.95
5.71
4.95
5.32
5.53
Table A1 (cont.):
3.80
4.82
3.19
4.33
4.44
0.13
-0.06
2.39
4.40
3.59
4.33
4.07
4.67
4.24
-0.39
3.01
3.13
2.21
2.91
3.46
2.92
22.50
23.60
25.00
23.80
20.30
24.30
22.40
26.00
27.90
19.00
23.30
25.70
22.70
25.50
20.10
17.00
21.50
9.80
16.30
20.60
10.80
-3.64
25.90
16.35
25.52
13.49
-1.00
0.87
12.16
22.36
21.64
20.36
23.21
20.18
19.69
10.11
20.61
18.15
10.91
19.06
18.51
13.11
-0.58
5.16
3.10
5.19
3.05
0.03
1.20
1.85
5.19
1.69
4.41
5.36
4.05
3.39
1.60
2.00
2.57
0.91
1.77
3.51
2.50
0.12
4.89
3.17
3.28
2.79
1.44
1.76
0.66
4.34
2.07
2.70
3.04
2.13
1.65
0.61
1.12
3.08
2.72
0.82
1.45
1.43
-50.59
-237.00
-223.55
-279.99
-364.41
-57.42
-82.31
-544.74
-115.25
-1020.49
-221.63
-109.75
-193.37
-162.79
-297.85
-763.02
-186.75
-597.37
-340.69
-506.89
-492.72
-4.24
-4.79
-4.57
-3.72
-5.02
2504.72
2504.72
-5.89
-4.40
-5.49
-6.38
-7.64
-7.05
-3.31
-8.22
-3.87
-4.24
-4.49
-3.94
-4.69
-3.36
--
-6.84
-5.29
-5.53
-6.75
--
--
-5.54
-5.80
-5.99
-8.66
-11.67
-12.24
-5.22
-4.61
-4.96
-5.04
-5.06
-4.75
-5.55
-4.77
YSO
AGB
AGB
AGB
AGB
YSO
YSO
AGB
AGB
AGB
AGB
AGB
AGB
AGB
OH/IR
AGB
AGB
AGB
AGB
AGB
AGB
M
.
S
c
h
u
l
t
h
e
i
s
e
t
a
l
.
:
N
e
a
r
-
I
R
s
p
e
c
t
r
o
s
c
o
p
y
o
f
I
S
O
G
A
L
s
o
u
r
c
e
s
1
5
16
M. Schultheis et al.: Near-IR spectroscopy of ISOGAL sources
Fig. B.1. Young stellar objects
Fig. B.2. Supergiants candidates
M. Schultheis et al.: Near-IR spectroscopy of ISOGAL sources
17
Fig. B.3. OH/IR stars (Wood et al. 1998)
18
M. Schultheis et al.: Near-IR spectroscopy of ISOGAL sources
Fig. B.4. Long period Variables (Glass et al. 2001)
M. Schultheis et al.: Near-IR spectroscopy of ISOGAL sources
19
Fig. B.5. AGB star candidates
20
M. Schultheis et al.: Near-IR spectroscopy of ISOGAL sources
Fig. B.5. (continued)
Fig. B.5. (continued)
M. Schultheis et al.: Near-IR spectroscopy of ISOGAL sources
21
22
M. Schultheis et al.: Near-IR spectroscopy of ISOGAL sources
Fig. B.5. (continued)
M. Schultheis et al.: Near-IR spectroscopy of ISOGAL sources
23
Fig. B.6. Red giant candidates
|
astro-ph/0702165 | 1 | 0702 | 2007-02-06T17:48:04 | Potential Magnetic Field around a Helical Flux-rope Current Structure in the Solar Corona | [
"astro-ph"
] | We consider the potential magnetic field associated with a helical electric line current flow, idealizing the near-potential coronal field within which a highly localized twisted current structure is embedded. It is found that this field has a significant axial component off the helical magnetic axis where there is no current flow, such that the flux winds around the axis. The helical line current field, in including the effects of flux rope writhe, is therefore more topologically complex than straight line and ring current fields sometimes used in solar flux rope models. The axial flux in magnetic fields around confined current structures may be affected by the writhe of these current structures such that the field twists preferentially with the same handedness as the writhe. This property of fields around confined current structures with writhe may be relevant to classes of coronal magnetic flux rope, including structures observed to have sigmoidal forms in soft X-rays and prominence magnetic fields. For example, ``bald patches'' and the associated heating by Parker current sheet dissipation seem likely. Thus some measurements of flux rope magnetic helicities may derive from external, near-potential fields. The predicted hemispheric preference for positive and negative magnetic helicities is consistent with observational results for prominences and sigmoids and past theoretical results for flux rope internal fields. | astro-ph | astro-ph | Potential Magnetic Field around a Helical Flux-rope Current
Structure in the Solar Corona
G.J.D. Petrie
National Solar Observatory, 950 N. Cherry Avenue, Tucson, AZ 85719; [email protected]
7
0
0
2
b
e
F
6
1
v
5
6
1
2
0
7
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
ABSTRACT
We consider the potential magnetic field associated with a helical electric line
current flow, idealizing the near-potential coronal field within which a highly
localized twisted current structure is embedded. It is found that this field has
a significant axial component off the helical magnetic axis where there is no
current flow, such that the flux winds around the axis. The helical line current
field, in including the effects of flux rope writhe, is therefore more topologically
complex than straight line and ring current fields sometimes used in solar flux
rope models. The axial flux in magnetic fields around confined current structures
may be affected by the writhe of these current structures such that the field
twists preferentially with the same handedness as the writhe. This property of
fields around confined current structures with writhe may be relevant to classes of
coronal magnetic flux rope, including structures observed to have sigmoidal forms
in soft X-rays and prominence magnetic fields. For example, "bald patches" and
the associated heating by Parker current sheet dissipation seem likely. Thus some
measurements of flux rope magnetic helicities may derive from external, near-
potential fields. The predicted hemispheric preference for positive and negative
magnetic helicities is consistent with observational results for prominences and
sigmoids and past theoretical results for flux rope internal fields.
Subject headings: magnetohydrodynamics: Sun, solar magnetic fields, solar corona
1.
Introduction
Magnetic fields emerge into the solar atmosphere with significant twist (Leka et al. (1996).
Solar coronal eruptions are believed to be caused by the ultimate failure of the confinement
of highly twisted magnetic fields (Low 1996, 2001). The new twisted flux does not immedi-
ately mix with the flux already there, but maintains its identity. This is because the nearly
perfectly conducting plasma and magnetic flux are frozen together except for field recon-
nection which occurs only on very small scales. Old and new flux systems find new mutual
-- 2 --
equilibrium states via successive reconnections and dynamical relaxations. Taylor's (1986)
theory has described turbulent relaxation to a minimum energy state, a linear force-free field,
under the postulated conservation of magnetic helicity contained in the plasma, a postulate
shown to be relevant to coronal plasma by Berger (1984). In the open environment of the
solar atmosphere, the dispersal of magnetic helicity occurs in an unbounded volume so that
the ultimate minimum energy state in the absence of further helicity injection is a potential
field. Well-confined structures may maintain their integrity within a near-potential ambient
field for long periods.
Examples of confined current structures embedded in relatively current-free surround-
ings are flux ropes associated with prominences and x-ray sigmoids. Observations of flux
ropes have repeatedly yielded evidence of helical structure (Rust & Kumar 1996, Chen et
al. 1997, 2000, Dere et al. 1999, Wood et al. 1999, Ciaravella et al. 2000, Gary & Moore 2004).
Observations (Leroy et al. 1984, 1989, Bommier et al. 1994, Lin et al. 1998) have con-
sistently emphasized the presence of strong axial components of prominence magnetic fields.
Amari & Aly (1989, 1992) modeled prominences by embedding line currents and current
sheets in twisted force-free magnetic fields. Employing line-current and current sheet solu-
tions, Low & Hundhausen (1995) demonstrated the necessity of spatially distributed currents,
associated with axial magnetic flux, for prominence support. Here the mathematical prob-
lem can be simplified by assuming that the prominence plasma is cold enough for the width
of the plasma condensation to be negligibly thin relative to the scale of the global field con-
figuration. This is reasonable since the hydrostatic scale height at prominence temperatures,
a few hundred km, is much smaller than the characteristic length of the large-scale magnetic
field, about 105 km.
On the other hand, sigmoids have been identified with plasma structures heated by
Parker current sheet dissipation (Titov & D´emoulin 1999, Low & Berger 2003). Flux sur-
faces consisting of field trajectories which graze the photosphere, known as bald patches
(Titov & D´emoulin 1999) are likely locations for the formation of magnetic tangential dis-
continuities between twisted coronal field and the photosphere. The two observed sigmoidal
shapes, the S-shape and its mirror-reflection, the Z-shape, are each found in both hemi-
spheres of the solar corona, but preferentially in the northern and southern hemispheres
respectively. Associating Z- and S-shaped sigmoids with flux ropes of negative and positive
magnetic helicities respectively, Low & Berger (2003) found a connection between the writhe
that the axial flux in fields around twisted current structures is biased by the writhe of these
current structures. This bias is such that the observed left-handed and right-handed twists
of flux ropes in the northern and southern hemispheres may be connected to the hemispheric
preferences for the two sigmoidal shapes via the fields around these flux ropes. Observations
-- 3 --
of both prominence and sigmoidal flux ropes indicate a hemispheric preference for the hand-
edness of twist of these structures (Canfield & Pevtsov 1999, Chae 2000, Pevtsov, Canfield
& Latushko 2001).
Titov & D´emoulin (1999) modeled a flux rope using a ring current, with current dis-
tributed across a ring of small but finite radius. An axial field component was produced by
threading a line current through the center of this ring. The field outside the ring is poten-
tial. Low & Berger (2003) constructed their flux rope using helically symmetric magneto-
hydrostatic solutions embedded within an ambient field described by Gold & Hoyle's (1960)
linear force-free field. In this paper we shift focus to the field around a localized twisted
current-carrying structure: using the potential field around a helical line current we study
the relatively unstressed field within which a flux rope with twisted axis (writhe) is embed-
ded.
2. Magnetohydrostatics and the helical line current
Consider the static equilibrium of a the magnetic field, denoted by B, in the absence of
significant plasma forces. The balance of forces is described by
j × B − ∇p − ρgz = 0,
1
4π
(∇ × B) = j,
∇ · B = 0,
(1)
(2)
(3)
Wherever the first term on the left-hand side of Equation (1) is much larger than the
other two terms, as is often the case for the solar corona, the force-free approximation applies:
Here α is a function of space which, by Equation (3), must be constant along field lines,
∇ × B = αB.
B · ∇α = 0.
(4)
(5)
Further sub-cases of interest for the study of the corona are the linear force-free case, in which
α = α0 is constant and equation (4) becomes linear, and the potential-field (current-free)
subcase where this constant α0 = 0. This paper focuses on a special magnetic field which
-- 4 --
falls into the category of potential field everywhere except a single helical line along which
an electric current of infinitesimally thin width flows. This line current represents a highly
twisted flux rope within which all forces of equation (1) may be important but outside which
the field has relaxed to a near-potential state. Such a field does not explicitly include forces
and the line current may be regarded as the limiting case of a class of nonlinear force-free
fields whose currents are concentrated along helical axes.
Using Cartesian coordinates (x, y, z) with y in the vertical direction, we define the
half-space y > 0 to be the corona. Defining the cylindrical coordinates with translational
symmetry in the z-direction,
x = r cos ϕ, y = r sin ϕ.
(6)
Coronal magnetic flux ropes usually lie along relatively straight polarity inversion lines (PILs)
at the base of the corona, and may be idealized to be locally 2D with invariance in the
direction of the PIL. Here we are more interested in the global 3D geometry of the ambient
field than with the central structure.We define the helical coordinate
ζ = ϕ − kz.
(7)
Lines of constant ζ wind around the cylindrical axis of symmetry x = y = 0 a full turn over
a distance 2π/k units along that axis.
In a Cartesian 2D field the magnetic surfaces are determined by the conservation of
longitudinal magnetic flux and in an axisymmetric field the azimuthal flux is conserved. The
flux surfaces of a helical field are determined by the conservation of the linear combination
ψ =
1
2π
(Ψz − Ψϕ),
of ϕ- and z-directed fluxes
Ψz =Z Bz(r, ζ)r dr dz, Ψϕ =Z Bϕ(r, ζ) dr dz =
1
k Z Bϕ(r, ζ) dr dϕ.
We define this flux function ψ(r, ζ) as follows:
rBr =
∂ψ
∂ζ
,
(8)
(9)
(10)
-- 5 --
Bϕ − krBz = −
∂ψ
∂r
.
(11)
This solves Equation (3), fixing the field components in the r and ∇ζ directions but leaving
the component in the r × ∇ζ direction free. From the ϕ- and z-components of Equation (1)
this component Bz + krBϕ is a free function F (ψ) of the flux function ψ (Neukirch 1999).
Therefore the magnetic field takes the form (Low & Berger 2003)
where
B =
1
r
∂ψ
∂ζ
r +
1
1 + k2r2
∂ψ
∂r
a +
F (ψ)
1 + k2r2
b,
a = ϕ − krz, b = kr ϕ + z,
(12)
(13)
and the equilibrium of Equation (1) is described in this case by the Grad-Shafranov equation
(Freidberg 1980, Low & Berger 2003),
1
r
∂
∂r (cid:18)
1
1 + k2r2
∂ψ
∂r(cid:19) +
1
r2
∂2ψ
∂ζ 2 −
2kF
(1 + k2r2)2 +
1
1 + k2r2 F
dF
dψ
+ 4π
dp
dψ
= 0.
(14)
This equation ignores gravity because helical symmetry and gravitational stratification are
mutually exclusive. Here p must be a function of the flux function ψ (Freidberg 1987,
Neukirch 1999). The force-free case is p ≡ 0 and the linear force-free and potential cases by
F = α0ψ and F ≡ 0.
In the present work we are ignoring plasma forces. Since we are modeling our magnetic
filament using a line current, our model does not resolve the physics of this current.
In
real systems the current will have a field-aligned component not associated with a Lorentz
force and a further component associated with such a force. Here, these details are collapsed
into a helical line of infinitesimally small width, within which all currents are confined. The
linear force-free form of this Equation (14) has eigenfunctions composed of Bessel functions
of r and cosines of ζ. Gold & Hoyle's (1960) force-free flux rope solution, whose field lines
are confined to lines of constant ζ, is the degenerate case ψ = constant with F a non-zero
constant.
To find the magnetic field of a single helical line current, we follow Morozov & Solov'ev (1966)
in first representing the potential fields B = ∇φi inside and B = ∇φe outside the cylinder
r = a associated with current flowing in the surface of the cylinder with components
-- 6 --
j = inka sin nζ ϕ + in cos nζ z
φ(r, ζ) =(cid:26) φi(r, ζ),
φe(r, ζ),
for r < a,
for r > a,
with the potential
where
φi = ai
nIn(nkr) sin nζ, φe = ae
nIn(nkr) sin nζ.
(15)
(16)
(17)
The boundary conditions at the boundary of the cylinder, derived from Equation (2), are
Bi
r − Be
r = 0, Bi
ϕ − Be
ϕ = −
4π
c
jz, Bi
z − Be
z =
4π
c
jϕ.
Using the identity I ′
n(x)Kn(x) − K ′
n(x)In(x) = 1/x the coefficients are found to be
ai
n =
4πka2in
c
K
′
n(nka), ae
n =
4πka2in
c
I
′
n(nka).
(18)
(19)
Now a single helical line current of strength I on the cylindrical surface is described by
jz =
I
k
δ(ζ − ζ0) =
jϕ = Ikδ(ζ − ζ0) =
I
2πk 1 + 2
2π 1 + 2
Ik
cos n(ζ − ζ0)! ,
cos n(ζ − ζ0)! ,
∞
∞
Xn=1
Xn=1
The zeroth harmonics are the constant fields
Bi
0 =
2Ik
c
z, Be
0 =
2I
cr
ϕ,
(20)
(21)
(22)
where the superscripts indicate internal and external solutions as usual, and the full po-
tentials for the interior and exterior fields are found by summing the other harmonics of
Equation (17),
-- 7 --
φi =
φe =
K
∞
2Ik
c z +
Xn=1
c ϕ + 2k2
Xn=1
∞
2I
′
n(nka)In(nkr) sin n(ζ − ζ0)! ,
n(nka)Kn(nkr) sin n(ζ − ζ0)! .
′
I
Therefore the magnetic field is given by
∞
B(r, ζ) = 4k2a
Xn=1
+ 4ka
Xn=1
r
∞
+2kI z,
nK
′
nK
n(nka)I
′
n(nkr) sin n(ζ − ζ0)! r
n(nka)In(nkr) cos n(ζ − ζ0)! a
′
inside the cylinder and
∞
B(r, ζ) = 4k2a
Xn=1
+ 4ka
Xn=1
r
∞
+2kI ϕ,
′
nI
n(nka)K
′
n(nkr) sin n(ζ − ζ0)! r
n(nka)Kn(nkr) cos n(ζ − ζ0)! a
′
nI
(23)
(24)
(25)
(26)
outside the cylinder.
The flux function ψ(r, ζ) is found in practice from the vector potential A, such that
B = ∇ × A: ψ = krAϕ + Az. This vector potential has been calculated by Tominaka (2004)
and the results are given in the Appendix. Sample contour plots of both the potential
phi(r, ζ) and the flux function psi(r, ζ) are shown in Figure 1 for selected values of helical
winding number k. Unlike in the 2D Cartesian and axisymmetric cases, the helical flux
function does not serve as a matched Euler potential for the magnetic field with the ignorable
coordinate, i.e., the vector ∇ψ × ∇ζ. Therefore the flux function ψ and the scalar potential
φ do not have contour curves perpendicular to each other everywhere in the r-ζ plane, as
Figure 1 clearly shows. The flux contours are squeezed in the region between the helical
axis and the axis r = 0 because of their relative proximity to the rest of the current and
-- 8 --
intense magnetic field in the configuration. This effect increases as k increases until a local
extremum of the flux function develops close to r = 0. As Morozov & Solov'ev (1966) report,
the presence of flux function extrema in regions where current flow does not occur. This
contrasts with the 2D Cartesian case where the maximum principle for Laplace's equation
forbids closed flux contours in current-free regions. These second extrema may be regarded
as low-lying loops beneath the flux rope.
There is a further departure from the simpler 2D Cartesian and axisymmetric cases.
The quantity
F = Bz + krBϕ = 2kI,
(27)
is non-zero, which contrasts with Equation (12) for the potential case F ≡ 0. The non-zero
F is entirely due to the zeroth harmonics of Equation (22), which are necessary for a correct
description of an infinitesimally thin helical tube of current. This field has a strong axial
(b-directed) component, unlike the simpler straight line current and circular ring current
solutions (Jackson 1965, Low & Hundhausen 1995). In constructing their circular current
structure, Titov & D´emoulin (1999) found it necessary to thread a straight line current
through their current ring in order to produce an axial magnetic field component. In contrast,
our helical line current naturally produces a significant axial field component because of the
skewed curvature of the current trajectory. Figure 2 shows sample field lines for the case
k = 1. The presence of axially-directed magnetic flux is clear. The magnetic field structure
has the familiar appearance of a twisted flux rope confined within a relatively unsheared
arcade.
The existence of strong axial field components has been a strikingly consistent feature
of prominence observations (Leroy et al. 1984, 1989, Bommier et al. 1994, Lin et al. 1998).
Low & Hundhausen (1995) proved that spatially distributed currents, associated with axially-
directed magnetic flux, are a necessary condition for prominence support. Here we are less
concerned with the weight support problem than with relating the axial flux around the flux
rope to the global structure of the rope, in particular the writhe. While these internal and
external axial fluxes play separate roles in the physics of the prominence flux rope, they may
be coupled in some way. In real systems it is likely that spatially distributed currents of
complex structure exist in the vicinity of the flux rope concentration itself. This flux could
flow in either axial direction producing flux twisting in either direction about the helical
axis. However, the fact that the field surrounding a flux rope with writhe has axial flux even
in the absence of local currents implies a preferred direction for axial flux determined by
the writhe of the rope. Thus we expect a preference for right-handed and left-handed twist
around right-handed and left-handed flux ropes respectively.
-- 9 --
Fig. 1. -- Magnetic potential contours (left pictures) and magnetic flux function contours
(right pictures) for the cases k = 0.5 (top), k = 1.0 (middle) and k = 2.0 (bottom).
-- 10 --
The axial flux of the helical line current potential field may also have a bearing on the
hemispheric preference of S-shaped sigmoids and their mirror reflections, which we call Z-
shaped after Low & Berger (2003). Z- and S-shaped sigmoids are associated with flux ropes
with positive and negative magnetic helicities respectively. There is no unique relationship
between magnetic Z and S shapes on the one hand and the sign of the magnetic helicity on
the other because the signs of the writhe of a flux rope and the internal magnetic helicity of
the field inside the rope are distinct. Moreover, a flux rope of a fixed helicity can have some
field trajectories project into S shapes and others into Z shapes.
The twist of the line current in the example in Figure 2 is right-handed. The twist of the
field trajectories around this line current is also right-handed. This line current is parallel to
a Gold-Hoyle magnetic field trajectory. If the shapes of this line current trajectory projected
on the atmospheric base y = 0 is identified with the shapes of observed sigmoids, this
solution implies an association of positive magnetic helicity, that is, a right-handed wind of
lines of force, with Z sigmoids. If we reverse the sign of the entire field, the field preserves
its handedness and sign of magnetic helicity because magnetic helicity and handedness are
independent of the direction of the magnetic field. If we reverse the sign of the parameter k
we obtain the mirror reflection of the solution in Figure 2. In the mirror-reflected solution,
the field now has negative helicity and a left-handed wind in its fields.
Its line current
projected onto y = 0 would then be shaped like an S sigmoid.
A direct association of the handedness of the line current twist to sigmoidal shape
identifies sigmoidal shapes with the writhes of flux ropes. This is equivalent to Low &
Berger's association of their Gold-Hoyle external fields with sigmoidal patterns since our line
currents are parallel to Gold-Hoyle fields. If the observed coronal sigmoids are interpreted
in terms of these associations, the implication is that positive and negative helicities are
preferentially found in the northern and southern hemispheres. As Low & Berger explain,
this contradicts the hemispherical preference effect for prominence and sigmoid flux ropes.
Low & Berger resolved this inconsistency by demonstrating that field trajectories within
their flux rope winding around a right-handed axis of helical symmetry with right-handed
twist (γ = α0/k > 0) grazing the atmospheric base project onto the atmospheric base as
S-shapes. Left-twisted lines about a right-twisted axis do not do this so readily. Anchored
field lines rising from the photosphere may wind more than once in the atmosphere before
exiting the atmosphere through their second footpoints. Such winding lines may tangentially
touch the atmospheric base from above only to return into the atmosphere. Low & Berger
demonstrated that the heated plasma in a sigmoid might arise from current sheets forming
within a flux rope as opposed to deriving from an external field, and that the sigmoidal
shape may be related more to the topology of the flux rope internal field than the topology
-- 11 --
of the external field. We have seen that the sigmoidal shapes that a Gold-Hoyle field would
produce would not conform to the hemispherical preference of sigmoidal shapes described
above. In contrast, a field within a helically symmetric flux rope of right-handed writhe,
with right-handed twist about the helical axis is likely to include lines winding about the
axis more than once, grazing the photosphere on the way. Such lines tend to project as
S-shapes.
On the other hand, the Gold-Hoyle field has linear field-aligned currents and does not
closely approximate an ambient coronal field in a relaxed, near-potential equilibrium. The
magnetic flux trajectories associated with this helical line-current wind more than a full turn
about the rope axis in the atmosphere whereas the Gold-Hoyle field is directed parallel to
it. For example, there is a particular field line in Figure 2 which dips very close to the
atmospheric base, almost grazing it just above the origin at the center of the domain. This
line projects as an S-shape and would show up as a bright heated structure.
Such a relationship between sigmoidal shapes and magnetic helicities invokes certain
physical properties of the heated plasma. The force-free field of the domain where our
model is meaningful, y > 0, may be regarded to match discontinuously a distinct field under
different physical conditions in y < 0, since the atmospheric layers between the corona and
the photosphere are comparatively very thin. The flux surface containing all fields grazing
the photosphere is tangential to the boundary y = 0 between a tenuous atmosphere and
a dense region below. Even if the field varies continuously across this boundary, the field
immediately above our grazing field is locally unanchored while the field immediately below
this grazing field locally threads across y = 0 and is rigidly anchored by the dense region
in y < 0. If perturbed, the field above has much more freedom of movement than the field
below, and so tangential discontinuities are likely to form and dissipate (Parker 1994) in the
neighborhood of the grazing field. Such a region of grazing field is called a "bald patch" by
Titov & D´emoulin (1999), who first cited such regions as likely sites of current sheet formation
and heating. Since thermal conduction is much more effective along than across magnetic
flux trajectories, thermal conduction heats up all plasmas that are magnetically connected
to the heated region. Therefore it is reasonable to interpret the sigmoidal emission patterns
as picking out flux trajectories threading threading through bald patches where tangential
discontinuities have formed.
Trajectories executing a couple of turns about the axis tend to project as S-shapes.
Therefore, like Low & Berger's internal field and unlike the Gold-Hoyle field, the potential
field around a helical line current is capable of reconciling the hemispheric preferences for
magnetic helicity and for sigmoidal shape.Fields both inside and outside current-carrying
flux ropes may be contributing to this hemispheric preference. The helical line current
-- 12 --
solution leads us to expect that magnetic fields surrounding flux ropes of a given handedness
prefer to twist about this current structure with the same handedness. In reality, a flux rope
which has initially acquired twist through the kink instability can readily transfer some of its
twist to writhe of the same handedness, and this can create twist of this handedness in the
surrounding field. This makes these fields a viable source of "bald patches", Parker current
sheets and sigmoidal emission patterns.
-- 13 --
Fig. 2. -- Some sample magnetic flux trajectories for the case k = 1, viewed from an oblique
angle (top left), from above (top right), along the x-axis (bottom left) and along the y-axis
(bottom right). The shading of the coronal base y = 0 represents the vertical field component
from white, maximum positive flux, to black, maximum negative flux.
-- 14 --
3. Conclusion
Using a simple solution for the magnetic field associated with a helical line current, we
have investigated some consequences of the writhe of coronal magnetic flux ropes for the near-
potential magnetic field within which they are embedded. Magnetostatic solutions in realistic
three-dimensional geometry are generally intractable because of the complex structure of the
governing equations. Two simplifications allow us to investigate some consequences of three-
dimensional structure. The idealization of highly twisted, confined magnetic flux ropes as
line currents allows us to focus on the relatively current free field around a flux rope without
having to treat the full magnetohydrostatic problem. No solutions are known which combine
a current-carrying flux rope field and a potential ambient coronal field. Such an equilibrium is
likely to require a numerical treatment. Helical symmetry is imposed so that the governing
equation becomes the helical Grad-Shafranov equation, an elliptic equation with a linear
differential operator whose source solution describes a line current and can be computed in
closed analytical form.
The most striking difference between potential fields around straight line currents and
ring currents on the one hand and potential fields around helical line currents on the other
is the existence of axially-directed magnetic flux in the helical case. The helical line current
field, in including the effects of flux rope writhe, is therefore more topologically complex than
these straight line and ring current fields sometimes used in solar flux rope models. Since
axial magnetic flux is central to the study of coronal flux ropes, this geometry-dependent
presence of axial flux may help us to understand some processes behind solar activity. The
two main manifestations of solar activity in the corona, flares and coronal mass ejections,
are believed to be related to the evolution of flux ropes associated with prominences and
x-ray sigmoids (Low 1996, 2001).
The axial field component of the helical line-current potential field manifests itself as a
twist of the potential field about the helical axis whose handedness is the same as that of
the axis writhe. This suggests that near-potential fields around flux ropes with writhe of a
given handedness twist preferentially with the same handedness as that writhe. This may
have a bearing on the interpretation of observations of certain classes of coronal magnetic
flux rope, including x-ray sigmoids and prominence magnetic fields. As we have seen, the
preference for flux-rope writhe and external magnetic field twist to have the same handedness
suggest that "bald patches" might tend to occur in such fields, and therefore the associated
heating by Parker current sheet dissipation seem likely. Thus some measurements of flux
rope magnetic helicities may derive from external, near-potential fields. The predicted sta-
tistical hemispheric preference for positive and negative magnetic helicities is consistent with
observational results for prominences and sigmoids and past theoretical results for flux rope
-- 15 --
internal fields.
I thank the referee for helpful comments. This work was conducted while the author was
a participant in the National Aeronautics and Space Administration (NASA) Postdoctoral
Program at Goddard Space Flight Center, and was based at National Solar Observatory,
Tucson.
A. Magnetic vector potential of helical line current
This calculation is given in full by Tominaka (2004) and we quote the results here.
The magnetic vector potential associated with a known electric line current is given by
(Jackson 1965)
A =Z
j dS
r − r′
(A1)
where j is given by Equations (15), describing an infinitely thin helical current embedded in
the cylindrical surface r = a, which we denote by S.
The three components of A are
Ar =(cid:26) −2kaIP∞
−2kaIP∞
n=1[Kn+1(nka)In+1(nkr) − Kn−1(nka)In−1(nkr)] sin n(ζ − ζ0),
n=1[In+1(nka)Kn+1(nkr) − In−1(nka)Kn−1(nkr)] sin n(ζ − ζ0),
for r ≤ a,
for r > a,
(A2)
n=1[Kn+1(nka)In+1(nkr) + Kn−1(nka)In−1(nkr)] cos n(ζ − ζ0),
n=1[In+1(nka)Kn+1(nkr) + In−1(nka)Kn−1(nkr)] cos n(ζ − ζ0),
for r ≤ a,
for r > a,
(A3)
n=1 Kn(nka)In(nkr) cos n(ζ − ζ0),
n=1 In(nka)Kn(nkr) cos n(ζ − ζ0),
for r ≤ a,
for r > a.
(A4)
Ik a2
Aϕ =( Ikr + 2IkaP∞
r + 2IkaP∞
Az =(cid:26) −2 log a + 4IP∞
−2 log r + 4IP∞
Amari, T., & Aly, J. J. 1989, A&A, 208, 261
REFERENCES
-- 16 --
Amari, T., & Aly, J. J. 1992, A&A, 265, 791
Anzer, U. 1989, in Dynamics and Structures of Quiescent Prominences, ed. E. R. Priest
(Dordrecht: Kluwer), 143
Bommier, V., Landi Degl'Innocenti, E., Leroy, J. L., & Sahal-Brchot, S. 1994, Sol. Phys.,
154, 231
Canfield, R.C., & Pevtsov, A.A. 1999, in Magnetic Helicity in Space and Laboratory Plasmas,
ed. M. R. Brown, R. Canfield, & A. Petsov (Geophys. Monogr. 111; Washington, DC:
AGU), 197
Casini, R., & Judge, P. G. 1999, ApJ, 522, 524
Casini, R., Lopez Ariste, A. L., Tomczyk, S., & Lites, B. W. 2003, ApJ, 598, L67
Chae, J. 2000, ApJ, 540, L115
Chen, J., Howard, R.A., Brueckner, G.E., Santoro, R., Krall, J., Paswaters, S.E., et al. 1997,
ApJ 409, L191
Chen, J., Santoro, R. A., Krall, J., Howard, R. A., Duffin, R., Moses, J. D., et al. 2000, ApJ
533, 481
Ciaravella, A., Raymond, J.C., Thompson, B.J., Van Ballegooijen, A., Strachan, L., Gardner,
L., et al. 2001, ApJ 529, 575
Dere, K.P., Brueckner, G.E., Howard, R.A., Michels, D.J., and Delaboudiniere, J.P. 1999,
ApJ 516, 465
Fan, Y., & Gibson, S. E. 2003, ApJ, 589, L105
Fan, Y., & Gibson, S. E. 2004, ApJ, 609, 1123
Freidberg, J.P. 1987, Ideal Magnetohydrodynamics (New York: Plenum Press)
Gary, A., and Moore, R.L. 2004, ApJ 611, 545
Gold, T. & Hoyle, F. MNRAS 120, 189
House, L.L. & Smart, R.N. 1982, Sol. Phys. 80, 53
Jackson, J.D. 1965, Classical Electrodynamics (New York: John Wiley & Sons)
-- 17 --
Leroy, J.L. 1989, in Dynamics and Structure of Quiescent Solar Prominences, ed. E.R. Priest
(Dordrecht: Kluwer), 77
Leroy, J. L., Bommier, V., & Sahal-Brchot, S. 1984, A&A, 131, 33
Lin, H., Penn, M.J. & Kuhn, J.R. 1998, ApJ 493, 978
Low, B.C. 1996, JGR 106, 25141
Low, B.C. 2001, Solar Phys. 167, 217
Low, B.C. 2006, ApJ 649, 1064
Low, B.C. & Hundhausen, J.R. 1995, ApJ 443, 81
Morozov, A.I. & Solov'ev, L.S. 1966, in Reviews of Plasma Physics, Vol. 2 (New York:
Consultants Bureau), 1
Morse P.M. & Feshbach H. 1953, Methods of Theoretical Physics (2 Vols.). (New York:
McGraw-Hill)
Neukirch, T. 1999, The Theory of MHD Equilibria lecture series, U. of St. Andrews (URL:
www-solar.mcs.st-and.ac.uk/ thomas/teaching/mhdlect.pdf)
Parker, E.N. 1994, Spontaneous Current Sheets in Magnetic Fields (New York: Oxford Univ.
Press)
Pevtsov, A.A., Canfield, R.C., & Latushko, S. M. 2001, ApJ, 549, L261
Rust, D.M., and Kumar,A. 1996, ApJ 464, L199
Taylor, J.B. 1986, Rev. Mod. Physics 58, 741
Wood, B.E., Karovska, M., Chen, J., Brueckner, G.E., Cook, J.W., and Howard, R.A.: 1999,
ApJ 512, 484
This preprint was prepared with the AAS LATEX macros v5.2.
|
astro-ph/0012110 | 1 | 0012 | 2000-12-05T17:43:07 | Optical Surface Photometry of a Sample of Disk Galaxies. II Structural Components | [
"astro-ph"
] | This work presents the structural decomposition of a sample of 11 disk galaxies, which span a range of different morphological types. The U, B, V, R, and I photometric information given in Paper I (color and color-index images and luminosity, ellipticity, and position-angle profiles) has been used to decide what types of components form the galaxies before carrying out the decomposition. We find and model such components as bulges, disks, bars, lenses and rings. | astro-ph | astro-ph |
A&A manuscript no.
(will be inserted by hand later)
Your thesaurus codes are:
3 (11.19.2; 11.19.5; 11.19.7)
ASTRONOMY
AND
ASTROPHYSICS
Optical Surface Photometry of a Sample of Disk Galaxies.
II Structural Components
M. Prieto1, J. A. L. Aguerri1,2,A. M. Varela1 & C. Munoz-Tun´on1
1.- Instituto de Astrof´ısica de Canarias, E-38200 La Laguna, Tenerife, Spain
2.- Astronomisches Institut der Universitat Basel, CH-4102 Binningen, Switzerland
Received ; accepted
Abstract. This work presents the structural decomposi-
tion of a sample of 11 disk galaxies, which span a range
of different morphological types. The U , B, V , R, and I
photometric information given in Paper I (color and color-
index images and luminosity, ellipticity, and position-angle
profiles) has been used to decide what types of components
form the galaxies before carrying out the decomposition.
We find and model such components as bulges, disks, bars,
lenses and rings.
Key words: Galaxies: optical photometry, galaxies:
structure, galaxies: fundamental parameters, galaxies: spi-
ral
1. Introduction
The fit of structural models to the luminosity distribution
of spiral galaxies is a complicated task. There are several
difficulties, among which we emphasize the following:
1) The presence of components other the bulge and disk.
Sometimes these structures are clearly present in the
form of, for example, large bars or lenses and we can
model them. Often times this is not possible particu-
larly at short wavelengths where the spiral arms, star
formation regions and the dust extinction can modify
the aspect of the smooth light distribution due to the
old population that contributes most of the mass.
2) Bulge models are not unique. All the light-profile func-
tions proposed for ellipticals can be used for the bulges
of spirals. In recent years the generalized exponential
bulge (Sersic 1968; Sparks 1988) has been the most
extensively used model and this law is intrinsicly vari-
able.
3) Although exponential disks are well established, it is
necessary to have photometric data at large radii in
order to avoid contamination from other more central
components (Prieto et al. 1992b). Most of the images of
nearby galaxies do not reach the low brightness levels
to achieve this.
In general, the B/D relation has very little reliability due
to these above-mentioned reasons and also because the
bulge scale length is sometimes smaller than the seeing
disk. Perhaps the most reliable decompositions are those
for galaxies with a very well defined disk. This means, well
sampled and free from other emission features.
In recent decades much work has been done in or-
der to improve the methods for ascertaining the principal
(bulge and disk) components of disk galaxies from their
optical photometric profiles. Kormendy (1977), Boroson's
(1981), Shombert & Bothum (1987), Kent (1987), Capac-
cioli, Held, & Nieto (1987), Andredakis & Sanders (1994),
De Jong (1996), Moriondo, Giovanardi, & Hunt (1997).
All these studies have resulted in gradual progress in this
difficult task.
The majority of the above-mentioned studies analyze
the bulge and disk without considering the presence of
other components. In this article we try to provide a fur-
ther step in the unraveling of the complicated structures
of spiral galaxies, by taking into account all other com-
ponents which may be present in disk galaxies. This is
done by using the multi-color photometric information (
Aguerri et al. (1999), here after Paper I), to determine
what the different components are that comprise each
galaxy before the decomposition is attempted. We do this
for a limited sample of 11 galaxies. In paper I, we pre-
sented the observations and data reduction for this sample
of galaxies.
2. Model Fitting
We use azimuthally averaged profiles to determine the
various components. This preserves information on non-
axisymmetics features, such as bars (m = 2), because
isophotal ellipse fitting allows the ellipticity and the po-
sition angle of the m = 2 structures to be fitted, as re-
flected in the ellipticity and position angle profiles (Woz-
niak et al. 1995; Prieto et al. 1997; Varela, Munoz-Tun´on,
& Simmoneau 1996). However when the bar is very large
and strong, we characterize the bulge and disk not by az-
imuthally averaging but by individual profiles transverse
to the projected bar.
2
M. Prieto et al: Structural components
The profiles are built from ellipses of variable ellipticity
and position angle. To consider this variation is a way to
correct for warps or any large-scale disk or bulge deforma-
tion. The small variations in the luminosity of the disk are
thus averaged and the S/N increased considerably. Points
in the profile affected by additional structures, like rings
or tight spiral arms, can be easily be omitted.
Moriondo et al. (1997) compared the 1D and 2D tech-
niques and concluded that the 1D technique is always less
accurate and sometimes inconsistent with 2D one and that
the inferior quality of the 1-D fits is attested by both by
the larger errors in the parameters and by the larger val-
ues of χ2. They obtained profiles by keeping the position
angle of the ellipses fixed, which introduces an error in the
luminosity profile since isophotes are frequently twisted.
Moreover, their 2D model considers the apparent bulge
and disk ellipticity to be fixed, while in the 1D model,
both parameters were allow to vary. We think that the
result of these two analyses are not comparable.
De Jong (1996) also compares the parameters of the
bulge and disk fitted in the 1D and 2D decompositions in
his Figure 6. The only model used for all the galaxies in
their 1D fit is that of bulge+disk neglecting the obvious
presence of bars in many of the galaxies of his sample;
however, the bars are taken into account in his 2D fit.
Again, these two analysis cannot be compared.
2.1. The components
We assume that the luminosity distribution of a galaxy
is the sum of the distributions of its individual compo-
nents. We use the following laws for fitting the different
components found in the galaxy sample:
Bulges: We approximate the bulge luminosity profile to
the r1/n law introduced by Sersic (1968):
where Iod is the central surface brightness of the disk and
h is the scale length of the disk.
Bars: We find two types of bars.
1. Elliptical bars. These bars have elliptical isophotes and
we have used the two-dimensional function given by
Freeman (1966) for fitting the luminosity distribution,
Ibar(x, y), of this type of bars:
Ibar(x, y) = Io,barr1 − (
x
abar
)2 − (
y
b bar
)2
(5)
where the free parameters are Io,bar, the bar central
surface brightness, and abar and bbar, which are the
scale lengths for the semi-major and semi-minor axes,
respectively, of the bar. This equation is valid within
the bar region (i.e. x ≤ abar, y ≤ bbar). In equation 5,
the origin is the bar center, the x-axis is in the direction
of the bar major axis and the y-axis is perpendicular
to this.
2. Flat bars. For fitting this kind of profile we use the
expression given in Prieto et al. (1997):
Ibar(r) =
Io,bar
1 + e
r−α
β
,
(6)
where α and β are constants which smooth the end of
the bar.
Lenses: A lens is characterized by having a smooth lu-
minosity gradient with a very sharp cut-off. Duval &
Athanassoula (1983), studying the galaxy NGC 5383,
found a lens in its luminosity profile and fitted it to the
expression:
Ibulge(r) = Ie10−bn(( r
re )1/n−1),
(1)
Ilens(r) = Iol(1 − (r/rol)2), r ≤ rol
(7)
where Ibulge(r) is the surface brightness of the bulge in
flux density at a distance r from the center, or, in surface
magnitude units:
µb(r) = µe + cn((
r
re
)1/n − 1),
(2)
with cn = 2.5bn, and where bn can be chosen in such a
way that the scale-radius, re, is the radius encircling half
of the total luminosity, LT, and Ie is the surface brightness
at this radius.
Disks: The disk luminosity profiles were fitted by an ex-
ponential law, where the surface brightness for the disk in
flux density, Idisk(r), is given by:
where Iol is the surface brightness in flux density at the
lens center , and rol is the typical size of the lens. We used
this expression for fitting the lenses found in this sample.
Rings: Rings and pseudo-rings are present in many spiral
galaxies. The transverse luminosity profiles of these struc-
tures are well fitted with Gaussian functions:
Iring(r) = Iore−
1
2 ( r−ror
σ
)2
,
(8)
where Ior is the surface brightness in flux density at ror,
which is the radius of the ring, and σ is the width of the
ring.
Idisk(r) = Iode−r/h.
Or, in surface magnitude units, by,
µd(r) = µo + 1.086
r
h
,
(3)
The total surface brightness model is the sum of all
the above functions corresponding to the components that
appear in the galaxy:
(4)
Itotal(r) = Ibulge(r)+Idisk(r)+Ibar(r)+Ilens(r)+Iring+...(9)
M. Prieto et al: Structural components
3
2.2. The fitting technique
The most important feature of the method is the detailed
study of the multicolor photometric information for each
of the galaxies, before performing the decomposition, to
determine the different structural components which form
the galaxies and make an estimate of their scale lengths.
After the decomposition we consider whether the param-
eters obtained are physically meaningful.
For an accurate decomposition, the intrinsic peculiar-
ities of each galaxy do not permit an automatic method.
The non-linear χ2 minimization routine method to fit the
model profiles to the data points is quicker than interac-
tive methods, and perhaps more suitable when treating
a more extensive sample of galaxies, but when we ana-
lyze a small number of galaxies in detail, the automatic
methods are not useful. For example, some points in the
profiles of galaxies with a blue ring must be avoided and
the model must pass below them. An automatic method
cannot do this. Moreover the χ2 method is not useful for
analyzing galaxies with complex structures. The param-
eters obtained with this method are often not physically
meaningful.
The process begins by examining the color and color-
index images, which give us a qualitative idea of what the
components are. Next, we examine the color-index pro-
files, where we can estimate the scale length of any addi-
tional components. Then, the confirmation of the assumed
components is obtained in the ellipticity and position-
angle profiles, where the geometry of the different struc-
tures projected onto the sky plane are well reflected. At
this point we decide the set of structural components that
make up the galaxy. Then we decompose the azimuthally
averaged or individual (for galaxies with strong bars) lu-
minosity profiles of the galaxy to obtain the parameters
of their components.
We use an interactive profile-fitting routine. The rou-
tine begins by fitting the parameters for the disk and bulge
over different ranges of the profile, defined by the linear
trends of the surface brightness against r and r1/n, re-
spectively, with the least-squares method in an iterative
process. Beginning with the estimated initial values, we
fit the other components (bars or lenses), by varying their
parameters until a good fit is obtained. These were then
subtracted from the original profile and the bulge and disk
fitted again. This process is repeated until all parame-
ters convergence. We define convergence to have occurred
when the difference between the structural parameters of
the bulge and disk for two consecutive steps is smaller
than the fit errors of those parameters.
For the galaxies, NGC 1300, NGC 7479, and
NGC 7723, which have prominent bars, we used profiles
along the major and minor axes of the bar instead of
the azimuthal profile. On the minor-axis profile, (which
is less affected by the bar component), corrected for incli-
nation and position angle, we fit the bulge, disk, and other
components using the method described above. The fitted
bulge and disk functions are subtracted from the major-
axis profile to characterize the bar in its long dimension.
This process is repeated until all parameters convergence
as described above. Some bars, such as NGC 7723, and
NGC 1300, show very strong star formation regions at
their ends, which were fitted with Gaussians. These re-
gions are associated with the beginning of spiral arms.
The features created in a profile due to spiral arm cross-
ing it are also well fitted by Gaussian functions.
3. Structural decomposition of the galaxies
All the photometric information needed to analyze the
galaxies is given in Paper I. The figures in that paper
should be viewed in conjunction with this section.
The calibration constants in Paper I include the cor-
rection due to galactic inclination and absorption; con-
sequently all parameters are corrected for these effects.
We have not corrected for internal extinction. According
to Xilouris et al. (1999), a typical face-on spiral galaxy
is transparent and is optically thick, at least in the cen-
tral regions, down to inclination angles of almost 60◦. The
galaxies in this sample are near to or below these limits.
In Figs. 1 -- 11 we present the decomposition of the lu-
minosity profiles in each filter and in Tables I-XI we show
the parameters of the models as defined in Sec. 2.1: the
fractional luminosity of each component, and the elliptic-
ity and position angle of the disk and bar defined as the
average value of the last points for the disk and the value
at the typical length for the bar.
The uncertainty given for the bulge and disk parame-
ters comes from the "standard errors" of the coefficients of
the lines of regression of the fits. These "standard errors"
are a measure of the residuals between the observations
and the fitted regression line. Clearly, these residuals de-
pend strongly on the deformation of the profile due to
structures which were not fitted, such as spiral arms, star
formations regions, etc...
The fit for bar and lens parameters was achieved by
varying by a fixed amount the parameters of their lumi-
nosity laws around the estimated values. This number is
the uncertainty in the parameters of the bars and lenses
given in Tables 1 -- 11.
Next, we describe the decomposition procedure for
each galaxy. The morphological classification given is that
of de Vaucouleurs et al. (1991).
NGC 1300. This is an SB(rs)bc galaxy. In the color and
color-index images and profiles (Fig.1a, and 2a, Paper
I), we can distinguish a prominent bar of about 70′′ (on
the major axis), a very blue region at the center, two
dust lanes along the bar, curved around the center, which
could be related to the presence of an ILR (Athanossoula
1992a,b). Two prominent blue spiral arms, and a uniform
color region inside 50′′ which suggests the presence of a
lens. The bar region has the same color as the disk, sug-
4
M. Prieto et al: Structural components
gesting that the stars of the bar are of the same type as
those in the surroundings. The galaxy seems to have the
following components: a bulge, a disk, a bar, and a lens.
The ellipticity profile (Fig. 2a, Paper I) confirms the pres-
ence of these components. In Fig. 1 we present the pro-
files perpendicular and parallel to the bar with the various
components fitted. The bulge is fitted with an index n=4
in all filters, but not inside 10′′ in the perpendicular bar
profile, probably due to the strong extinction in this re-
gion. The bar is very well fitted by an elliptical function.
The model for the parallel profiles inside 30′′ does not fit
the observations well, due to the two dust-lanes parallel
to the bar. At the end of the bar there are star forma-
tion regions, which are well fitted by a Gaussian function.
However, in the parallel profile at 150′′ there is a hump
in the luminosity due to a spiral arm which is not well
fitted by a Gaussian, due the presence in this direction
of a very narrow peak, probably caused by a giant star
formation region. The hump in the perpendicular profiles
around 120′′ is a spiral arm which is well fitted by a Gaus-
sian profile. In this galaxy, the errors of the scale lengths
of the disk are so large that no conclusions can be drawn
concerning the relative values of the parameters of the
different components or the bulge-to-disk (B/D) ratio.
NGC 5992. The morphological type of this galaxy is not
clear; it is classified as simply S-type by de Vaucouleurs et
al. (1991) and as Markarian 489 by Mazzarella & Balzano
(1986). It presents an active starburst nucleus (Balzano
1983; Bicay et al. 1995). In the B−I color-index image
(Fig. 1b, Paper I) some structure appears inside 10′′, a red
region to the NE from the center, and a blue one to the
E. In the ellipticity profile (Fig. 2b, Paper I) we observe
a feature indicating a bar with a length of about 20′′.
We fit a luminosity profile for the bulge with an index
of 1.5 in all filters, an exponential disk, and a flat bar.
The observational feature appearing above the model at
about 6.5′′ is probably due to the structure within 10′′.
The B/D ratio is high, which suggests that it is an early-
type spiral with structural deformations probably due to
an interaction with NGC 5993. In Fig. 2 we present the
fit of the all these parameters to the surface-brightness
profiles of the galaxies.
NGC 6056: This is a lenticular barred galaxy, classified
as SB(s)0. In the color-index map (Fig. 1c, Paper I) we
observe a structure of constant color, which probably cor-
responds to a lens ending at about 20′′. The ellipticity and
position-angle profiles reveal the presence of a bar about
8′′ in length. The lens is very faint in the optical profiles,
but is prominent in the infrared filters, as we will show in
a future paper. The bar is flat and has a misalignment of
about 10◦ with respect to the line of nodes of the disk (see
Fig. 2c, Paper I). We obtained a very good fit with a flat
bar feature. The re of the bulge increases, and the scale
length of the disk decreases, with redder filters. The scale
length of the bar in I is smaller than in the other filters.
NGC 6661. This is a lenticular galaxy, classified as
SA(s)0/a. The color-index images and profiles (Figs. 1d,
and 2d, Paper I) present a constant-color region from
about 10′′ to 30′′, which suggests the presence of a lens. A
red ring appears at about 30′′ in the color-index profiles
and B−I image. We need to add an elliptical lens to the
bulge and disk functions for a good fit to the luminosity
profiles in all filters (Fig. 4). The bulge is fitted with an
index of about 2 in all filters, the disk is smaller in redder
filters, and the scale length of the lens is smaller in I than
in the other filters.
is a late-type spiral classified as
NGC 6946. This
SAB(rs)cd, with an active nucleus, type HII (Keel 1984;
Engelbracht et al. 1996). In the B and U filters it was not
possible to fit ellipses to the isophotes due to the presence
of strong spiral arms and many high luminosity regions,
probably star formation regions. The color and color-index
images (Fig. 1e, and 2e, Paper I) show an extensive cen-
tral region of constant color, suggesting the presence of a
lens. In the color index profile, we can estimate the length
scale of this component, about 70′′, and this can be con-
firmed in the ellipticity profiles where there is a change in
the ellipticity at just about this distance. The misalign-
ment between the inner and the outer isophotes seen in
the position-angle profiles (Fig. 2e, Paper I) may indicate
the presence of a triaxial bulge.
The fluctuations of the observations about the disk
model (Fig. 6) are due to the presence of spiral arms; there
is a very blue arm at about 20′′, which causes the model to
fall below the observations in the bluer filters. The bulge
fits an r1/4 law, including in the central part, and its rel-
ative luminosity increases toward the redder filters. The
scale length and relative luminosity of the disk decreases
greatly with the redder filters. This galaxy presents the
smallest B/D ratio of the sample.
NGC 7013. This is another lenticular galaxy, classified as
SA(r)0/a. Optical and Hα images (Lynds 1974) show a
small bulge, and an inner ring. The H i distribution is in
two rings (Eskridge & Pogge 1991; Knapp et al. 1984).
The larger ring is situated just inside the edge of the opti-
cal disk, and the smaller one is associated with the inner
stellar ring. In the color-index images and profiles (Figs.
1g, and 2g, Paper I) we observe a blue ring feature cen-
tered at about 25′′, and a constant-color region from 10′′
to 60′′, suggesting the presence of a lens. In the ellipticity
profile we can confirm these components. At about 25′′ the
isophotes have an ellipticity close to 0.65 corresponding to
the position of the ring; this is similar to that of the outer
disk. After this, there are a plateau until about 60′′, which
corresponds to the lens. This lens probably has a non-zero
intrinsic ellipticity because it is considerably smaller than
that of the outer disk, and its position angles are quite
different. Beyond this, the ellipticity and position angle of
the ellipses are due to the inclination of the galactic disk.
We have modeled the observed luminosity profiles of this
galaxy with four components (Fig. 6): a bulge, a disk, a
M. Prieto et al: Structural components
5
lens, and a ring. The value of the index n for the bulge de-
pends on the filter, and is about 1.5. The scale length for
the disk and lens is smaller in I than in the other filters.
NGC 7217. This galaxy is classified as (R)SA(r)ab. It is
a LINER (Ho, Filippenko, & Sargent 1993; Hummel et
al. 1987). In the B−I and B−V color map and profiles
(Figs. 1g, and 2g, Paper I) we can see the ring structures,
with three nuclear rings (red, blue and another red) at
about 8′′, 10′′ and 15′′, a blue inner ring, and a blue outer
ring at about 30′′, and 75′′, respectively, measured in the
B−I image. The ellipticity and PA profiles indicate that
the rings are quite circular, and that their isophotes have
similar position angles to those of the disk. Buta et al.
(1995) and Verdes-Montenegro, Bosma, & Athanassoula
(1995) studied the ring structures of this galaxy and their
locations are in agreement with our values. The bulge is
fitted with an r1/2 law in all filters. In addition to the bulge
and disk in the B and V filters, we have fitted the blue
nuclear ring, and the blue inner and outer rings; in the
R and I filters we have also fitted the red innermost ring
and the blue outer ring. All rings are fitted with Gaussian
functions. In Table 7, we give the positions of the rings in
the different filters from the model fit, as shown in Figure
7.
NGC 7479. This is a barred galaxy, cataloged as SB(s)bc;
it is classified as a LINER type by Keel (1983a,b) and
Devereux (1989), and as HII (Hummel et al. 1987). This
galaxy has been studied by many authors. Dynamic stud-
ies by Laine (1996) reveal the presence of an interaction
with another galaxy, which could explain the asymmetry
of the arms. It presents large, bright H ii regions along the
bar (Hua, Donas, & Duun 1980). The central blue peak
that appears in the B−I color-index profile is probably
due to nuclear activity. This galaxy shows a prominent
asymmetric bar, and strong spiral arms in the color and
color-index images (Figs. 1h, and 2h, Paper I). In the bar
region there are strong dust lanes, which present an asym-
metric distribution. The color of the bar and bulge is red-
der than that of the disk. The ellipticity and position angle
profiles (Fig. 2h, Paper I) have the features of a strong bar
about 100′′ in semi-length, its position angle being about
100◦, different from that of the disk. The presence of a
triaxial bulge is suggested by the misalignment between
the inner and outer isophotes. The bar is fitted with a flat
function along the both axes (Fig. 8). In these profiles, it
is possible to see two spiral-arm features, at around 40′′
and 100′′. This inner spiral arm is very smooth in the I
band (it is almost invisible) but is clearly present in the V
band. Elmegreen & Elmegreen (1985), studying the bar --
interbar intensities, also found that this bar has a flat light
profile. The fit is better in the I band than in R and V ,
due the I profile is less affected by the star formation re-
gions present along the bar's major axis (Hummel et al.
1987). The bulge profile fits an exponential function well.
As in NGC 1300, the large error in the scale length of the
disk prevents us from reaching any conclusions concerning
the trend of the parameters with the filters.
NGC 7606. This galaxy is cataloged as SA(s)b with a
LINER-type nucleus (Keel 1983b). The color and color-
index images (Figs. 1i, and 2i, Paper I) show a red nucleus,
a constant-color region inside 20′′, and beyond this there
are very prominent spiral arms. In the B−I map there are
three red pseudo-rings, probably due to the inter regions.
The ellipticity and position-angle profiles (Figs. 1i, and
2i, Paper I) confirm these structures; inside 20′′ the bulge
geometry dominates. From here out to about 60′′ a dif-
ferent structure appears, due to the arms and rings, and
beyond this distance the ellipticity and position angle are
constant due the inclination of the disk. The bulge+disk
model provides a good fit to the observations, including
the central part of the bulge, with an index n of around 2
for all filters. Andredakis & Sanders (1994) also found an
exponential bulge (Fig. 9). The early type of this galaxy,
together with its high inclination, cause the strong spiral
arms to look like rings; they are prominent in the isophotal
luminosity profiles above the bulge-plus-disk model.
NGC 7723. This is an SB(r)b galaxy with HII-type nu-
clear activity (Keel 1983b; Hummel et al. 1987; Giurcin
et al. 1994). In the color and color-index maps and pro-
files (Figs. 1j, and 2j, Paper I) there is clear evidence for
a bar of semi-length about 30′′ and a pseudo-ring at the
end of the bar. The bar is redder than the disk and shows
two straight dust-lanes emanating from the nucleus. The
structure of these dust-lanes, which are not curved inward
toward the center, suggests that there is no ILR in this
galaxy (Athanassoula 1992a,b). The central pixels of this
galaxy are saturated and the results are not significant
inside the central 5′′. The color index of the disk is quite
constant with radius. In the ellipticity and position-angle
profiles (Fig. 2j, Paper I) the bar and pseudo-ring are
clearly present, with different position angles from that of
the disk. The bar is well fit by an elliptical bar function,
and the pseudo-ring by a Gaussian function. The hump
that appears in the perpendicular profiles around 30′′ cor-
responds to the pseudo-ring. This galaxy has smooth and
very broad spiral arms, which correspond to the region
(from 50′′ to 90′′), where the model does not fit the ob-
served profiles. The bulge is fitted with an index n=1 (
exponential law). The re for the bulge and the h for the
disk are constant in all the bands.
NGC 7753. It is in interaction with another galaxy (Salo
& Laurikainen 1993) and is classified as SAB(rs)bc. In
the color and color-index maps (Figs. 1k, and 2k, Paper
I) a small (about 10′′ long) bar is evident. The existence
of such structure is also indicated by the ellipticity and
position-angle profiles (Fig. 2k, Paper I). The luminosity
profiles were modeled with a bulge, a disk, and an ellip-
tical bar (Fig. 11). At around 40′′ there is an increase in
luminosity above the model which is due to a spiral arm.
The bulge was fitted with an r1/2 law.
6
4. Discussion
M. Prieto et al: Structural components
To analyze the results of the structural decomposition of
this sample of galaxies, we divide these in three groups:
those with a well defined disk, those with an ill-defined
disk and those with strong bars and well developed arms.
1.- Galaxies with a well defined disk.
These galaxies have an extensive part of their disk free
from contamination by other components, which implies
that the fitted bulge and disk parameters are reliable,
and that we can make comparisons between them. Our
criterion for selecting these galaxies is that they have at
least 1/3 of their luminosity profile coming from the disk
without overlapping other components. Only four galax-
ies obey this condition: NGC 6056, NGC 6661, NGC 6946,
and NGC 7606 (NGC 6661 fault this condition in the I
band.
The following interpretations, although of an outstand-
ingly coherent behavior, are made with caution due the
small number of galaxies and the reduced statistical sig-
nificance.
In Fig. 12 we present the shape index, n, versus the
filter. There is a trend where n increases with the redder
filters. This can be interpreted as the older population be-
ing more extended than the younger, which could indicate
an end of the star formation in the bulge from outside to
inside. Extinction would have the reverse trend. The ab-
solute variation of n with filter is ±0.3 around the mean
value. The n index is bigger for the Sb galaxy than for the
Scd one. We are not going to compare this index between
spirals and lenticulars.
These galaxies are of different morphological types
(SO, SO/a, Sb, and Sc). Their inclinations are the largest
of the sample, near the transparent limits (60◦), (Xilouris
et al. (1999)), so we expect to find some effect of the inter-
nal extinction on the scale length of the disk, h, depending
on the filter. In Fig. 13 we see how h varies with the filter
for each galaxy. The scale length of the disk decreases in
all galaxies when the filter is redder, as we would expect
an optical thickness increasing towards the center of the
galaxy. However, we can not exclude an additional effect
due to stellar population. The effective radii of the bulges
do not present any systematic trend with the filter.
The ratio of the scale length of the disk to the effec-
tive radius of the bulge, Fig. 14, presents a systematic
decrease to the red. We also find a systematic increase
of re/h from late to early type galaxies, contrary to that
found by Courteau, de Jong, & Broeils (1996) and de Jong
(1996). This problem have been rigorously treated in Gra-
ham & Prieto (1999).
2.- Galaxies with an ill-defined disk.
Although in these galaxies the disk is not well defined,
we have carried out the structural decomposition in order
to determine which components are present and we have
obtained an estimate of their scale lengths.
Fig 12. The index n of the bulge model vs filter for those
galaxies with well defined disk and variable index.
Fig 13. Scale length of the disk against filter for those
galaxies with well defined disk.
NGC 7013. This galaxy has a blue outer ring, seen
clearly in the B−V and B−I images and in the luminosity
profiles. This ring hinders the clear detection of the disk
and this has been fitted to the external edges of the ring.
The galaxy has an inclination close to the transparency
M. Prieto et al: Structural components
7
limit and there is a trend toward decreasing h in redder
filters. We can conclude that this galaxy has a ring at 20′′,
a lens of scale length is 50′′, and a bulge shape index of
1.5.
NGC 7217. This galaxy has multiple blue and red rings
which hinder clear detection of the disk. We have fitted
the disk to the zone we consider to be between rings, but
these rings are very wide and numerous except for the I
filter where they are smoother.
NGC 5992. This galaxy is probably interacting with
NGC 5993. Except in the B filter, there is too little data
for the outer disk to make a good estimate of the disk pa-
rameters. We have fitted the disk to the outermost points
of the galaxy, outside the bar region. We can conclude that
this galaxy has a flat bar with scale length is 19′′.
NGC 7753. This galaxy has very strong spiral arms.
We have carry out the fit of the disk in the inter-arm
regions.
Fig 14. The ratio h/re vs the filter for those galaxies with
well defined disk.
3.- Galaxies with large bars and well defined spiral
arms.
These galaxies are, NGC 1300, NGC 7479, and
NGC 7723. None of these galaxies have a well observed
region of the disk due to the presence of powerful spiral
arms. We have fitted the disk in individual profiles and
on the existing points between the arms. There is no clear
zone of disk in any of these galaxies. However, we have
achieved the fittings to know the general features of the
other components.
NGC 1300. This galaxy has a bar, a lens, prominent
spiral arms and star formation regions.
NGC 7479 and NGC 7723. These galaxies have very
few points to fit the disk. They have an elliptical bar with
scale length of 51′′ and scale wide 25′′ in NGC 7479 and
scale length of 22′′ and scale wide of 12′′ for NGC 7723.
The bulges of the spiral galaxies in this sample, with
the exception of NGC 1300, have an n index which follows
the general trend observed in spiral galaxies. The earliest,
NGC 7217 (Sab), has an index n = 2.8, and the rest have
values between 1 and 2.
We did not find Freeman type II profiles in this sample
of 11 galaxies probably because we considered components
other than the bulge and disk.
We found bars in half of the galaxies analyzed. Half
of these bars have the same color as the underlying struc-
ture (NGC 1300, NGC 6056, and NGC 5992), and the
other half are redder than such structure (NGC 7479,
NGC 7723, and NGC 7753). The cause of this segrega-
tion is a question still to be understood. This could give
us some indication of the state of the bar; the red bars are
probably the oldest, and the young bars retain the color
of the disk stars which give them their form.
5. Summary and Conclusions
We have performed a structural decomposition for a sam-
ple of 11 disk galaxies of different morphological types.
The bulges were fitted with an r1/n law, the disks with an
exponential law, the bars with elliptical or flat functions,
rings and spiral arms with Gaussian functions, and the
lenses with a quadratic expression. Prior to the fit, we used
the U , B, V , R, and I photometric data (color and color-
index images and luminosity, ellipticity, and position-angle
profiles) to decide the type and number of different compo-
nents which form the galaxies and to estimate their scale
lengths. We have written an interactive profile-fitting rou-
tine for the decomposition, which fits the parameters of
the models in an iterative process. We find and model all
components which form the galaxies: bulge, disk, bar, lens,
ring, etc. Only for galaxies with well defined disks do we
give reliable parameters for the bulge and disk.
For the galaxies with well defined disks we find that :
-- The scale length of the disk decreases when the filter
is redder, as would be expected for an optical thickness
increasing towards the center of the galaxy.
-- There is a increased trend in the index n of the bulge
with the redder filters. This can be interpreted as the older
population being more extended than the younger.
-- The ratio between the scale length of the disk and the
effective radius of the bulge, shows a systematic decrease
to the red. We also find a systematic increase in this ratio
from late to early types of galaxies.
For the galaxies with ill-defined disk we have given the
position and length-scales of components other than the
bulge and disk.
8
M. Prieto et al: Structural components
We do not find Freeman type II profiles in this sample
of galaxies, probably because we considered components
other than the bulge and disk.
We found bars in half of the galaxies analyzed, which
are either elliptical or flat. Half of these bars have the same
color as the underlying structure (NGC 1300, NGC 6056,
and NGC 5992), and the other half have redder colors
(NGC 7479, NGC 7723, and NGC 7753).
We are most sincerely grateful to Alister Graham for
useful conversations and kindly reading and improving the
manuscript. We express our thanks to Terry Mahoney for
correcting the English of the manuscript. The 2.5-m INT
is operated on the island of La Palma by the Royal Green-
wich Observatory at the Spanish Observatorio del Roque
de Los Muchachos of the Instituto de Astrof´ısica de Ca-
narias. Support for this work comes from project PB97-
1107 and PB97-0219 of the Spanish DGES. The observa-
tions received financial support from the European Com-
mission through the Access to Large-Scale Facilities Ac-
tivity of the Human Capital and Mobility Programme.
References
Kormendy, J. 1977, ApJ, 214, 359
Laine, S. J. 1996, PhD Thesis, Florida University
Lynds, B. T. 1974, ApJS, 28, 391
Mazzarella, J. M., & Balzano, V. A. 1986, ApJS, 62, 751
Moriondo, G., Giovanardi, C., & Hunt L.K., 1997, A&AS, 130,
81
Prieto, M., Beckman, J. E., Cepa, J., & Varela, A. M. 1992b,
A&A, 257, 85
Prieto, M., Gottesman, S. T., Aguerri, J. A. L., & Varela, A.
M. 1997, AJ, 114, 1413
Prieto, M., Longley, D. P. T., P´erez, E., Beckman, J. E., Varela,
A. M., & Cepa, J. 1992a, A&AS, 93, 557
Salo, H., & Laurikainen, E. 1993, ApJ, 410, 586
Schombert, J.M., & Bothun, G.D., 1987, AJ, 92, 1, 60
Sersic, J. L. 1968, Atlas de galaxias australes (C´ordoba: Ob-
servatorio Astron´onico)
Sersic, J..L., 1982, Extragalactic Astronomy. Reidel, Dor-
drecht.
Spark, W.B., 1988, AJ, 1569 NASA ADS
de Vaucouleurs, G., de Vaucouleurs, A., Corwin, H. G., & Buta,
R. 1991, Third Reference Catalog of Bright Galaxies (New
York: Springer) (RC3)
de Vaucouleurs, G. 1958, Ann, d'Astrophys, 11, 247
Varela A.M., Munoz-Tun´on, C. & Simmoneau E., 1996, Astron.
Astrophys., 306, 381.
Aguerri, J. A. L., Varela, A. M., Prieto, M. & Munoz-Tun´on,
Verdes-Montengro, L., Bosma, A., & Athanassoula, E. 1995,
A&A, 300, 65
Wozniak, H., Friedli, D., Martinet, L., Martin, P., & Bratschi,
P. 1995, A&AS, 111, 115
Xilouris, E.M., Byun, Y.I., Kilafis, N.D., Paleologon, E.U., &
Papamastorakis, J., 1999 (in press).
C. 1999, AJ, 119, 1638.
Andredakis, Y. C., & Sanders, R. M. 1994, MNRAS, 267, 283
Athanassoula, E. 1992a, MNRAS, 259, 328
Athanassoula, E. 1992b, MNRAS, 259, 345
Balzano, V. A. 1983, ApJ, 268, 602
Bicay, M. D., Kojoian, G., Seal, J., Dickinson, D. F., & Malkan,
H. A. 1995, ApJS, 98, 369
Boroson, T. 1981, Ap.JS, 46, 177
Buta, R., van Driel, W., Braine, J., Combes, F., Wakamatsu,
K., Sofue, Y., & Tomita, A. 1995, ApJ, 450, 593
Capaccioli, M., Held, E.V., & Nieto J. -- L.,1987, AJ, 94,1519
Courteau, S., de Jong, R.S., Broeils. A.H. 1996, ApJ, 457, L73
Devereux, N. A. 1989, ApJ, 346, 126
Duval, M. F., & Athanassoula, E. 1983, A&A, 121, 297
Elmegreen, B. G., & Elmegreen, D. M. 1985, ApJ, 288, 438
Elmegreen, B. G., Elmegreen, D. M., Chromey, F. R., Hassel-
bacher, D. A., & Bisseli, B. A. 1996, AJ, 111, 2233
Engelbracht, C. W., Rieke, M. J., Rieke, G. M., & Latter, W.
B. 1996, ApJ, 467, 227
Eskridge, P. B., & Pogge, R. W. 1991, AJ, 101, 2056
Freeman, K. C. 1966, MNRAS, 133, 47
Giurcin, G., Tamburini, L., Mardirossian, F., Mezzetti, H., &
Monaco, P. 1994, ApJ, 427, 202
Graham, A., & Prieto, M., 1999, ApJ, 524, 23L
Ho, L. C., Filippenko, A. V., & Sargent, W. L. W. 1993, ApJ,
417, 63
Hummel, E., van der Hulst, J. M., Keel, W. C., & Kennicutt,
R. C. 1987, A&AS, 70, 517
Hua, C. T., Donas, J., & Duun, H. H. 1980, A&A, 90, 8
de Jong, R.S., 1996, A&AS, 118, 557
Keel, W. C. 1983a, ApJ, 269, 466
Keel, W. C. 1983b, ApJ, 268, 632
Keel, W. C. 1984, ApJ, 282, 75
Kent, S. M. 1987, AJ, 93, 816
Knapp, G. R., van Driel, W., Schwarz, V. J., van Woerden, H.,
& Gallager, J. S. 1984, A&A, 133, 127
M. Prieto et al: Structural components
9
Table 1. Structural parameters of NGC 1300.
Ellipticity
0.39 ± 0.11
Structure
Bulge
Parameters
re(")
µe (mag/arcsec2)
n
% LBulge/LT otal
Disk
h(")
µ0(mag/arcsec2)
Bar
Position Angle
% LDisk/LT otal
B/D
Major axis (Type)
µ0(mag/arcsec2)
a(")
Minor axis (Type)
b(")
Position angle
Ellipticity (b/a)
% LBar/LT otal
rol
Position angle
% LLens/LT otal
B
R
2.86 ± 0.17
18.45 ± 0.16
5.58 ± 0.15
18.5 ± 0.1
4.0
30 ± 1
58 ± 21
23 ± 0.8
96 ± 10
52 ± 30
0.5 ± 0.1
Elliptical
4.0
29 ± 1
70 ± 10
21.9 ± 0.5
0.45 ± 0.07
97 ± 7
55 ± 20
0.5 ± 0.1
Elliptical
95 ± 1
Elliptical
32.4 ± 0.5
100 ± 4
87 ± 1
Elliptical
43.2 ± 0.5
101 ± 2
0.34 ± 0.01
0.49 ± 0.01
9 ± 3
7.4 ± 0.2
64.8 ± 0.5
100 ± 4
78.3 ± 0.5
100 ± 4
I
3.9 ± 0.3
17.2 ± 0.2
4.0
19 ± 1
101 ± 25
21.5 ± 0.8
0.42 ± 0.08
94 ± 8
68 ± 29
0.3 ± 0.1
Elliptical
85 ± 1
Elliptical
416 ± 0.5
101 ± 2
0.43 ± 0.01
7.67 ± 0.05
21.83 ± 0.03
67.5 ± 0.5
100 ± 3
23.61 ± 0.07
22.61 ± 0.03
21.52 ± 0.02
Lens
µl0(mag/arcsec2)
23.75 ± 0.07
22.75 ± 0.03
8.15 ± 0.03
7.97 ± 0.14
5.58 ± 0.05
Table 2. Structural parameters of NGC 5992.
Structure
Parameters
B
V
R
I
Bulge
re(")
µe (mag/arsec2)
n
%LBulge/LT otal
Disk
h(")
Bar
µ0(mag/arsec2)
Ellipticity
Position Angle
%LDisk/LT otal
B/D
Type
µ0(mag/arsec2)
α(")
β(")
Position Angle
4.24 ± 0.06
20.27 ± 0.02
3.65 ± 0.08
19.78 ± 0.03
3.85 ± 0.05
19.28 ± 0.02
3.91 ± 0.09
19.14 ± 0.02
1.5
41.2 ± 0.6
8.63 ± 0.27
20.75 ± 0.27
0.22 ± 0.12
26 ± 18
47.98 ± 0.34
0.85 ± 0.05
Flat
23.3 ± 0.1
19.44 ± 0.5
1.1 ± 0.5
89 ± 3
1.5
36 ± 5
8.18 ± 0.14
20.3 ± 0.1
0.23 ± 0.15
26 ± 23
51 ± 4
0.75 ± 0.06
Flat
22.8 ± 0.1
18.9 ± 0.5
1.1 ± 0.5
93 ± 3
1.5
45.6 ± 0.6
8.85 ± 0.25
20.26 ± 0.07
0.21 ± 0.12
39 ± 18
42.8 ± 0.6
1.06 ± 0.01
Flat
22.5 ± 0.1
18.9 ± 0.5
1.1 ± 0.5
93 ± 3
1.5
47 ± 2
11.15 ± 0.04
20.76 ± 0.04
0.21 ± 0.15
36 ± 35
38 ± 2
1.25 ± 0.07
Flat
22.1 ± 0.1
18.9 ± 0.5
1.1 ± 0.5
93 ± 3
Ellipticity
0.49 ± 0.04
0.51 ± 0.04
0.50 ± 0.03
0.51 ± 0.04
%LBar/LT otal
11 ± 5
13 ± 6
12 ± 6
15 ± 6
10
M. Prieto et al: Structural components
Table 3. Structural parameters of NGC 6056.
Structure
Parameters
B
V
R
I
Bulge
re(")
µe (mag/arcsec2)
n
% LBulge/LT otal
Disk
h(")
µ0(mag/arcsec2)
Ellipticity
Position Angle
% LDisk/LT otal
B/D
Type
Bar
1.94 ± 0.02
21.65 ± 0.02
2.1 ± 0.2
20.74 ± 0.02
2.06 ± 0.07
20.15 ± 0.02
2.22 ± 0.13
19.70 ± 0.06
1.0
7.54 ± 0.02
7.57 ± 0.07
21.29 ± 0.02
0.45 ± 0.06
53 ± 5
83.8 ± 0.4
0.10 ± 0.01
1.1
9 ± 2
7.30 ± 0.12
20.34 ± 0.06
0.47 ± 0.03
53 ± 6
83 ± 3
1.1
9.7 ± 0.5
7.02 ± 0.05
19.67 ± 0.07
0.44 ± 0.06
53 ± 6
88 ± 8
1.3
9.10 ± 0.06
6.89 ± 0.15
18.93 ± 0.08
0.42 ± 0.16
54 ± 15
84 ± 5
0.11 ± 0.02
0.11 ± 0.01
0.11 ± 0.01
Flat
Flat
Flat
Flat
µ0(mag/arcsec2)
23.17 ± 0.03
22.53 ± 0.02
21.92 ± 0.02
21.09 ± 0.02
α(")
β(")
Position Angle
Ellipticity
% LBar/LT otal
8.6 ± 0.5
1.1 ± 0.5
65 ± 1
0.48 ± 0.01
8.62 ± 0.06
9.2 ± 0.5
1.1 ± 0.5
67.7 ± 0.5
0.52 ± 0.01
8.12 ± 0.08
8.6 ± 0.5
1.1 ± 0.5
67.5 ± 0.5
0.51 ± 0.01
7.45 ± 0.03
8.1 ± 0.5
1.1 ± 0.5
65.4 ± 0.7
0.51 ± 0.01
7.40 ± 0.03
Structure
Parameters
B
V
R
I
Table 4. Structural parameters of NGC 6661.
Bulge
re(")
µe (mag/arsec2)
n
% LBulge/LT otal
Disk
h(")
µ0(mag/arsec2)
Ellipticity
Position Angle
% LDisk/LT otal
B/D
Lens
µl0(mag/arsec2)
7.51 ± 0.07
21.13 ± 0.21
7.52 ± 0.09
20.15 ± 0.01
7.56 ± 0.07
19.67 ± 0.01
7.54 ± 0.03
19.40 ± 0.01
1.7
2.0
2.1
27.8 ± 0.5
35.8 ± 0.4
32.8 ± 0.2
1.7
17 ± 3
42 ± 2
21.81 ± 0.06
0.46 ± 0.12
60 ± 10
76 ± 6
28 ± 1
21.0 ± 0.1
0.34 ± 0.14
56 ± 14
63 ± 1
0.22 ± 0.05
21.62 ± 0.04
0.44 ± 0.01
21.33 ± 0.02
rol
Position Angle
% LLens/LT otal
23.7 ± 0.5
53 ± 6
7.1 ± 0.1
24.3 ± 0.5
46 ± 9
8.9 ± 0.1
28 ± 1
21.0 ± 0.1
0.30 ± 0.14
53 ± 16
56 ± 2
0.64 ± 0.01
21.21 ± 0.04
23.2 ± 0.5
49 ± 3
19 ± 1
19.62 ± 0.14
0.3 ± 0.1
51 ± 12
64 ± 6
0.51 ± 0.01
21.14 ± 0.05
17.8 ± 0.5
47 ± 2
7.91 ± 0.03
3.58 ± 0.01
Table 5. Structural parameters of NGC 6946.
Structure
Parameters
Bulge
re(")
µe (mag/arcsec2)
V
R
I
3.73 ± 0.01
17.81 ± 0.07
3.62 ± 0.02
17.32 ± 0.04
3.5 ± 0.1
16.78 ± 0.05
n
1.0
1.0
% LBulge/LT otal
0.61 ± 0.03
0.71 ± 0.12
Disk
h(")
µ0(mag/arcsec2)
Ellipticity
Position angle
% LDisk/LT otal
174 ± 21
19.94 ± 0.06
0.67 ± 0.11
101 ± 7
98 ± 1
114 ± 7
19.29 ± 0.05
0.63 ± 0.06
119 ± 5
98 ± 1
1.0
3 ± 2
68 ± 2
18.73 ± 0.03
0.58 ± 0.11
131 ± 8
95 ± 2
B/D
0.006 ± 0.002
0.011 ± 0.001
0.030 ± 0.001
Lens
µl0(mag/arcsec2)
rol
Position angle
% LLens/LT otal
19.8 ± 0.1
27.5 ± 0.5
110 ± 4
19.6 ± 0.1
27.5 ± 0.5
109 ± 4
0.67 ± 0.04
0.64 ± 0.03
19.3 ± 0.1
27.5 ± 0.5
109 ± 3
2.2 ± 0.4
M. Prieto et al: Structural components
11
Table 6. Structural parameters of NGC 7013.
Structure
Parameters
B
V
R
I
Bulge
re(")
µe (mag/arcsec2)
7.37 ± 0.15
20.14 ± 0.02
7.56 ± 0.13
19.10 ± 0.03
7.74 ± 0.1
19.10 ± 0.01
6.41 ± 0.10
18.40 ± 0.04
n
1.3
1.5
1.6
1.5
% LBulge/LT otal
12.6 ± 0.8
25.88 ± 0.25
23.53 ± 0.16
19.42 ± 0.14
Disk
h(")
µ0(mag/arcsec2)
Ellipticity
Position Angle
% LDisk/LT otal
B/D
Ring
Ior(mag/arcsec2)
ror(")
σr(")
Ellipticity
Position Angle
% LRing/LT otal
µl0(mag/arcsec2)
rol
Position Angle
% LLens/LT otal
Lens
75 ± 9
22.3 ± 0.1
0.67 ± 0.05
62 ± 5
83 ± 2
0.15 ± 0.03
21.3 ± 0.1
21.6 ± 0.5
8.1 ± 0.5
0.48 ± 0.09
78 ± 7
0.83 ± 0.13
23.4 ± 0.1
54 ± 1
71 ± 3
50 ± 5
21.3 ± 0.1
0.65 ± 0.08
63 ± 5
65 ± 1
0.29 ± 0.05
21.3 ± 0.1
21.1 ± 0.5
6.5 ± 0.5
0.44 ± 0.11
80 ± 8
0.52 ± 0.05
22.1 ± 0.1
51.3 ± 0.5
71 ± 3
55 ± 4
21.3 ± 0.1
0.68 ± 0.05
64 ± 4
66.5 ± 0.7
0.34 ± 0.024
20.9 ± 0.1
20.5 ± 0.5
6.2 ± 0.5
0.56 ± 0.14
76 ± 12
0.72 ± 0.04
21.9 ± 0.1
54 ± 1
71 ± 6
45 ± 3
20.40 ± 0.14
0.67 ± 0.11
65 ± 6
66.6 ± 0.2
0.3 ± 0.1
20.9 ± 0.1
20.5 ± 0.5
5.4 ± 0.5
0.5 ± 0.1
82 ± 9
0.33 ± 0.02
20.7 ± 0.1
48.6 ± 0.5
74 ± 2
3.97 ± 0.02
8.23 ± 0.06
9.22 ± 0.01
14.67 ± 0.01
12
M. Prieto et al: Structural components
Table 7. Structural parameters of NGC 7217.
Structure
Bulge
Disk
Parameters
re(")
µe (mag/arcsec2)
n
% LBulge/LT otal
h(")
µ0(mag/arcsec2)
Ellipticity
Position Angle
% LDisk/LT otal
B/D
Blue nuclear ring
Ior(mag/arcsec2)
ror(")
σr(")
B
V
R
I
11.21 ± 0.19
20.70 ± 0.04
10.72 ± 0.22
20.03 ± 0.02
6.96 ± 0.08
18.82 ± 0.03
6.75 ± 0.10
18.21 ± 0.02
2.5
21.30 ± 0.23
26.9 ± 0.7
20.04 ± 0.08
0.09 ± 0.20
91 ± 8
78.1 ± 0.9
0.26 ± 0.01
22.08 ± 0.05
10.8 ± 0.5
2.7 ± 0.5
2.8
22.78 ± 0.58
26.68 ± 0.17
19.48 ± 0.03
0.12 ± 0.05
92 ± 12
76.6 ± 0.9
0.29 ± 0.01
21.13 ± 0.02
10.8 ± 0.5
2.7 ± 0.5
2.5
15.11 ± 0.05
24.26 ± 0.21
18.51 ± 0.03
0.12 ± 0.09
92 ± 29
2.5
15.57 ± 0.21
22.9 ± 0.7
17.88 ± 0.06
0.09 ± 0.23
92 ± 57
84.58 ± 0.17
0.18 ± 0.01
84.11 ± 0.48
0.18 ± 0.01
Ellipticity
Position Angle
0.91 ± 0.03
0.91 ± 0.02
73 ± 11
78 ± 9
% LN uclear ring/LT otal
0.07 ± 0.01
0.1 ± 0.06
Red nuclear ring
Ior(mag/arcsec2)
ror(")
σr(")
Ellipticity
Position Angle
Inner ring
% LN uclear ring/LT otal
Ior(mag/arcsec2)
Outer ring
ror(")
σr(")
Ellipticity
Position Angle
% LInner ring/LT otal
Ior(mag/arcsec2)
ror(")
σr(")
Ellipticity
Position Angle
23.1 ± 0.1
31.3 ± 0.5
2.7 ± 0.5
22.35 ± 0.07
30.2 ± 0.5
5.4 ± 0.5
0.83 ± 0.03
0.86 ± 0.03
83 ± 7
0.08 ± 0.02
23.4 ± 0.2
75.6 ± 0.5
8.1 ± 0.5
0.85 ± 0.09
73 ± 20
83 ± 8
0.18 ± 0.08
23.6 ± 0.2
75.6 ± 0.5
10.8 ± 0.5
0.90 ± 0.05
83 ± 15
20.33 ± 0.05
19.44 ± 0.05
8.1 ± 0.5
2 ± 1
8.1 ± 0.5
2.7 ± 0.5
0.97 ± 0.03
0.92 ± 0.02
75 ± 25
74 ± 7
0.06 ± 0.01
0.12 ± 0.02
23.0 ± 0.1
75.1 ± 0.5
9.72 ± 0.5
0.89 ± 0.13
89 ± 21
22.9 ± 0.1
75.6 ± 0.5
10.8 ± 0.5
0.94 ± 0.06
69 ± 27
% LOuter ring/LT otal
0.44 ± 0.03
0.28 ± 0.05
0.24 ± 0.06
0.18 ± 0.09
M. Prieto et al: Structural components
13
Table 8. Structural parameters of NGC 7479.
Structure
Bulge
Parameters
re(")
µe (mag/arcsec2)
V
R
I
7.22 ± 0.09
20.45 ± 0.19
5.7 ± 0.4
5.3 ± 0.4
19.39 ± 0.04
18.77 ± 0.09
n
1
1
1
% LBulge/LT otal
11.32 ± 0.18
10.98 ± 0.32
9.68 ± 0.08
Disk
h(")
µ0(mag/arcsec2)
Ellipticity
Position Angle
% LDisk/LT otal
B/D
Major axis (Type)
µ0(mag/arcsec2)
Bar
45 ± 5
21.86 ± 0.19
0.24 ± 0.06
118 ± 9
62.65 ± 0.15
0.17 ± 0.02
36 ± 6
20.9 ± 0.8
0.26 ± 0.05
116 ± 7
57.05 ± 0.15
0.18 ± 0.05
40 ± 3
20.44 ± 0.14
0.25 ± 0.08
119 ± 11
61.93 ± 0.15
0.15 ± 0.04
Flat
Flat
Flat
20.83 ± 0.01
20.03 ± 0.01
19.39 ± 0.01
α(")
β(")
51.0 ± 0.5
9.2 ± 0.5
Minor axis (Type)
Flat
51.0 ± 0.5
9.2 ± 0.5
Flat
19.0 ± 1.0
9.8 ± 0.5
99 ± 2
52.0 ± 0.5
8.6 ± 0.5
Flat
17.5 ± 0.5
9.2 ± 0.5
99 ± 2
19.0 ± 1.0
9.8 ± 0.5
98 ± 2
α(")
β(")
Position Angle
Ellipticity
% LBar/LT otal
Structure
Parameters
Bulge
re(")
µe (mag/arcsec2)
n
% LBulge/LT otal
Disk
h(")
0.76 ± 0.02
26.03 ± 0.13
0.74 ± 0.02
31.95 ± 0.11
0.72 ± 0.02
28.38 ± 0.18
Table 9. Structural parameters of NGC 7606.
B
V
R
I
5.83 ± 0.11
21.84 ± 0.02
4.99 ± 0.05
20.73 ± 0.01
7.4 ± 0.2
20.74 ± 0.03
5.86 ± 0.06
19.74 ± 0.02
2.1
3 ± 1
51 ± 6
2.0
4 ± 1
43 ± 5
20.9 ± 0.3
0.59 ± 0.02
56 ± 2
96 ± 1
2.5
5.84 ± 0.24
43 ± 4
20.39 ± 0.25
0.59 ± 0.02
56 ± 2
94 ± 2
2.3
6 ± 1
40 ± 4
19.71 ± 0.25
0.59 ± 0.02
56 ± 2
94.7 ± 0.7
µ0(mag/arcsec2)
Ellipticity
Position Angle
% LDisk/LT otal
21.85 ± 0.25
0.59 ± 0.03
53 ± 2
96.8 ± 0.7
B/D
0.034 ± 0.001
0.041 ± 0.003
0.067 ± 0.007
0.058 ± 0.002
14
M. Prieto et al: Structural components
Table 10. Structural parameters of NGC 7723.
Structure
Bulge
Parameters
re(")
B
3 ± 1
µe (mag/arcsec2)
20.66 ± 0.06
V
R
I
3.55 ± 0.14
19.63 ± 0.08
3.79 ± 0.04
19.46 ± 0.04
3.91 ± 0.02
18.49 ± 0.01
n
% LBulge/LT otal
Disk
h(")
µ0(mag/arcsec2)
1
5 ± 4
23 ± 7
21 ± 2
1
9 ± 1
22 ± 4
21 ± 1
1
8 ± 1
21 ± 3
20 ± 1
Ellipticity
0.31 ± 0.08
0.31 ± 0.06
0.32 ± 0.08
1
9.25 ± 0.05
22 ± 5
19.12 ± 0.01
0.31 ± 0.02
126 ± 2
86.1 ± 0.7
0.1 ± 0.1
Elliptical
20.3 ± 0.01
23 ± 1
Elliptical
11 ± 1
156 ± 3
Bar
Position Angle
% LDisk/LT otal
B/D
Major axis (Type)
µ0(mag/arcsec2)
a(")
Minor axis (type)
b(")
Position angle
Ellipticity
127 ± 3
91 ± 14
0.06 ± 0.35
Elliptical
129 ± 7
82 ± 10
0.1 ± 0.1
Elliptical
128 ± 9
86 ± 8
0.09 ± 0.07
Elliptical
22.41 ± 0.02
21.15 ± 0.01
21.02 ± 0.01
22 ± 1
Elliptical
11 ± 1
159 ± 4
0.5 ± 0.1
22 ± 1
Elliptical
13 ± 1
157 ± 5
21 ± 1
Elliptical
14 ± 1
157 ± 4
% LBar/LT otal
3.81 ± 0.32
0.59 ± 0.07
8.87 ± 0.08
0.66 ± 0.07
6.79 ± 0.07
0.47 ± 0.06
5.02 ± 0.04
Table 11. Structural parameters of NGC 7753.
Structure
Parameters
Bulge
re(")
V
4 ± 1
R
3 ± 1
I
4 ± 1
µe (mag/arcsec2)
20.64 ± 0.12
19.54 ± 0.23
19.45 ± 0.09
n
%LBulge/LT otal
Disk
h(")
2
9 ± 2
22 ± 3
2
7 ± 1
21 ± 4
2
12 ± 2
20 ± 3
Bar
µ0(mag/arcsec2)
20.64±0.02
20.18 ± 0.01
19.64 ± 0.01
Ellipticity
Position angle
%LDisk/LT otal
0.2 ± 0.1
150 ± 10
90 ± 2
0.2 ± 0.1
147 ± 11
91 ± 1
0.2 ± 0.1
150 ± 10
85 ± 3
B/D
Type
0.09 ± 0.04
Elliptical
0.08 ± 0.05
Elliptical
0.14 ± 0.09
Elliptical
µ0(mag/arcsec2)
22.29 ± 0.01
21.20 ± 0.03
20.41 ± 0.05
a(")
Position angle
Ellipticity
%LBar/LT otal
8.6 ± 0.5
170 ± 3
0.45 ± 0.03
0.88 ± 0.08
10.8 ± 0.5
171 ± 2
0.45 ± 0.02
2.24 ± 0.12
12.4 ± 0.5
175 ± 3
0.44 ± 0.03
4.31 ± 0.23
M. Prieto et al: Structural components
15
1. Structural decomposition of the surface-brightness profiles along the semi-minor bar axis (left) and semi-major bar axis
(right) of NGC 1300 in B, R, and I.
Fig
16
M. Prieto et al: Structural components
2
c
e
s
c
r
a
/
g
a
m
2
c
e
s
c
r
a
/
g
a
m
16
18
20
22
24
26
0
16
18
20
22
24
26
0
NGC5992 B
bulge
disk
total
bar
10
20
30
Radial distance(arcsec)
NGC5992 R
bulge
disk
total
bar
10
20
30
Radial distance(arcsec)
16
18
20
22
24
2
c
e
s
c
r
a
/
g
a
m
40
26
0
16
18
20
22
24
2
c
e
s
c
r
a
/
g
a
m
40
26
0
NGC5992 V
bulge
disk
total
bar
10
20
30
Radial distance(arcsec)
NGC5992 I
bulge
disk
total
bar
10
20
30
Radial distance(arcsec)
2. Structural decomposition of the average surface-brightness profiles of NGC 5992 in B, V , R, and I.
40
40
Fig
M. Prieto et al: Structural components
17
3. Structural decomposition of the average surface-brightness profiles of NGC 6056 in B, V , R, and I.
Fig
18
M. Prieto et al: Structural components
4. Structural decomposition of the average surface-brightness profiles of NGC 6661 in B, V , R, and I.
Fig
M. Prieto et al: Structural components
19
5. Structural decomposition of the average surface-brightness profiles of NGC 6946 in V , R, and I.
Fig
20
M. Prieto et al: Structural components
6. Structural decomposition of the average surface-brightness profiles of NGC 7013 in B, V , R, and I.
Fig
M. Prieto et al: Structural components
21
NGC7217 B
bulge
disk
total
ring
40
20
Radial distance (arcsec)
60
80 100 120
NGC7217 R
bulge
disk
total
ring
40
20
Radial distance (arcsec)
60
80 100 120
2
c
e
s
c
r
a
/
g
a
m
2
c
e
s
c
r
a
/
g
a
m
18
20
22
24
0
16
18
20
22
0
2
c
e
s
c
r
a
/
g
a
m
2
c
e
s
c
r
a
/
g
a
m
16
18
20
22
24
0
16
18
20
22
24
0
NGC7217 V
bulge
disk
total
ring
40
20
Radial distance (arcsec)
60
80 100 120
NGC7217 I
bulge
disk
total
ring
40
20
Radial distance (arcsec)
60
80 100 120
7. Structural decomposition of the average surface-brightness profiles of NGC 7217 in B, V , R, and I.
Fig
22
M. Prieto et al: Structural components
8. Structural decomposition of the surface-brightness profiles along the semi-minor bar axis (left) and semi-major bar axis
(right) of NGC 7479 in V , R, and I.
Fig
M. Prieto et al: Structural components
23
2
c
e
s
c
r
a
/
g
a
m
2
c
e
s
c
r
a
/
g
a
m
18
19
20
21
22
23
24
25
0
17
18
19
20
21
22
23
24
0
NGC7606 B
bulge
disk
total
40
20
Radial distance (arcsec)
60
80 100 120
NGC7606 R
bulge
disk
total
40
20
Radial distance (arcsec)
60
80 100 120
2
c
e
s
c
r
a
/
g
a
m
2
c
e
s
c
r
a
/
g
a
m
17
18
19
20
21
22
23
24
0
16
17
18
19
20
21
22
23
0
NGC7606 V
bulge
disk
total
40
20
Radial distance (arcsec)
60
80 100 120
NGC7606 I
bulge
disk
total
40
20
Radial distance (arcsec)
60
80 100 120
9. Structural decomposition of the average surface-brightness profiles of NGC 7606 in B, V , R, and I.
Fig
24
M. Prieto et al: Structural components
2
c
e
s
c
r
a
/
g
a
m
2
c
e
s
c
r
a
/
g
a
m
2
c
e
s
c
r
a
/
g
a
m
2
c
e
s
c
r
a
/
g
a
m
18
20
22
24
26
0
16
18
20
22
24
26
0
16
18
20
22
24
26
0
16
18
20
22
24
26
0
NGC7723 B
bulge
disk
total
bar
40
20
Radial distance (arcsec)
60
80 100 120
NGC7723 V
bulge
disk
total
bar
40
20
Radial distance (arcsec)
60
80 100 120
NGC7723 B
star formation
disk
total
bar
bulge
40
20
Radial distance (arcsec)
60
80 100 120
NGC7723 V
star formation
disk
total
bar
bulge
2
c
e
s
c
r
a
/
g
a
m
14
16
18
20
22
24
26
0
2
c
e
s
c
r
a
/
g
a
m
16
18
20
22
24
26
0
20 40 60 80 100 120 140
Radial distance (arcsec)
NGC7723 R
NGC7723 R
bulge
disk
total
bar
2
c
e
s
c
r
a
/
g
a
m
16
18
20
22
24
26
star formation
disk
total
bar
bulge
40
20
Radial distance (arcsec)
60
80 100 120
0
20 40 60 80 100 120 140
Radial distance (arcsec)
NGC7723 I
NGC7723 I
bulge
disk
total
bar
40
20
Radial distance (arcsec)
60
80 100 120
2
c
e
s
c
r
a
/
g
a
m
16
18
20
22
24
26
0
star formation
disk
total
bar
bulge
40
20
Radial distance (arcsec)
60
80 100 120
Fig
10. Structural decomposition of the surface-brightness profiles along the semi-minor bar axis (left) and semi-major bar axis
(right) of NGC 7723 in B, V , R, and I.
M. Prieto et al: Structural components
25
NGC7753 R
ring
bulge
disk
total
20
40
60
80
Radial distance (arcsec)
NGC7753 V
ring
bulge
disk
total
20
40
60
Radial distance (arcsec)
NGC7753 I
ring
bulge
disk
total
20
40
60
Radial distance (arcsec)
16
18
20
22
24
2
c
e
s
c
r
a
/
g
a
m
80
0
80
2
c
e
s
c
r
a
/
g
a
m
2
c
e
s
c
r
a
/
g
a
m
18
20
22
24
0
16
18
20
22
24
0
11. Structural decomposition of the average surface-brightness profiles of NGC 7753 in V , R, and I.
Fig
|
astro-ph/0105327 | 1 | 0105 | 2001-05-18T13:02:35 | A turbulent MHD model for molecular clouds and a new method of accretion on to star-forming cores | [
"astro-ph"
] | We describe the results of a sequence of simulations of gravitational collapse in a turbulent magnetized region. The parameters are chosen to be representative of molecular cloud material. We find that several protostellar cores and filamentary structures of higher than average density form. The filaments inter-connect the high density cores. Furthermore, the magnetic field strengths are found to correlate positively with the density, in agreement with recent observations. We make synthetic channel maps of the simulations and show that material accreting onto the cores is channelled along the magnetized filamentary structures. This is compared with recent observations of S106, and shown to be consistent with these data. We postulate that this mechanism of accretion along filaments may provide a means for molecular cloud cores to grow to the point where they become gravitationally unstable and collapse to form stars. | astro-ph | astro-ph |
Mon. Not. R. Astron. Soc. 000, 1 -- 7 (2001)
Printed 24 October 2018
(MN LATEX style file v1.4)
A turbulent MHD model for molecular clouds and a new
method of accretion on to star-forming cores
D. Balsara1, D. Ward-Thompson2, R. M. Crutcher1,3
1NCSA, University of Ilinois, Champaign-Urbana, Illinois, USA
2Dept of Physics and Astronomy, Cardiff University, PO Box 913, Cardiff CF2 3YB
3Dept of Astronomy, University of Ilinois, Champaign-Urbana, Illinois, USA
Accepted 2001 April 1; Received 2001 March 13; in original form 1999 December 1.
ABSTRACT
We describe the results of a sequence of simulations of gravitational collapse in a tur-
bulent magnetized region. The parameters are chosen to be representative of molecular
cloud material. We find that several protostellar cores and filamentary structures of
higher than average density form. The filaments inter-connect the high density cores.
Furthermore, the magnetic field strengths are found to correlate positively with the
density, in agreement with recent observations. We make synthetic channel maps of
the simulations and show that material accreting onto the cores is channelled along
the magnetized filamentary structures. This is compared with recent observations of
S106, and shown to be consistent with these data. We postulate that this mechanism
of accretion along filaments may provide a means for molecular cloud cores to grow
to the point where they become gravitationally unstable and collapse to form stars.
Key words: stars: formation
1
INTRODUCTION
Surveys of the densest cores of molecular clouds (eg: Ben-
son & Myers 1989; Ward-Thompson et al. 1994) have shown
them to be sites of active star formation, which can collapse
under self-gravity to form objects known as protostars. A
protostar is an object which will evolve into a star, but which
is currently in the process of accreting the major part of
its final main sequence mass (eg: Andr´e, Ward-Thompson
& Barsony 1993 ; Shu et al 1993). The nature of this ac-
cretion process is still a matter of debate, although it is
known that for low-mass stars (M ≤ 4 -- 5 M⊙) the proto-
star finally emerges from its enveloping cloud of material
onto a well-defined pre-main sequence track on the H-R di-
agram (Stahler, Shu & Taam 1980), and hence to the main
sequence.
Much current debate centres around how dense cores in
molecular clouds form initially, and how they subsequently
evolve. One school of thought suggests that cores form by
Jeans-type gravitational instabilities (eg: Blitz & Williams
1997), and subsequently evolve to higher densities by means
of ambipolar diffusion (eg: Ciolek & Mouschovias 1994). This
is a process whereby a large-scale magnetic field threading
the dense core supports the ionised component of the mate-
rial against collapse while the neutral gas diffuses under self-
gravity towards the centre of mass. This causes the centre
of the core to increase in density until a critical mass-to-flux
ratio is reached, and runaway gravitational collapse sets in.
c(cid:13) 2001 RAS
However, there is a growing body of observational evi-
dence that these quasi-static equilibrium processes are not
the only important physical processes in the evolution of
the interstellar medium. Excess emission at higher veloci-
ties than purely thermal emission (e.g. Falgarone & Phillips
1990) have been explained in terms of intermittent velocity
behaviour, which is a characteristic of turbulence. It has also
been shown that the linewidths of molecular gas emission
can be decomposed into a thermal and a non-thermal com-
ponent (e.g. Casselli & Myers 1995), where the non-thermal
component is produced by turbulence. Likewise, it has been
seen (Crutcher 1999) that there is a strong tendency for
molecular cloud material to have velocities that are super-
sonic and mildly sub-Alfvenic. Therefore, any complete un-
derstanding of the physics of the interstellar medium should
include consideration of a self-gravitating magnetohydrody-
namic (MHD) fluid undergoing turbulence.
There have been a number of MHD models that have
recently been published. For example, Gammie & Ostriker
(1996) showed that MHD turbulence can inhibit gravita-
tional collapse. However, that model was limited to a slab ge-
ometry for molecular clouds. Subsequently, Ostriker, Gam-
mie & Stone (1999) presented the results of a 2.5-D MHD
simulation, and showed that the ratio of magnetic to ther-
mal pressures in a molecular cloud provides an indicator as
to how the cloud will evolve to form stars. A number of
authors have shown that turbulence tends to decay on rela-
tively short time-scales compared to the life-times of molec-
2 D. Balsara, D. Ward-Thompson, R. M. Crutcher
Figure 2. Isosurfaces of density and magnetic field magnitude
superposed. Cyan, yellow and red indicate isosurfaces of density
that are four, six and eight times the mean density. Purple and
magenta indicate isosurfaces of the magnitude of the magnetic
field that are five and seven times the mean. Note the spatial
coincidence of density and magnetic field strength.
1999; Benson & Myers 1989 and references therein). Periodic
boundary conditions were utilized. The initial conditions are
very similar to the unforced models described in Balsara &
Pouquet (1999) with the only difference that the simulation
presented here is self-gravitating.
This simulation is one of several isothermal and mildly
adiabatic (adiabatic index of 1.2) simulations that we have
carried out over a range of realistic turbulent rms Mach num-
bers and Alfven Mach numbers. Truelove et al. (1997) found
that isothermal collapse can form gravitationally condensed
structures on all scales, which is unphysical, but that an
adiabatic equation of state can arrest collapse on certain
scales. A compromise to this issue was suggested by Boss
et al. (2000), who advocated using a barotropic equation of
state. We intend to explore this solution in subsequent pa-
pers, but we note that the compromise between adiabatic
and isothermal equations of state that we present herein ap-
pears to overcome the problems noted by Truelove et al.
(1997).
The simulation was carried out on a fixed grid, and it
is well documented that the grid-based criterion needs to be
satisfied for the condensed objects (cores) in order for their
inner structure to be properly analyzed (e.g. Truelove et al.
1997). Satisfying the grid-based Jeans criterion is not always
possible on all scales. However, our focus in this paper is not
on the inner detailed structure of the cores, rather it is on the
accretion that takes place on to the cores. For those length-
scales our conclusions are unaffected by the grid-based Jeans
criterion. We evolved the model using the RIEMANN code
for numerical MHD (Roe & Balsara 1996; Balsara 1998a&b;
Balsara & Spicer 1999a&b).
2.1 Results of the simulations
Figure 1(a) shows the simulated total intensity, assuming op-
tically thin emission, from the model cube after a few cross-
ing times. It can be seen that the model molecular cloud
has some regions of lower than average density, as well as
compact regions of much higher density. Furthermore, the
regions of highest density appear to be connected to one an-
other by linear features. Examination of the whole model
c(cid:13) 2001 RAS, MNRAS 000, 1 -- 7
Figure 1. Panel (a) (upper) shows the simulated total intensity,
assuming optically thin emission, from the model cube once pro-
nounced cores have formed. The red outline identifies the core
magnified in Figure 1(b). Panel (b) (lower) shows the integrated
intensity from the core delineated in Figure 1(a). Velocity channel
maps for this core are shown in Figure 3.
ular clouds (e.g. Stone et al., 1998; MacLow 1999). However,
there are many potential driving mechanisms of turbulence
that could counteract this decay, such as stellar winds, jets
and outflows, as well as large-scale motions such as shear-
ing caused by Galactic differential rotation. In this paper
we present the results of fully three-dimensional simulations
of turbulent MHD flows in molecular clouds and compare
them with observations.
2 THE MODEL
A three dimensional cubic computational domain with 2563
zones that is initially 2 pc across was set up with a density of
10−20 g cm−3. A tapered exponential spectrum of velocity
and magnetic fluctuations was initially used (see Balsara &
Pouquet 1999; Balsara, Crutcher & Pouquet 1996; Balsara,
Crutcher & Pouquet 1999). The initial conditions consisted
of flow that had a turbulent root-mean square (rms) Mach
number of 1.5 and an Alfven Mach number of unity. These
initial conditions are consistent with the observations of sev-
eral molecular clouds, particularly those 'starless' clouds
in which star formation has not yet begun (e.g. Crutcher
Turbulent MHD model for molecular clouds
3
Figure 3. Panels (a) -- (c) (left-hand column) and (d) -- (f) (right-hand column) show simulated channel maps of the same dense core shown
in Figure 1(b) and its surrounding filamentary structure. Each panel represents a single velocity slice through the region. The ratios of
the line of sight velocity to the thermal width are −1.04, −0.80, −0.66, −0.41, −0.27 and −0.07 respectively.
c(cid:13) 2001 RAS, MNRAS 000, 1 -- 7
4 D. Balsara, D. Ward-Thompson, R. M. Crutcher
cube reveals that these linear features are essentially one-
dimensional filaments, rather than two-dimensional sheets
seen 'edge-on'. All of the filaments terminate in at least one
high density region, while many of the filaments link two
high density regions. The filaments arise as a result of con-
verging gas flows in a turbulent medium.
We also see in our simulations that matter accretes on to
the cores along the filaments. The high density cores are all
compact, and have diameters ranging from a few hundredths
of a parsec to little over a tenth of a parsec. The larger cores
are reasonably well-resolved, occupying over twenty zones.
However, all of the cores satisfy the numerical resolution
criterion (e.g. Truelove et al. 1997). The densities of the
cores are high, having typical values of n(H2) ≥6×104cm−3.
These sizes and densities are consistent with those observed
in dense cores in actual molecular clouds (e.g. Benson &
Myers 1989).
We have analyzed the correlation between the magnetic
field strength and the density in such simulations (Balsara
et al. 1999) where we showed that statistically there is a
very strong positive correlation. In Figure 2 we show the
correlation visually using iso-surfaces of density and mag-
netic field magnitude superposed on the same image of an
isolated core. Thus cyan, yellow and red indicate iso-surfaces
of density that are four, six and eight times the mean density.
Purple and magenta indicate iso-surfaces of the magnitude
of the magnetic field that are five and seven times the mean.
We see that both the density and magnetic field correlate
positively and also that they are stretched out in elongated
filamentary structures. Furthermore, the magnetic field di-
rection typically lies along the cores, where it could serve to
'funnel' gas onto cores (see below).
Quantitative analysis shows that several of the cores are
magnetically supercritical, in agreement with the observa-
tional findings (Crutcher 1999). The model has, therefore, il-
lustrated the topological structures of density and magnetic
field in turbulent magnetized molecular clouds and cores.
2.2 Accretion onto cores
Figure 3 shows simulated channel maps of one of the dense
cores and its surrounding filamentary structures. The chan-
nel maps were made assuming optically thin emission. The
region of Figure 1(a) that we have magnified in Figure 3 is
shown by the red outline in Figure 1(a) and is also shown in
close-up in Figure 1(b). The channel maps were obtained by
convolving the line of sight velocity with the thermal broad-
ening. Analysis of the channel maps shows that the magni-
fied region consists of two cores that are very well separated
in velocity space. We focus on one of them here.
Figure 1(b) shows the integrated intensity of the se-
lected core. Each of the panels of Figure 3 represents a single
velocity slice through the region. The caption gives the ratio
of the line of sight velocity to the thermal width. The range
of those numbers clearly shows that the motions are indeed
supersonic. Figure 3(c) corresponds to the core's rest veloc-
ity. From Figure 1(b) we notice that there is one filament
heading in a north-easterly direction from the core.
In most situations it is reasonable to expect that matter
will accelerate along one part of a filament and also deceler-
ate along another part. Gravity should cause the mass to ac-
celerate towards a core, while certain magnetic field topolo-
gies, as well as the presence of shocks or density gradients in
the accreting material, would cause matter to decelerate as
it approaches a core. The simulated channel maps allow us
to dissect the acceleration or decelaration of mass as it flows
along the filaments. In general the filaments can be curved,
but a simple model that assumes linear filaments that make
contact with a core at one of their ends allows us to draw
some simple deductions on how to interpret channel maps.
Figure 4(a) shows a schematic diagram of a core that
has two filaments with accreting matter that is decelerat-
ing towards the core. A little reflection shows that, if there
is pure deceleration of accreting matter that is flowing in
a linear filament towards the core, the local maximum in
the emission intensity should appear to move away from the
core as one steps away (in velocity space) from the core's
rest velocity. This is schematically illustrated in Figure 4(b),
where we show the motion of the intensity maxima in the
channel maps. Pure accceleration of matter in a linear fila-
ment should show the opposite trend in the channel maps. It
should also be pointed out that constant velocity flow along
a curved filament can also appear to be an acceleration or a
deceleration depending on the filament's orientation relative
to the observer.
The schematic diagrams in Figure 4 allow one to inter-
pret the simulated channel maps in Figure 3. Figure 3(a)
shows only the core with a very faint suggestion of filamen-
tary structure. Figures 3(a) -- (f), viewed in sequence, show
the intensity in the filament become progressively stronger
as the core itself becomes fainter. In Figure 3(e) the core is
decidedly fainter than the filament and in Figure 3(f) it is
very much fainter than the filament. By viewing Figures 3(a)
through to 3(f) in sequence one also notices that the inten-
sity maximum in the filament moves away from the core in
a north-westerly direction traced out by the filament in the
total intensity map of Figure 1(b). In view of the discussion
above, we interpret this as deceleration of matter along the
filament in the simulated channel maps in Figure 3.
From the range of velocities quoted in Figure 3 we notice
that the velocity ranges over an entire Mach number in this
case. We have carried out such imaging for numerous cores
in the simulation presented here as well as in numerous other
simulations that we have carried out. While a filamentary
structure that is aligned along the line of sight will not stand
out prominently, we found that most cores have filamentary
structures around them. Larger cores show more filamentary
structures, a fact that finds a natural explanation in this
scenario, as will be shown below. The velocities that develop
along these filaments correspond to a range of one or more
Mach numbers.
What we appear to be observing in Figure 3 is material
decelerating along a filament towards a dense core. Remem-
ber that the densest regions in the simulated molecular cloud
are also the regions of highest magnetic flux density. Thus
the filaments, as well as being regions of higher density, are
also regions of higher magnetic flux density. Figure 2, as well
as our earlier results (see Balsara et al. 1998) have shown us
that the magnetic field lines tend to lie along the filaments.
Hence these filaments are effectively magnetic flux tubes
in the molecular cloud, which are linked to the dense cloud
cores. So we interpret Figure 3 as illustrating how material
can be channelled by magnetic flux tubes and flow onto the
dense cores. Thus the model simulation has indicated one
c(cid:13) 2001 RAS, MNRAS 000, 1 -- 7
V = - 3
X
V = - 2
X
V = + 3
X
V = + 2
X
V = - 1
V = + 1
X
X
Fig 4b
Core
Filament
Eye
Fig 4a
Figure 4. (a) Schematic diagram of a core undergoing accretion
along filaments. The shaded area denotes the core. The two lin-
ear structures correspond to two filaments. The arrows along the
filaments denote velocity vectors. The rest velocity of the core is
taken to be zero. The image of the physical system is shown in
plan view. To break the degeneracy of having the two filaments
appear in the same line we ask the reader to tip the edge of the
figure that is farther from the eye a little above the plane of the
paper. (b) Schematic diagram of the channel maps that would be
produced by the situation in (a). The 'X' in the channel maps
shows the core's location. The shaded blob shows the location of
the intensity maximum at that scaled velocity (which is shown
immediately above each channel map).
method in which dense cores may evolve and grow in mass
by accretion along magnetic flux tubes. We also document
that this process is seen to occur very often for other cores
that form in several of our simulations.
This suggests that the amount of mass built up by a
core over time will depend strongly on the topological struc-
ture of the magnetic fields that link to that core. The fact
that larger cores seem to have more filaments is now consis-
tently explained by the fact that larger cores are connected
to more magnetic flux tubes which, therefore, supply them
with accreting matter at a faster rate. Our new paradigm
for core formation seems to form a substantial number of
magnetically supercritical cores. While consistent with re-
cent observations (e.g. Crutcher 1999) this fact is at variance
with a previous paradigm for core formation (Mouschovias
and Spitzer 1976) which suggests that cores that form are
initially magnetically subcritical and only a portion of the
core becomes supercritical through the workings of ambipo-
lar diffusion.
The difference in the two paradigms stems from the fact
that matter is accreted directly along field lines in these dy-
namical models while the previous models thought of core
formation as taking place through a sequence of quasi-steady
states. In other work (Balsara, Crutcher and Pouquet 1999)
we showed that ambipolar diffusion does not play a signif-
icant role in damping out the turbulence because the non-
linear eddy turnover times are shorter than the ambipolar
diffusion time. Hence the inclusion or exclusion of ambipolar
diffusion does not strongly affect these results. In the next
section we compare the model findings with one particular
molecular cloud region, to discover whether the predictions
of the model may actually occur in reality.
c(cid:13) 2001 RAS, MNRAS 000, 1 -- 7
Turbulent MHD model for molecular clouds
5
3 COMPARISON WITH S106
In Figure 5 we show 13CO channel maps of the S106 molecu-
lar cloud. The black contours trace the 800-µm dust contin-
uum emission and, therefore, serve to identify the location
of the protostellar core. The numbers at the corner of the
figures show the velocity in kms−1 for each of the channel
maps. Figure 5(d) corresponds to the core's rest velocity.
Based on temperature measurements of S106 the thermal
sound speed can be estimated to be 0.5 kms−1
S106 contains at its eastern edge S106-IR, and at its
western edge the core which was identified from its sub-
millimeter dust emission as S106-FIR (Richer et al. 1993).
The latter source has no known outflow or associated pro-
tostar, and so is a candidate core for comparison with our
model. There are two filaments which join this core from the
north-east that can be seen partly over-lapping in Figure 5.
There is also a much smaller filamentary structure that joins
it from the south-east. The two north-eastern filaments over-
lap each other to a great extent in the total intensity maps
but can be separated from each other in the channel maps.
The temperature in the filamentary gas has been measured
and is too low to permit one to interpret the filaments as
outflowing gas.
The channel maps in Figure 5 of the velocity struc-
ture taken with the BIMA interferometer show evidence for
velocity deceleration in the gas that is accreting along the
length of the filaments if one assumes that the filaments are
not curved along the line of sight (Roberts & Crutcher 2001).
However, one of the filaments in S106 appears to be curved in
the plane of the sky, making it difficult to deduce whether we
are seeing acceleration or deceleration along it. In any case,
the existence of filamentary structures and the motion of the
intensity maxima along the filaments as one steps through
the channel maps in Figure 5 shows that these data are in
close agreement with the model predictions shown in Figure
3. Thus we conclude that the accretion process in this core
of S106 is taking place preferentially along the filaments.
Polarimetric observations (Hildebrand et al 1995; Hol-
land et al 1996) suggest that the magnetic field is also pref-
erentially aligned along the long axis of the filaments, a fact
that is also consistent with the simulations. Thus the data
for this cloud suggest a scenario where the magnetic fields
serve to channel the accreting material along the filaments,
and the data are consistent with our hypothesised model sce-
nario. If this hypothesis is proved by further observations,
then we have shown how cores grow to become super-critical,
and hence collapse to form stars.
4 CONCLUSIONS
Using simulated channel maps from a numerical MHD sim-
ulation in conjunction with observational data of S106, we
have been able to arrive at several insights into the process
of accretion onto the molecular cloud cores in which stars
form:
(i) Turbulent MHD processes cause the formation of
high density cores;
(ii) The cores are linked by extended filamentary struc-
tures;
(iii) Several of the simulated cores are found to be mag-
netically supercritical;
6 D. Balsara, D. Ward-Thompson, R. M. Crutcher
Figure 5. 13CO channel maps of the S106 molecular cloud. The black contours trace the 800-µm dust continuum emission and therefore
serve to identify the location of the protostellar core. The numbers at the corners of the figures show the velocity in kms−1.
c(cid:13) 2001 RAS, MNRAS 000, 1 -- 7
Turbulent MHD model for molecular clouds
7
Roe P. L., Balsara D. S., 1996, SIAM J. Appl. Math., 56, 57
Shu F. H., Najita J., Galli D., Ostriker E., Lizano S., 1993, in:
Levy E. H., Lunine J. I., eds., 'Protostars and Planets III', 3,
University of Arizona Press, Tucson
Stahler S., Shu F. H., Taam R. E. 1980, ApJ, 242, 226
Truelove J. K., Klein R. I., McKee C. F., Holliman J. H., Howell
L. H., Greenough J. A., 1997, ApJ, 489, L179
Ward-Thompson D., Scott P., Hills R. E., Andr´e P., 1994, MN-
RAS, 268, 276
(iv) The higher density filaments are aligned with the
magnetic field structure;
(v) The accretion onto the cores takes place along the
filaments;
(vi) This suggests a scenario where the material is chan-
nelled by the magnetic field and flows under the gravita-
tional influence of the dense cores at the ends of the fila-
ments, accreting onto the cores themselves;
(vii) The amount of matter accreting on to a core thus
depends on the topological structure of the magnetic fields
that link to that core;
(viii) Our observations of S106 appear to show this pro-
cess occurring in a real molecular cloud.
If observations of other regions show similar results,
then this would show that we have distinguished for the
first time an important way in which star-forming cores ac-
crete matter to the point where they have gained sufficient
mass to undergo dynamical collapse and form stars.
ACKNOWLEDGMENTS
DB wishes to thank A.Pouquet and C.McKee for useful dis-
cussions. We also wish to thank S.Levy and R. Patterson for
help with the images. The use of SDSC supercomputer time
is also gratefully acknowledged.
REFERENCES
Andr´e P., Ward-Thompson D., Barsony M. 1993, ApJ 406, 122
Benson P. J., Myers P. C., 1989, ApJS, 71, 89
Balsara D. S., Crutcher R. M., Pouquet A., 1996, in: Holt, S.,
Mundy L. G., eds., 'Star Formation Near and Far', p. 89, AIP
Press
Balsara D. S., Crutcher R. M., Pouquet A., 1999, ApJ, submitted
Balsara D. S., 1998a, ApJS, 116, 119
Balsara D. S., 1998b, ApJS, 116, 133
Balsara D. S., Pouquet A., 1999, Physics of Plasmas, 6, 89
Balsara D. S., Pouquet A., Ward-Thompson D., Crutcher R. M.,
1999, in: Franco J., Carraminana A., eds., 'Interstellar Tur-
bulence' Cambridge Contemporary Astrophysics, p. 261
Balsara D. S., Spicer D. S., 1999a, J. Comp. Phys., 148, 133
Balsara D. S., Spicer D. S., 1999b, J. Comp. Phys., 149, 270
Benson P., Myers P. C., 1989, ApJS, 71, 89
Blitz L., Williams J. P., 1997, 488, L145
Boss A. P., Fisher R. T., Klein R. I., McKee C. F., 2000, ApJ,
528, 325
Caselli P., Myers P. C., 1995, ApJ, 446, 665
Ciolek G. E., Mouschovias T. Ch., 1994, ApJ, 425, 142
Crutcher, R. M., 1999, ApJ, 520, 706
Falgarone E., Phillips T. G., 1990, ApJ, 231, 438
Gammie C. F., Ostriker E. C., 1996, ApJ, 466, 814
Hildebrand R. H., Dotson J. L., Dowell C. D., Platt S. R., Schle-
uning D., Davidson J. A., Novak, G. 1995, in: Haas M. R.,
Davidson J. A., Erickson E. F., eds., 'Airborne Astronomy
Symposium on the Galactic Ecosystem' ASP Conference Se-
ries, 73, 97
Holland W. S., Greaves J. S., Ward-Thompson D., Andr´e P., 1996,
A&A, 309, 267
MacLow M., 1999, ApJ, 524, 169
Mouschovias T. Ch., Spitzer L. 1976, ApJ, 210, 326
Ostriker E. C., Gammie C. F., Stone J. M., 1999, ApJ, 513, 274
Richer J. S., Padman R., Ward-Thompson D., Hills R. E., Harris
A. I., 1993, MNRAS, 262, 839
Roberts, D., Crutcher, R. M. 2001, ApJ, submitted
c(cid:13) 2001 RAS, MNRAS 000, 1 -- 7
|
astro-ph/0306594 | 1 | 0306 | 2003-06-27T18:50:53 | Star Formation History and Other Properties of the Northern HDF | [
"astro-ph"
] | The original analysis of the star formation history in the NICMOS Deep images of the NHDF is extended to the entire NHDF utilizing NICMOS and WFPC2 archival data. The roughly constant star formation rate from redshifts 1 to 6 found in this study is consistent with the original results. Star formation rates from this study, Lyman break galaxies and sub-mm observations are now in concordance The spike of star formation at redshift 2 due to 2 ULIRGs in the small Deep NICMOS field is smoothed out in the larger area results presented here. The larger source base of this study allows comparison with predictions from hierarchical galaxy formation models. In general the observation are consistent with the predictions. The observed luminosity functions at redshifts 1-6 are presented for future comparisons with theoretical galaxy evolution calculations. Mid and far infrared properties of the sources are also calculated and compared with observations. A candidate for the VLA source VLA 3651+1221 is discussed. | astro-ph | astro-ph |
Astrophysical Journal
Star Formation History and Other Properties of the Northern
HDF
Rodger I. Thompson
Steward Observatory, University of Arizona, Tucson, AZ 85721
ABSTRACT
The original analysis of the star formation history in the NICMOS Deep
images of the NHDF is extended to the entire NHDF utilizing NICMOS and
WFPC2 archival data. The roughly constant star formation rate from redshifts
1 to 6 found in this study is consistent with the original results. Star formation
rates from this study, Lyman break galaxies and sub-mm observations are now
in concordance The spike of star formation at redshift 2 due to 2 ULIRGs in the
small Deep NICMOS field is smoothed out in the larger area results presented
here. The larger source base of this study allows comparison with predictions
from hierarchical galaxy formation models. In general the observation are consis-
tent with the predictions. The observed luminosity functions at redshifts 1-6 are
presented for future comparisons with theoretical galaxy evolution calculations.
Mid and far infrared properties of the sources are also calculated and compared
with observations. A candidate for the VLA source VLA 3651+1221 is discussed.
Subject headings: Early Universe -- galaxies:evolution -- galaxies:distances and
redshifts
1.
Introduction
The Hubble Deep Fields (HDFs) are rich sources of data on cosmology and the evolution
of galaxies. The large number of observations at many different wavelengths make the
Northern HDF (NHDF) particularly useful. The NHDF is the only HDF that has complete
coverage with both WFPC2 and NICMOS. The NICMOS coverage, however, is not to the
same depth as the WFPC2 due to the smaller area of the NICMOS Camera 3. This is
somewhat compensated for by the red nature of evolved galaxies and the redshift of visible
light into the infrared bands for young blue galaxies.
-- 2 --
This work studies the star formation history in the NHDF and is an extension of a similar
study utilizing the smaller NICMOS Deep field (Thompson, Weymann and Storrie-Lombardi
2001) (hereinafter TWS). The increased area and number of galaxies greatly improves the
statistical significance of the results over TWS. Shallower coverage, however, increases the
photometric error, making it comparable or dominant over the large scale structure error
which decreased due to the larger field.
Other researchers, Lanzetta, Yahil & Fern´andez-Soto (1996); Fern´andez-Soto, Lanzetta,
& Yahil (1999); Lanzetta et al. (2002) have utilized the same data set, plus additional ground
based data to investigate star formation history in the NHDF. Our analysis differs signifi-
cantly from those studies by including extinction in the SED templates used to determine
the photometric redshifts. This produces results quite different from Lanzetta et al. (2002)
who do not include extinction. Contrary to Lanzetta et al.
(2002) we find that the star
formation rate (SFR) is essentially constant in the epoch from z = 1 to z = 6 and that this
rate is significantly larger than the present day SFR (Lilly et al. 1996), consistent with the
results presented in TWS.
2. Observations
Observations utilized in this study are all from the Hubble Space Telescope (HST)
archives. The first set is the processed WFPC2 images of the NHDF produced by Williams
et al.
(1996). The second set is the archival data from the NICMOS survey of the entire
NHDF by Dickinson (2000). The NICMOS F110W and F160W data were reprocessed as
described in § 3. Only the areas corresponding to the three wide field WFPC2 chips are
used since the WFPC2 PC chip images have substantially different signal to noise statistics.
NICMOS Deep NHDF images and data reductions from Thompson et al. (1999) and TWS
are used in the error analysis.
All NICMOS images were taken in the SPARSE64 mode with 24 samples after the first
read and integration times of 1344 seconds. Images at 8 positions completely covered the
NHDF. At each position 9 images were taken in a dithered pattern. The 8 positions with 9
dither points in the F110W and F160W filters required 144 exposures.
3. Data Reduction
The data reduction procedures are almost identical to the procedures described in
Thompson et al. (1999) and TWS. For this reason they will only be briefly outlined except
-- 3 --
where there are differences. The WFPC2 images were not altered from the Version 2 images
available in the HST archives. The NICMOS images, however, were reduced from the raw
images obtained in the HST cycle 7 program 7817 by Mark Dickinson. The images were
examined for effects of cosmic ray persistence from SAA passages and 17 of the 144 images
were rejected. The background flux was removed by taking the median of the images in each
filter that were not affected by cosmic ray persistence and subtracting it from each image.
The 9 dithered images for each of the 8 positions were combined using the standard
Drizzle procedures with 0.1 arc second pixels, one half the linear size of the NICMOS Cam-
era 3 pixels. The IDL procedure IDP3 (Lytle et al. 1999) produced a mosaic of the 8 images.
IDP3 also reduced the WFPC images to the NICMOS resolution with a bi-cubic spline in-
terpolation and aligned them with the NICMOS mosaic. Inspection of the images showed
regions of increased noise at the boundaries between images. These areas were masked out
as shown in Figure 1. Since this study is primarily a statistical study, the small amount of
masked area does not affect the conclusions.
-- 4 --
Fig. 1. -- The masked areas shown on the WFPC2 F814W filter image of the NHDF.
-- 5 --
Source positions were found using the method of Szalay, Connolly, & Szokoly (1999)
as described in TWS. A Gaussian fit to the histogram of pixel values determined the global
variance of the image mosaic in each filter. The histogram is dominated by the background
noise with true sources creating a deviation from the Gaussian fit at large positive values.
The 1 σ noise levels in both the F110W and F160W NICMOS filters are 5 × 10−4 ADUs per
second. For comparison the NICMOS Deep Field has a 1 σ noise level of 2.2 × 10−4 ADUs
per second. 5 × 10−4 ADUs per second corresponds to 1.4 × 10−9 Janskys. The drizzling
procedure, however, introduces correlation, therefore, the true noise level is about 1.6 times
larger than the measured Gaussian noise.
The Szalay source identification procedure requires images with a local variance of one.
The local variances for the F110W and F160W fields were found by drizzling and mosaicing
the flat fields of the two filters in exactly the same way as the images. This accounts for
differences in sensitivity over the area of the NICMOS Camera 3 detector. The flats have a
median of 1, therefore, the variance images also have a median of 1. The ratio of the imaged
fields to the variance image was scaled to the average variance creating a local variance of
1. The variance in the WFPC filters was assumed constant across the field. This procedure
differs slightly from the procedure used in TWS.
The threshold value of the Szalay R parameter, √χ, was set to 4.5, higher than the 2.3
value used in TWS but more appropriate for the six dimensional χ2 analysis of the combined
WFPC and NICMOS fluxes. Unlike TWS all six fluxes are included in the Szalay procedure
to insure that small blue galaxies with little infrared flux are not missed in the source
extraction. The source selection criterion for determining the star formation history (§ 5.1),
however, does discriminate against galaxies with no F110W or F160W flux. Figure 2 shows
the distribution of R image pixels. All pixels with R values higher than or equal to 4.5 are
considered legitimate source pixels. Pixels with R values less than 4.5 are assigned random
values with average values 106 below the valid pixels. This forces the source extraction
program to only consider pixels chosen by the Szalay procedure. The source extraction
program crashes if all non-valid pixels are set to zero.
The source extraction program SExtractor (SE) (Bertin and Arnouts 1996) extracts
the sources in two passes. The first pass on the R has the extraction threshold set at 4.0,
selecting all valid pixels. As in TWS the DETECT MINAREA parameter is set to 3 so a true
source must have 3 contiguous valid pixels. The source positions, except for the contiguous
area criterion are, therefore, determined by the Szalay procedure, not by SE. SE then runs
again on the six photometric images in a mode that extracts sources in the pixel pattern
determined by the first pass. Run in this manner, SE does not produce daughter objects.
Sources which have several components may be broken into individual objects by this process.
-- 6 --
Fig. 2. -- R image pixel distribution from the Szalay procedure. The asterisks represent the
actual data and the solid line is the expected χ2 distribution with six degrees of freedom.
The boxes show the difference between the data minus the expected distribution. The long
tail at high R is the real source distribution.
-- 7 --
Although listed as separate objects in the catalog the star formation history is unaffected
as long as the calculated photometric redshift is the same for both components. In some
cases a separate listing for two components is an advantage, allowing different extinctions
for different components. § 5.1 discusses additional criteria for final source selection.
The PHOT-APERTURES parameter in the SE configuration file was set to 6, 10, 15 to
produce 0.6′′, 1.0′′, and 1.5′′ diameter aperture fluxes. The photometric, redshift, extinction
and SED calculations use the 0.6′′ aperture fluxes. The total flux for determining the SFR
is found by summing all of the pixels identified by the Szalay procedure as belonging to a
single source.
4. Photometric Redshift, Extinction and SED Determination
The basic procedure for photometric redshift, extinction and SED determination is a
χ2 comparison of the observed fluxes in the six bands to fluxes calculated from numerically
redshifted and extincted SEDs. As described in TWS we draw our templates from three
sources. The first source is the four observed SEDs of Coleman, Wu, and Weedman (1980)
utilized by several authors. These galaxy SEDs have been corrected for galactic extinction
but not for extinction in the galaxies themselves. The galaxies, however, were selected
on the basis of their very low internal extinction. The unreddened SED of the set of mean
SEDs of Calzetti, Kinney & Storchi-Bergmann (1994) provides an additional observed active
star-forming galaxy template (Calzetti (1999a)). A final and even hotter template is a 50
million year old continuous star formation SED calculated from the Bruzual and Charlot
models (Bruzual and Charlot
(1996)) with a Salpeter IMF and solar metallicity. This
theoretical SED does not have emission lines, therefore we have added Hα, (O[III]+ Hβ)
and O[II] emission lines by scaling up the lines from the Calzetti SED by the ratio of the UV
fluxes in the Calzetti and Bruzual -- Charlot SEDs. Template 6 is substantially bluer than the
Calzetti SED.
Extinction is calculated from the starburst and star forming galaxy optimized obscura-
tion law of Calzetti, Kinney & Storchi-Bergmann (1994). Since the primary purpose of this
paper is a determination of the star formation rate it is appropriate to utilize an extinction
law determined from star forming galaxies. The absorption due to extragalactic neutral
hydrogen is calculated from an updated version of the formulation of Madau et al.
(1996).
The calculated fluxes are then interpolated between the 6 SEDs to produce 51 different SED
fluxes for each redshift and extinction point. The details of this procedure are given in TWS.
The grid for the χ2 analysis includes 100 different redshifts between 0 and 8, 15 different
extinctions between E(B-V) of 0 to 1.0 and 51 different interpolated template types for a
-- 8 --
total of 7.65×104 choices. The resulting redshifts, extinctions and SEDs are given in Table 2.
The earliest SED has a value of 1.0 incremented by 0.1 to the latest SED of 6.0. Integer
values refer to the 6 basic SED templates. Extinctions in E(B-V) are incremented by 0.02
from 0 to 0.1 and by .1 from 0.1 to 1.0, the maximum used in the analysis.
The χ2 procedure is identical to TWS except that negative fluxes are not replaced
by zero. The technique minimizes the χ2 residuals between the observed fluxes and those
predicted by the numerically redshifted, extincted and Lyman absorption attenuated SEDs.
It alters the usual error term in the denominator by adding a second term proportional to
the measured flux. The χ2 residual is
6
χ(z, E)2 =
!2
(1)
Xi=1 (fi − A · f mod(z, E)i)
i + (0.1fi)2
pσ2
The index i refers to the six fluxes, fi is the measured flux and f mod(z, E) is the flux
predicted by a template at a redshift of z and extinction E(B-V) = E. Since this is not a
formal χ2 calculation the quantitative probabilities associated with χ2 values are not strictly
valid. The normalization constant A minimizes χ(z, E)2.
A =
6
Xi=1
fi · f mod(z, E)i
i + (0.1fi)2 /
σ2
(f mod(z, E)i)2
σ2
i + (0.1fi)2
6
Xi=1
(2)
The limit of the expression at very low flux levels is the standard form with the formal
background, σ, dominating the denominator and at high flux levels it is the flux difference
between the observations and the model divided by 10% of the flux instead of the σ. At
high flux levels the errors in the fit are proportional to flux since the dominate errors are
intra-pixel sensitivity variations and systematic flux errors. The accuracy of the NICMOS
flux levels may be better than this. It is estimated to be 4 - 5% for this level of dithering
but the output does not appear to strongly depend on the precise coefficient of the flux in
the denominator of equation( 1).
4.1. Photometric Redshift Accuracy
Photometric redshifts have gained a reasonable reputation for accuracy, particularly
at high redshifts where the clear signature of the Lyman break is easy to observe. Two
tests of the photometric redshifts derived in this work are comparison to available spectro-
scopic NHDF redshifts and comparison with previous NHDF photometric redshifts (TWS,
Fern´andez-Soto, Lanzetta, & Yahil (1999), Wang, Bachall, & Turner (1998)).
-- 9 --
4.1.1. Comparison with Spectroscopic Redshifts
The primary source of spectroscopic redshifts is Cohen et al.
(2000), enhanced and
modified by Cohen (2001). The small reductions in the field area described in § 2 reduces
the number of comparison redshifts to 135. Figure 3 shows the results of the comparison.
The standard deviation is 0.29 in redshift, however, it is not constant with redshift. Most of
the error occurs in the redshift range between 0 and 2, as is expected, since the prominent
Lyman break does not appear in our data for that range. Note that galaxies in the redshift
range between 0 and 0.5 are not used in this study.
Catastrophic redshift errors, errors of more than 0.5 in redshift, are overploted with
squares in Figure 3. Analysis of the 11 catastrophic failures reveals three basic types of
failures. The first type is where there is no clear minimum in the χ2 value along the range of
redshifts, extinctions or SED templates. These are characterized by a steadily decreasing χ2
value with a parameter, such as template, until the end of the parameter space is reached.
This failure of the parameters to span the space of galaxy types is seen in 6 galaxies. The
second category contains 5 galaxies that are most probably superpositions of two galaxies
that have not been separated by the SE source extraction process. Both galaxies are con-
tained in the LRIS 1′′ slit width used in the spectroscopic observations (Cohen et al. 2000).
In this case the spectroscopic redshift would be determined by the galaxy with the most
prominent spectral features while the photometric redshift would be most influenced by the
galaxy with the highest continuum level.
The third failure in 2 galaxies is an apparent mild degeneracy in parameters which tends
to favor a redshift of about 0.5 for galaxies with spectroscopic redshifts between 0.5 and 1.0.
These galaxies appear to be well fit at the correct redshift with an intermediate SED and
low value of extinction or at a redshift of ∼.5 with a very hot template and a high value
of extinction. This degeneracy creates a pile up of galaxies at redshifts of 0.48 and 0.56.
Examination of the photometric redshifts for these sources where the extinction is forced
to be zero did not give statistically better fits except in the extreme cases such as those in
Table 1. This type of error can reduce the calculated SFR in the redshift 0.5 to 1.5 bin by
moving some sources out of the bin to a redshift of 0.48 or reducing the calculated UV flux for
a source that has been moved to a lower redshift erroneously. This would make the SFR for
redshift a lower limit, subject to other errors. Note that some galaxies have multiple failure
modes so that the number of failures listed is larger than the number of failed galaxies.
-- 10 --
Table 1 lists the 11 galaxies with redshift errors larger than 0.5 along with their param-
eters and the probable cause of the error. The general conclusion is that the photometric
redshifts are quite adequate for statistical studies but can be subject to significant errors for
individual galaxies.
4.1.2. Comparison with Previous Photometric Redshifts
The comparison between the spectroscopic and photometric redshifts is dominated by
low redshift fairly bright galaxies except for the few faint, high redshift galaxies that have
spectroscopic redshifts through very long spectroscopic observations on the largest ground
based telescopes. It might be expected that the agreement on bright sources may be better
than for the more numerous faint galaxies. One test is a comparison between the photometric
redshifts obtained in the Deep NICMOS field (TWS) with those found in this study. The
NICMOS F110W and F160W images are completely independent between the two data sets
while the WFPC2 image are in common except that KFOCAS was used for source extraction
in TWS and SE is used here. The comparison primarily determines the differences introduced
by noise in the F110W and F160W images. Redshifts from the NICMOS deep NHDF images
should be more accurate than the redshifts from the present study due to the higher signal
to noise of the deep images.
A total of 283 galaxies in the Deep images also appear as legitimate galaxies in this set
of images. Figure 4 shows a histogram of the differences in the redshifts between the deep
and present images. The differences are peaked at zero as expected but do not appear to be a
Gaussian distribution away from the peak. The differences are reasonably equally distributed
between positive and negative values with a slight skew toward negative values (the present
redshift higher than the deep redshift). The percentage of "catastrophic failures", a difference
greater than 0.5, is 17%. Some of the failures create large differences in redshift. Examination
of some cases with large differences reveals significant differences in either the F110W or
F160W flux. As expected there is little variation in the WFPC2 fluxes measured by KFOCAS
and SExtractor. The 17% significant redshift error from this comparison is higher than the
8% error found from the spectroscopic redshift comparison. In the final error analysis we
will assume that 17% of the galaxies have redshift errors that are randomly distributed up
to a maximum of 4.
-- 11 --
Fig. 3. -- Plot of the spectroscopic redshifts from Cohen et al. (2000) versus the photometric
redshifts determined in this work. Some high redshift objects have been added from other
sources. Objects that have redshift differences greater than 0.5 are overplotted with a square
symbol and are discussed in the text and Table 1.
-- 12 --
Fig. 4. -- A histogram of the values of the Deep NICMOS NHDF photometric redshifts
minus the photometric redshifts determined from this study.
-- 13 --
4.1.3. Comparison with Photometric Redshifts from Other Work
There have been several other determinations of photometric redshift in the NHDF but
only two (Wang, Bachall, & Turner 1998; Fern´andez-Soto, Lanzetta, & Yahil 1999) have
readily accessible catalogs of redshift and object position. Figure 5 shows the comparison
between this work and the redshifts obtained by the two groups. Coincidence of objects was
determined by position correspondence within 0.3′′. The redshifts of both previous studies
are offset to larger redshifts than this study. This effect is most probably due to both groups
not correcting for extinction. Without extinction the only mechanism to make a galaxy
redder is to redshift it or choose an earlier template.
-- 14 --
Fig. 5. -- Histograms of the values of the photometric redshifts determined from this study
minus the redshifts from Fern´andez-Soto, Lanzetta, & Yahil (1999) on the left and Wang,
Bachall, & Turner (1998) on the right. In both case the redshifts from the cited works were
subtracted from the redshifts found in this work.
-- 15 --
5. Catalog of Individual Source Properties
Table 2 lists the individual sources, ordered by RA, that satisfy the criteria listed in §5.1.
Column 1 gives the NICMOS identification number. Column 2 is the WFPC2 identification
number of the galaxy from Williams et al.
(1996) if there is positional coincidence within
0.3′′. Columns 3 and 4 are the the redshift and extinction. Column 5 contains the SFR
in solar masses per year as determined in § 6. The bolometric luminosity of the galaxy is
in column 6. The flux of a galaxy is obtained by integrating over the unextincted selected
template scaled by the factor A from equation( 2). The bolometric luminosity then follows
from the redshift and our adopted cosmology. The fraction of the luminosity that is extincted
and therefore goes into far infrared flux is in column 7 followed by the calculated 6, 15 and
850 µm flux (mJy) in columns 8, 9 and 10 for comparison with measured ISO and SCUBA
fluxes (§ 10). The luminosities, ISO fluxes and SCUBA fluxes of galaxies with a redshift of
zero are all set to zero.
Columns 11 and 12 give the template number T of the best fit, and the modified χ2
value of the fit from Equation 1. It should be noted that the distribution of the modified
χ2 values will not rigorously follow a true χ2 distribution. The values are provided to give a
qualitative indication of the relative goodness of fit for the best fit values of redshift, template
type and E(B-V). Column 13 has the total F160W AB magnitude, determined by adding
all of the pixels designated by SE as part of an object, followed by the 0.6 aperture F160W
magnitude (Ap. mag) in column 14. Note that variations in local background as interpreted
by SE can result in an aperture magnitude brighter than a total magnitude for faint galaxies.
If no F160W magnitude is listed the object had a zero or negative measured F160W flux.
Columns 15 and 16 are the right ascension and declination positions of the object. The RA
listing contains only seconds and the DEC listing only minutes and seconds. 12.h 36.m should
be added to the RA and 62.◦ to the DEC. Note that a few sources have RA values of 12.h 37.m
plus seconds. These sources are found at the end of the catalog with seconds values between
0 and 1.2.
5.1. Source Selection Criteria
Data reduction is performed on all sources detected by the Szalay procedure. Additional
criteria for source selection are introduced before sources appear in the catalog of objects
used for determining the star formation history. First all sources that are known stars are
rejected. Next all objects that are too near the edge of the image to insure that no flux is
lost are rejected. This step uses the SE internal flags and visual inspection of the image.
Next a list of suspect objects are rejected from a visual inspection of the images. These
-- 16 --
include objects with significant overlap that have been counted as a single object by SE and
stellar diffraction spikes that were counted as objects. Next all objects with SE internal flags
of 16 or greater, indicating that the photometric aperture was corrupted by some means,
are rejected. Next is a criterion that all valid sources must have a signal to noise of 3.5 or
greater in one band or a signal to noise of 2.5 or greater in two bands. Finally an accepted
source is required to have a flux greater than 5 × 10−4 ADUs per second in both the F110W
and F160W bands. All sources that meet these criteria are included in the catalog but only
sources with redshifts between 0.5 and 6.5 are used in the star formation history analysis.
Two other comments are in order on source selection. At this point, for the small
number of objects that have known redshifts, the photometric redshift is replaced by the
spectroscopic redshift in the analysis procedures. Second the bright galaxy NHDF 3-610 is
rejected from the analysis due to impingment of the diffraction spike from the adjacent star
and a clear overlap with another galaxy. For this reason, even though it is quite bright, it
does not appear in Table 2.
5.2. Distribution of Source Properties
Figure 6 shows the distribution of F160W magnitudes with redshift and the calculated
tracks of an early, a late, and an extincted late L∗ galaxy. The early galaxy is template 1.0,
the late galaxy template 6.0 and the extincted late galaxy is template 6.0 with an extinction
of E(B-V) = 0.2. The sources are color coded with the earliest type galaxy deep red and the
latest galaxy deep blue as indicated by the legend on the figure. As expected the majority
of the early type galaxies lie at low redshift with some intermediate type galaxies up to
redshifts of five. The majority of galaxies have template numbers between 5 and 6.
Particularly at low values, there is apparent redshift banding. The band at zero redshift
is due to galaxies with χ2 values that were still falling as they approached zero redshift,
indicating that the were not matched by any of the galaxy templates. The band at redshift
0.5 does correspond to an over density observed in the spectroscopic redshift analysis of
Cohen (2001) which also sees a small over density at a redshift of 1.
It is tempting to
interpret the banding in Figure 6 as sheets or nodes in large scale structure encountered by
the pencil beam of the NHDF but the redshift analysis in § 4 does not seem to warrant that
interpretation. This is particularly true for the band at 0.5 which may be due to the effect
discussed in § 4.1.1. Investigation of the significance of the banding will be deferred to a
later paper.
Most of the galaxies in Figure 6 fall above the lines of the L∗ (3.4 × 1010 L⊙) galaxy
-- 17 --
Fig. 6. -- The distribution of F160W AB 0.6′′ diameter aperture magnitudes versus photo-
metric redshift. The solid and dashed lines indicate the F160W aperture magnitude of an
early and late-type L∗ galaxy. The dash dot line shows the track of a late-type L∗ galaxy
with an extinction of E(B-V) equal to 0.2. L∗ is defined as a total luminosity of 3.4 × 1010
L⊙
-- 18 --
tracks. Past a redshift of ∼3 there are several galaxies that lie significantly below the L∗
galaxy line. This is expected since there certainly are galaxies brighter the L∗. It is legitimate
to ask whether these galaxies have non-physical parameters and whether their position in
the diagram is due to a bad fit with the models. Fig. 7 shows the fit between the observed
fluxes of the 9 galaxies that most exceed L∗ and the best fit fluxes from the photometric
redshift program. In all cases the fit is excellent, therefore, their position in the diagram
is not due to a bad fit. Table 3 gives a summary of the properties of the nine sources.
Although all but one galaxy has the luminosity of a Luminous InfraRed Galaxy (LIRG), the
low extinction does not re-emit a large enough fraction of the luminosity to put them in that
category. All but 2 galaxies fall in the top 50 most luminous galaxies in the field but none
of them are in the top 10. This leads to the conclusion that the physical properties of the
galaxies are realistic and that there is no reason to exclude them as legitimate sources based
on non-physical parameters.
Note that 398.0 and 410.0 are probably two parts of the same galaxy. Their photometric
redshifts of 3.04 and 2.96 are similar but not exactly the same. The spectroscopic redshift
for the the galaxy is 2.799, close to the photometric redshift. The spectroscopic redshift is
the redshift used in calculating the luminosity for source 410. If the spectroscopic redshift
for source 410 is also the proper redshift for source 398, the 398 luminosity should be reduced
by a factor of 0.74.
There are three sources without equivalent WFPC2 identifications. Source 398 is an
extension of source 410 identified as WFPC2 4-555.1. The closest WFPC2 source to 398 is
actually 4-555.11 which is 0.32′′ away, just beyond our association radius of 0.3′′. Similarly the
closest WFPC sources for 1630 and 1636 are 3-367.0 (0.33′′) and 3-839 (0.40′′). Although not
listed in Table 2,the identifications are listed in Table 3 since inspection of the images clearly
indicates that they are the proper associations. It is often the case that the F160W center
of a galaxy is offset from the optical position. The F160W position is more representative
of the underlying stellar distribution while the optical position measures rest frame UV star
formation activity.
6. Determination of the Star Formation Rates
As in TWS the SFR is determined from the broad band 1500 A flux relation given by
Madau, Pozzetti & Dickinson (1998).
UV1500 = 8.0 × 1027 · SF R(M⊙/yr) ergs second−1 Hz−1
(3)
-- 19 --
Fig. 7. -- The best fit fluxes matched to the observed fluxes for 9 galaxies that are significantly
brighter than the L* galaxy line. Luminosities are given in L⊙.
-- 20 --
The value of the UV flux is determined from the 1500 A flux of the unextincted template
found in the χ2 analysis averaged over a 200 A band centered on 1500 A. The reshift and the
normalization factor A (equation 2) set the absolute 1500 A flux. This procedure provides
the extinction corrected 1500 A flux for the SFR calculation. It is obviously a simplification
to assume a single extinction for a whole galaxy but it is clearly better than assuming a
single average extinction for all galaxies or no extinction at all. A much more extensive
discussion of the extinction is presented in TWS and it is not repeated here.
Rather than using the isophotal or total flux values returned by SE, the flux in each
band is computed as the sum of the flux from each pixel that the Szalay procedure and SE
determine is part of an object. The SFR computed with the A value determined from the
0.6′′ aperture is then multiplied by the ratio of the total flux to the aperture flux to give the
values reported in Table 2. The sum of the SFRs in Table 2 gives the observed SFR in the
appropriate redshift bins. At high redshift the observed SFR is significantly below the actual
SFR due to surface brightness dimming. The correction for this effect, using the extinction
corrected star formation rate per pixel is described in § 7.
7. Correction for Surface Brightness Dimming
It is well known that surface brightness dimming at high redshifts limits the number as
well as the extent of galaxies that are detected, creating errors in the measured SFR. In TWS
we used the star formation intensity distribution function (Lanzetta et al. 1999) to correct
for star formation missed due to surface brightness dimming. Lanzetta et al.
(2002) has
also used the distribution to correct for missed star formation but did not use a distribution
derived from extinction corrected UV fluxes.
7.1. Star Formation Intensity Distribution Function
The star formation intensity x is defined as the SFR in solar masses per year per proper
square kiloparsec. The intensity is calculated for each pixel that is part of a galaxy. Within a
given redshift interval the distribution function, h(x), is defined as the sum of all the proper
areas in a interval of specific intensity, divided by the interval and by the comoving volume in
cubic megaparsecs defined by the field and redshift interval (Lanzetta et al. 1999). Defined
in this manner the values of h(x) determine the star formation rate per cubic comoving
megaparsec through equation(4).
-- 21 --
sf r =Z ∞
0
xh(x)dx
(4)
The distribution functions for both extinction corrected and uncorrected star formation
intensities are shown in Figure 8.
-- 22 --
Fig. 8. -- Plots of the SFR intensity distribution for the Northern HDF. The abscissa is
-- 23 --
The top panels of Figure 8 show the 3 point smoothed fit to both the extinction corrected
and uncorrected distribution function at a redshift of 1. At higher redshifts the smoothed
curve is matched to the high intensity portion of the distribution by simply moving the
smoothed distribution up or down in the log-log plot. It is clear that the smoothed curve
for the uncorrected distribution does not match well at higher redshifts. This is discussed
further in § 7.3 but for now we will concentrate only on the extinction corrected distribution.
The deviation of the observed data from the smoothed distribution in the lower panels
shows the effect of surface brightness dimming. As in TWS we correct for surface brightness
dimming by integrating equation(4) over the observed data where the smoothed function
matches the data and then over the smoothed function (adjusted in height) when the ob-
servations fall below the curve. Note that the integral reaches 59% of its total value at
log x = −0.25 and 94% of its total value at log x = −1.25. § 7.3 discusses the validity of the
correction.
Table 4 shows the SFR at each stage of correction. The second column gives the rate
with no correction for either extinction or surface brightness dimming. The third column
shows the SFR after correction for extinction only and the fourth column shows the SFR after
correction for both extinction and surface brightness dimming. At low redshifts extinction is
the dominant correction while at high redshifts extinction and surface brightness dimming
are comparable. All of the calculations without extinction correction are made from the
output of the χ2 photometric redshift program with the extinction held to zero. In some
cases this alters the photometric redshift since the degree of freedom to redden a galaxy by
extinction is removed. This is the cause of the odd case at redshift 4 where the extinction
corrected SFR is less than the uncorrected SFR.
7.2. Selection Function
Lanzetta et al. (2002) introduced the concept of a selection function in conjunction with
the star formation intensity distribution function h(x). The selection function is the angular
area over which a parameter at a given depth can be detected. The maximum angular area
is the total area of the observation. At some depth of a parameter such as flux, the area goes
to zero when it becomes undetectable. The parameter used in (Lanzetta et al. 2002) and
here is the star formation intensity x. The main reason for introducing this concept is the
variance in sensitivity of the NICMOS Camera 3 in different regions of the detector. The
sensitivity varies by a factor of 3 from maximum to minimum.
The star formation intensity is a function of redshift, SED, extinction, and flux, there-
-- 24 --
fore, each pixel included in h(x) at a specific x may have a different selection function
depending on the values of the parameters. Given the criterion of 3.5 sigma detection in
a flux band for 3 contiguous pixels the selection function determines for each source pixel
the percentage of the field area where the source could be detected. The star formation
intensity distribution function program then corrects the area of each pixel by the inverse of
the selection function. The maximum allowed correction is a factor of 10 to prevent marginal
detections from dominating the distribution function. In practice only a small percentage of
pixels have selection functions less than 100%.
7.3. Validity of the Distribution and Comparison With Other Work
Initially in TWS the star formation intensity distribution was treated strictly empirically
with no physical motivation. The identical match of the smoothed distribution derived here
from largely independent data is further empirical confirmation of the distribution as is the
excellent match at high SFR intensities at higher redshifts. Thompson (2002), however,
provides a physical motivation by showing that the distribution is a natural consequence
of the Schmidt law, the majority of star formation in occurring in exponential disks, and a
Schechter distribution of galaxy masses. It also showed that effects of a smaller characteristic
mass and a smaller average radius of the exponential disk at higher redshift have opposite
effects on the shape of the distribution leading to a generally invariant distribution. Although
we lack detailed knowledge of the changes of characteristic mass and exponential disk radius
with redshift, the assumption of an invariant distribution appears to be justified. Even if the
surface brightness dimming corrections in Table 4 varied by 30% the error would be small
compared to the errors from other effects such as photometric error and large scale structure.
Lanzetta et al.
(2002) also utilized the star formation intensity distribution function
to correct for surface brightness dimming but found a SFR that steadily increases from
the present day to high values at a redshift of 8, in contrast to the essentially steady star
formation rate between a redshift of 1 and 6 as presented in § 8. Only the SFR in the
common ground of redshifts will be discussed in the following comments. The comparison
will be with the rates found from scaling the h(x) distribution vertically in Lanzetta et al.
(2002) since that is the technique used in this paper.
The critical difference between the technique applied here versus Lanzetta et al. (2002)
(2002)
is that this work uses the extinction corrected distribution where as Lanzetta et al.
(2002) refer to
use the distribution that is uncorrected for extinction. Lanzetta et al.
this rate as the unobscured rate. At a redshift of 1 they find a SFR of 0.03M⊙ per Mpc−3
whereas the value in this work is 0.3M⊙ per Mpc−3. This is simply understood as the
-- 25 --
difference between correcting for extinction and not correcting for extinction. At a redshift
of 4 Lanzetta et al. (2002) find a rate of 0.25M⊙ per Mpc−3 comparable to our value of 0.1.
At a redshift of 6 they find a value of 0.45 compared to our value of 0.2.
The difference between our relatively steady values and their increasing values is due to
two effects. As can be seen from Figure 8 and Lanzetta et al.
(2002) the bright end of the
distribution has a steeper slope in the uncorrected for extinction case than the corrected one.
This is due to extinction removing intrinsically bright pixels from the distribution. Second
at higher redshifts the average extinction of the sample becomes smaller due to surface
brightness dimming removing the highly extincted galaxies from the sample. Matching the
steep slope of the uncorrected distribution function to the continuously lower extinction
sample at higher redshifts results in an over correction and an apparent steadily increasing
star formation rate. Figure 8 shows this effect. The dimming correction in the right hand
column is significantly higher than in the left hand extinction corrected column.
8. Star Formation History
Initial work on the star formation history of the universe by (Madau et al. 1996) showed
a sharp rise in SFR from the present day to a peak at z between 1 and 2 then a decline to
lower rates at higher redshifts. Later work by Madau, Pozzetti & Dickinson (1998), which
applied a small correction for extinction, showed a shallower decline in SFR at high redshifts.
Subsequent work by Steidel et al.
(1999), Hopkins, Connolly, & Szalay (2000) and TWS
gave SFRs that were roughly constant at redshifts higher than 1. This is consistent with
the results presented here. The SFR values found in this paper are also consistent with the
SFRs found in the sub-millimeter observations of Barger et al.
(2000) at redshifts of 1-3
and 3-6. The concordance of results between the various studies is discused in § 11.
Figure 9 shows the SFR versus redshift corrected for the surface brightness dimming
and the selection function as described in § 7. The redshift binning unit is 1 centered on
integer redshifts. The lowest redshift included in the study is 0.5 and the highest is 6.5. The
NHDF statistics are not adequate at redshifts below 0.5 to determine an accurate SFR. In
that redshift range the number of sources is low and the percentage error in redshift is high,
leading to high SFR errors. The accuracy of large area ground based surveys is much greater
in this redshift range. Although Figure 6 shows galaxies at redshifts above 6.5 we do not
use them in this analysis.
-- 26 --
Fig. 9. -- Plots of the extinction and surface brightness dimming corrected SFR as a function
of redshift. The solid error bars indicate the photometric errors and the dashed error bars
indicate the uncertainty in the global SFR. The SFR from Yan et al.
(1999) indicated by
the diamond is for star formation in the range between a redshift of 0.8 to 2.0. The SFRs
from Barger et al. (2000) are indicated by the crosses and are for redshifts 1-3 and 3-6 with
a Λ = 2/3 universe.
-- 27 --
8.1. Star Formation Time History
The usual form of the 'Madau Diagram' plots SFR versus redshift since redshift is the
measured quantity and is not subject to the choice of cosmology. This form of the diagram
can be misleading in terms of when the majority of star formation occurs since there is
very little time at high redshifts. A more instructive plot is the star formation rate versus
age of the universe.
In the era of 'precision cosmology' one can dare to produce such a
plot with reasonable assumptions on the cosmology. For ages in the redshift range of 0.5
to 6.5 Figure 10 indicates the measured SFR at ages of the universe between 1.25 and 7.25
gigayears. The non integer age units are dictated by the ends of the redshift range.
Within the errors the SFR is roughly constant from an age of the universe between 1
and 7 gigayears at a rate significantly above the present day rate. The trend of the data
indicates an increase in the formation rate from 1 gigayear to 7 gigayears but the errors are
too large to validate this impression. Note that the binning of the objects is quite different
between the SFR versus redshift and versus age of the universe plots.
8.2. Error Analysis
The error analysis for this data set was carried out in the same manner as the analysis
performed in TWS where there is an extensive discussion of the analysis procedures. Table 5
gives the error sources from number statistics, photometric error and large scale structure.
The greatly increased number of sources in this data set has reduced number statistics to a
negligible error compared to other error sources.
8.2.1. Photometric Error Propagation
Photometric errors can propagate into errors in redshift, extinction, and SED. These
in turn create errors in the SFR. As in TWS, the photometric error analysis computes the
SFR in each redshift bin for 100 runs of the data with the flux values of each object altered
randomly by the gaussian distribution of the flux errors calculated for each object. The flux
error is the total flux error which includes both the random and systematic error shown in the
denominator of equation 1. Due to the variation of sensitivity across the NICMOS detector,
the errors in NICMOS flux are calculated individually for each galaxy. The distribution of
SFRs between the runs determines the SFR error level from photometric errors. The solid
line photometric error bars in Figure 9 are the 16% and 84% points of the SFR distribution
in quadrature with the 17% redshift error discussed in § 4.1. This procedure integrates the
-- 28 --
Fig. 10. -- Plots of the extinction and surface brightness dimming corrected SFR as a function
of age of the universe for the cosmology indicated in the figure. The solid error bars indicate
the photometric errors and the dashed error bars indicate the uncertainty in the global SFR.
-- 29 --
photometric error induced redshift, extinction and SED errors into a measured SFR error.
The procedure preserves the redshift, extinction, and SED calculated in each run, there-
fore, we can plot the individual errors in each parameter as a function of the F160W AB
magnitude. Figures 11, 12 and 13 show the histograms of redshift, extinction and template
SED errors in F160W AB magnitude bins ranging from 20 to 29 for all of the sources listed
in Table 2. In each case the unperturbed value is subtracted from the perturbed value. Note
that since the histograms include the more error prone objects in the 0.0-0.5 and 6.6-8.0
redshift ranges they are a conservative measure of the error.
The error histograms do not show any significant trend with magnitude. The basic
reason is that the systematic part of the error scales with flux so that the perturbations for
the strong fluxes are larger than for the weaker fluxes, accurately reflecting the observations.
As mentioned in § 4, the 10% systematic error is probably an overestimation so errors at
brighter magnitudes may be overestimations. If this is the case the error bars in Figure 9
may be exaggerated.
Both the redshift and extinction errors show a trend of being asymmetric around zero
error with an over abundance of negative errors showing larger redshifts and extinctions for
the unperturbed values than for the perturbed values. Both of these effects indicate larger
SFRs for the unperturbed versus the perturbed values. This trend is reflected in the overall
SFR errors shown in Figure 9 where the error bars extend further on the low side than the
high side for most of the redshift bins.
8.2.2. Large Scale Structure Error
As in TWS the two-point correlation function ξ(ro, γ) = (ro/r)γ as given by Peebles
(1980) is used to calculate the large scale structure error. The fractional error is taken as
σN /N where
and
σN /N =p1 + N × I2/√N
I2 =Z Z ξ(ro, γ)dV1dV2/V 2
(5)
(6)
For each redshift bin it is a very good approximation to consider the volume as a long
thin tube of square cross section with sides of (comoving) dimension D and (comoving) length
-- 30 --
Fig. 11. -- The redshift error histograms for all sources in Table 2. Negative values indicate
that the unperturbed redshift is higher than the perturbed redshift.
-- 31 --
Fig. 12. -- The E(B-V) error histograms for all sources in Table 2. Negative values indicate
that the unperturbed extinction is higher than the perturbed extinction.
-- 32 --
Fig. 13. -- The template number error histograms for all sources in Table 2. Negative values
indicate that the unperturbed template SED is later than the perturbed SED. Template
numbers run from 1, the earliest, to 6, the latest in increments of 0.1
-- 33 --
L. Then the expression for I2 has the form
I2 ≃ C(γ) × (ro/D)γ × (D/L)
(7)
The quantity L/D is simply the number of cubes of dimension D × D × D that can be
placed end to end in the tube of length L. The dimensionless coefficient C(γ) is a double
integral in which dV1 is taken over the unit cube and dV2 covers that same cube plus a
large number of cubes on either side of this unit cube. We have evaluated this coefficient
numerically which is vastly simplified by the large number of symmetries in this geometry.
(1998) we adopt γ = 1.8 and ro = 5h−1 Mpc and find
From the work of Adelberger et al.
C(γ = 1.8) = 8.22. We use the value of D and L for our adopted cosmology appropriate to
the center of each bin.
The calculation was made for a square field equal to one WFPC2 chip and then the errors
for the three fields were added in quadrature. The larger field and the increased photometric
error means that large scale structure is no longer the dominant but still an important error
in extrapolating the SFR history in the NHDF to the SFR history of the universe. The
dashed error bars in Figures 9 and 10 are the quadrature sum of the photometric and large
scale structure errors.
9. Magnitude Functions and Comparisons with Predictions
Ground based K magnitude functions have provided constraints on galaxy formation
models, eg. Kauffmann and Charlot
(1998). The much deeper F160W magnitude func-
tion determined here provides an even more powerful constraint on theoretical models. A
great advantage of this function is that is completely independent of models or cosmology.
Figure 14 shows the observed F160W magnitude function for the NHDF. The roll off at
magnitudes greater than 28 is due to surface brightness dimming and is not a true feature of
the magnitude function. In the range of validity, the magnitude functions from the WFPC2
F814W band (Williams et al. 1996) and the NICMOS F160W band are remarkably similar.
Kauffmann and Charlot (1998) proposed two tests to discriminate between hierarchi-
cal galaxy formation and pure luminosity evolution (PLE) of galaxies. Both models were
constrained to fit the present day K band luminosity function. Since the predictions are for
the K band, K band magnitudes were calculated from the observed 1.6 micron flux using
the SEDs found by the analysis. Any error in this transformation is very small compared to
the differences in the predictions between the hierarchical and PLE models. The first test
compares the number density of galaxies versus K magnitude at redshifts of 0.6, 1.0, and 2.0.
-- 34 --
Fig. 14. -- The NICMOS F160W (1.6 micron) magnitude function with the WFPC2 0.814
micron magnitude function superimposed.
-- 35 --
The PLE model predicts many more bright galaxies at each redshift than do the hierarchical
models. Since the predictions only go as faint as a K magnitude of 21.5 the number statistics
in the NHDF are small. The PLE model predicts a factor of 10 more bright galaxies than
are observed in the NHDF at redshifts of 1 and 2. The hierarchical predictions are within
the error bars of the observations.
A more decisive test from Kauffmann and Charlot (1998) is their prediction of the rela-
tive fraction of galaxies versus redshift in K magnitude bins. Figure 15 shows the predictions
and the observations. The first two magnitude bins have 1 and 0 galaxies in the NHDF but
the third K 19-21 bin has a significant number of galaxies. It clearly favors the hierarchical
model with the data and prediction being virtually indistinguishable. Although the predic-
tions end at a K magnitude of 21, the data for fainter magnitudes follows an extrapolation
of the hierarchical prediction at those magnitudes. The net results of the two tests are
consistent with the hierarchical prediction and inconsistent with the PLE prediction.
Our analysis determines the redshift, extinction, and intrinsic SED therefore a combi-
nation of the flux, redshift, and SED determines the total luminosity of the galaxy to the
accuracy of the individual parameters. The great depth of NHDF observations provides
further constraints on galaxy formation models by observing the evolution of the luminosity
function with time. Figure 16 shows the luminosity function at 6 different epochs, z = 1-6.
The roll off of each curve at low luminosities is simply due to galaxies falling below the
detection limit. To observational accuracy, the shapes of the luminosity functions at each
epoch appear remarkably similar. The low absolute values for the luminosity function at a
redshift of 5 may be due to a void in the NHDF at that redshift.
It should be noted that even though the luminosity function appears invariant with red-
shift to z = 6, this does not mean that the mass function is also invariant. Very young galaxies
at a redshift of 6 require much less mass to achieve the same luminosity as old, present day
galaxies. In fact, the invariant luminosity function appears to favor the hierarchical model
of galaxy assembly.
10.
Infrared and Sub-mm Constraints
The observed infrared and sub-mm backgrounds place constraints on the extinction
values found in our χ2 analysis. If the extincted luminosity is significantly different from the
observed infrared background levels then the extinction values may be considered suspect. A
calculated background flux that is much higher than the observed background would indicate
an over correction for extinction and a higher SFR than allowed by the infrared and sub-mm
-- 36 --
Fig. 15. -- Plots of the relative fraction of galaxies with redshifts greater than z in 7 different
magnitude bins. Predictions only exist for the first three bins. The first two bins have only
1 and 0 observed galaxies respectively.
-- 37 --
Fig. 16. -- The evolution of the 1.6 micron luminosity function with redshift.
-- 38 --
constraints. On the other hand, if the extincted luminosity is much lower than the infrared
background levels it would be a sign that our analysis has missed a large part of the highly
obscured star formation activity. In addition the calculated sub-mm luminosity of individual
sources should not exceed the detections and limits set in the NHDF by Hughes et al. (1998).
Note that although comparison with the observed background and emission from individual
objects is a constraint on the extinction values, agreement is not a validation of the values.
The difference between E(B-V) values of 1 and 2 is a large extinction error but both predict
essentially the same amount of re-emitted flux since they both remove almost all of the UV
and optical luminosity.
Both Calzetti et al. (2000) and Adelberger and Steidel (2000) have explored the errors
in transforming between extinctions derived from the optical flux and the observed infrared
emission. Calzetti et al. (2000) estimates that the error is a factor of 2 for individual sources
which reduces to 20% for a group of about 50 galaxies. Adelberger and Steidel (2000) cite
errors of 0.2 - 0.3 dex (1.6 - 2.0) and ascribe the same range in SFR error. The error bars in
Figures 9 and 10 are well within these limits, particularly since the points represent hundreds
of galaxies rather than individual galaxies.
10.1. Predicted and Observed Backgrounds
The fraction of luminosity removed by extinction is easily calculated for each SED from
the Calzetti obscuration law. This luminosity is assigned to a dust emission SED. Dust SEDs
vary significantly (Rigopoulou et al. 1999) but there is no easy way to predict which SED
is appropriate from just the optical and near-IR observations. For consistency a single dust
SED is used from Figure 1 of Aussel et al. (1999) for the PAH features joined with the Arp
220 ULIRG spectrum given in Figure 4 of Rowan-Robinson and Efstathiou (1993). The
SCUBA 850 µm flux as well as the ISO 6 and 15 µm fluxes listed in Table 2 are calculated
from the SED, E(B-V), redshift and luminosity.
Integration of all of the predicted 850 µm fluxes gives a 850 µm background of 3.9×10−10
watts m−2 steradian−1. The 850 µm flux is consistent with the value of 5 ± 2 × 10−10 from
Blain et al. (1999) and 4 × 10−10 for COBE (Fixsen et al. 1998). The agreement with the
probably more accurate COBE flux is remarkable but it should be interpreted only that the
predicted sub-mm flux does not indicate an error in the derived parameters
Integration of the observed 1.6 µm fluxes gives a 1.6 µm discrete source background of
7.0 × 10−9 watts m−2 steradian−1. The previous 1.6 µm background in the Deep NICMOS
NHDF from TWS was 6 × 10−9. The F160W background found here is significantly less
-- 39 --
than the J (54 × 10−9) and K (27.8 × 10−9) backgrounds found by Cambresy et al. (2001).
It is also less than but closer to the K band background of 12.4 × 10−9 found by Wright and
Reese (2000). Our background subtraction technique will remove any component of the
true background not due to detected discrete sources.
10.2.
Individual Sources
Although it is not the primary purpose of the infrared flux calculation, the calculated
fluxes of individual sources can be compared with observations and observational limits.
10.2.1. ULIRGs
We define an Ultra Luminous Infrared Galaxy (ULIRG) as a galaxy with greater than
1012 L⊙ total luminosity with half or greater of the luminosity emitted in the mid and far
infrared. One of the surprising aspects of TWS was the observation of two ULIRGs in the
small 50′′x50′′ field of the Deep NICMOS observations. They are sources 439 and 800 in the
present catalog. Source 800 retains it ULIRG status in this analysis. Source 800 may also
contain an AGN component. Although not listed in their detection catalog, the Chandra
X-Ray map of (Brandt et al. 2001b) clearly shows a source at the position of source 800.
Source 439 (166 in TWS), however, falls to LIRG status with a luminosity of 1.95× 1011 L⊙.
The primary reason is that the best fit in this analysis is a template 5.7 and E(B-V) = 0.5
fit as compared to a template 5.9 and E(B-V) = 1.0 fit in TWS. Although the differences in
predicted fluxes is small, the luminosity difference between the two fits is large. Given the
higher signal to noise, the parameters found in TWS are more likely correct.
There are two very high redshift galaxies, 22 and 396, with ULIRG level luminosities but
their redshifts of 8.0 and 7.44 put them beyond the galaxies considered in this paper. The
validity of these source will be left to another study. An additional two candidates included
in this analysis are sources 475 and 1364 at redshifts of 6.24 and 4.88 respectively. These do
not correspond to any WFPC2 NHDF source and are both outside of the field of the Deep
NICMOS image. Source 1364 has a small amount of F814W flux while 475 has no detectable
flux in any of the optical bands. Both galaxies have significant F110W and F160W fluxes
with very red F160W-F110W colors. The χ2 plots show two very broad minima at redshifts
between 1 and 2 and at the higher redshifts where the fit is the best. Both of the high
redshift minima are extremely broad, covering a redshift range of approximately 2 around
the selected redshift. Given the nature of the χ2 distribution, the ULIRG status of these
-- 40 --
sources is extremely uncertain.
With no reliable detection of additional ULIRGs in this study we can throw no light
on the question of whether ULIRGs were much more prevalent in the past, reaching a peak
near a redshift of two. If the two possible ULIRGs above were at a redshift of 2 rather than
their high redshifts, their luminosities would be considerably sub-ULIRG. The existence of
any ULIRGs in the area of the NHDF is still highly improbable, however, given the local
space density of ULIRGs. Resolution of this question will require deep imaging in additional
fields.
10.2.2. SCUBA Sources
Probably the most enigmatic sources in the NHDF are the sub-mm sources detected by
SCUBA (Hughes et al. 1998). Of the 5 detected HDF sources, source 2 actually lies just
outside of the NHDF field and source 3 is in the PC chip which was not analyzed in this
work. The remaining sources 1, 4, and 5 are discussed below.
VLA observations by Downes et al.
(1999) indicate that the galaxy corresponding to
HDF850.1 is either WFPC2 3-586 or 3-593 if it is associated with an optically visible galaxy.
It could also be an optically obscured galaxy lying between the two galaxies. Examination of
the F160W image does not reveal an optically undetected galaxy lying between the sources,
however, the images of the two galaxies slightly overlap at a level below the threshold set
for positive detection in § 3. WFPC2 3-586 corresponds to source 1247 in Table 2. It is
an elliptical galaxy (Template 1.2) with no detected extinction at a redshift of 0.88. The
luminosity is 1.5 × 1010 L⊙ and it is an unlikely candidate for HDF850.1.
WFPC2 3-593.0 and 3-593.2 correspond to source 1212 in Table 2. This is a very late
galaxy (Template 6.0) with an E(B-V) of 0.4 at a redshift of 1.84. It has a luminosity of
1.76× 1011 but the predicted 850 µm flux is only 0.16 mJy, below the detected flux of 7 mJy.
This is a case where the available parameter space of the analysis may have been inadequate
to identify this object as the SCUBA source. The analysis picked the hottest SED with a
relatively high E(B-V). If an even hotter SED were available it might have given a better
fit with a correspondingly higher E(B-V). The combination of the hotter SED and higher
E(B-V) could dramaticly increase the expected 850 µm flux. Of the two visible sources 3-593
(1212) is by far the better candidate for HDF850.1.
An alternative source is 1277 which lies closer than 1′′ to the position of HDF850.1 given
by Hughes et al. (1998). It is barely visible in the F814W band but is a strong source in the
F110W and F160W bands. It is a high redshift object (z = 4.8) with a extinction of E(B-V)
-- 41 --
= 0.3 and a very late template of 5.9. Source 1277 has a luminosity of 1.49 × 1011 L⊙ and a
predicted 850 µm flux of 0.125 mJy which is also below the observed flux. The location of
the VLA source however makes this identification much less likely.
Dunlop et al. (2002) have proposed an identification of HDF850.1 as a source located
between 3-586.0 and 3-593.1 that is only visible after subtraction of 3-586 in ground based
K band images. They also quote a possible detection after subtraction in the F160W images
used here. In this scenario the flux of the sub-mm source is lensed with an amplification
factor of 3. If this is the true source, it has been missed in the analysis performed here and
would only be visible through special processing which is beyond the scope of this paper.
HDF850.4 and HDF850.5 lie very close to each other with a separation of about 12′′.
The sources are in fact blended and must be deconvolved (Serjeant et al. 2002). The only
nearby source with strong calculated 850 µm flux is source 1108 which lies almost midway
between the sources. Its calculated 850 µm flux is 1.36 mJy close to the observed fluxes of
2.3 mJy and 2.1 mJy for HDF850.4 and HDF850.5. The source has a spectroscopic redshift
of 0.19, an E(B-V) of 0.6 and a luminosity of 1.55 × 1011L⊙.
If the shift in position for
HDF850.1 to lie on top of 3-593 is applied to HDF850.4 and HDF850.5, HDF850.5 would
be within 4′′ of 1108. In this case sources 896 and 905 could be identified with HDF850.4.
Their combined predicted flux is equal to 0.89 mJy which is near to the observed flux of 2.3
mJy for HDF850.4.
If the stated detection limit of 2 mJy is taken for the SCUBA observations in the NHDF
(Hughes et al. 1998) there is one source in Table 2 that has a predicted 850 µm flux above
the limit but was not detected by the SCUBA observations. The source is the ULIRG,
source 800, discussed in § 10.2.1. It corresponds to source 277.211 in TWS where its flux
was predicted to be 1.5 mJy rather the 3.46 mJy found here. The present analysis found
the same redshift, E(B-V) and template as TWS, but the observed F110W and F160W
magnitudes were brighter. If the ULIRG contains a AGN component, as is indicated by the
x-ray component, the variation might be real but is of surprising magnitude.
VLA 3651+1221
(1999) observed a second source, VLA 3651+1221, very near the elliptical
Downes et al.
galaxy 3-659.1 which is source 1193 in Table 2. The F160W image shown in Figure 17 clearly
shows a strong source at the position of the VLA detection that is very faint or absent in
the F814W image.
It is clearly a very red source. The correspondence was first noticed
by Dickinson et al (2000) who claim that this is the second reddest source in the NHDF.
Papovich, Dickinson and Ferguson (2001) point out that the object is also an x-ray source
-- 42 --
(Hornschemeier et al. 2000; Brandt et al. 2001a) and therefore probably an AGN. The SE
source extraction program linked the source with the elliptical galaxy in producing Table 2,
which resulted in an identification of a very late galaxy with a high extinction. This is clearly
an artifact of the superposition of fluxes from two very different types of galaxies.
11. Concordance of Star Formation Rates
There has been significant debate (Barger et al. 2000; Blain et al. 1999) on whether
optically based studies miss the majority of the star formation in the universe because it is
hidden in highly obscured galaxies. The SFRs in Figure 9 for the optical, optical plus near
infrared, and submillimeter, however, are all in agreement. On the face of it this would seem
to indicate that there is no large missing component of star formation. On the other hand
the agreement may be fortuitous or the definitions may be misleading. This section examines
the various studies to see what is missing and how the agreement has been achieved.
The only inconsistent measurements in Figure 9 are the measurements of Madau (1999)
which show a fall off of star formation at a redshift of 4 and generally lower SFR at all
redshifts. The SFRs in Madau (1999) are derived from the numbers presented in Madau,
Pozzetti & Dickinson (1998), (note erroneous reference in Madau (1999)) which do not
appear to be corrected for surface brightness dimming. The lack of dimming correction
appears to be the principal reason for the fall off at high redshifts. The average extinction
correction in Madau (1999) is A1500 = 1.2, which is an underestimate according to this work
and accounts for the lower values at redshifts 1 to 2.
The extinction corrected SFRs at redshifts 3 and 4 from Steidel et al.
(1999), cor-
rected for the cosmology adopted here, are coincident with the SFRs found in this work.
The extinction corrected values were found by multiplying the uncorrected SFRs by 4.7,
corresponding to an E(B-V) = 0.15 (Steidel et al. 1999). The extinction correction value in
Steidel et al. (1999) is well motivated and is interpreted there as the proper correction to the
UV flux from observed galaxies in the study. It could also be interpreted as a lower E(B-V)
correction to the observed galaxies plus a factor that accounts for galaxies not observed due
to very high extinction. Without determining an extinction for each galaxy in the sample
it is not possible to separate the two interpretations. It still leaves the possibility that the
actual average extinction to the observed galaxies is less than E(B-V)= 0.15 and that there
are other undetected galaxies that contribute to the SFRs observed by Barger et al. (2000).
Note that the comparison with the Steidel et al.
(2000) Figure 11
is to the uncorrected rate and a Λ = 0 universe.
(1999) in Barger et al.
-- 43 --
Fig. 17. -- The extremely red object associated with VLA 3651+1221.
-- 44 --
The SFRs in this work are both extinction corrected on an individual galaxy basis
and corrected for surface brightness dimming. The extinction correction at redshifts 2 and
4 are 4.9 and 6.1, near the range of the Steidel et al.
(1999) correction. The surface
brightness dimming corrections are 2.0 in both cases, therefore, the extinction correction is
the dominant correction rather than galaxies or parts of galaxies undetected due to surface
brightness dimming. This would appear to indicate that the observed sub-mm SFR can
be accounted for by galaxies that are observable in deep HST optical and near infrared
observations if not necessarily optical ground based observations.
It can also be argued that the present work has been unsuccessful in identifying the
3 observed sub-mm sources in the fields of the 3 WF chips. There is no doubt that there
can be luminous galaxies that are so highly extincted that even deep HST observations
will miss them. At a redshift of 3 the F160W NICMOS band is measuring the 4000 A
flux which can be severely extincted.
It is also possible that the optical depth that can
be reached with the F160W filter is less than the total optical depth of the galaxy so that
the present analysis determines a lower luminosity and extinction than is actually present
in the galaxy. Another possibility is that the assumed dust SED is not appropriate for the
bright sub-mm sources. The luminosity in the 850 µm SCUBA band is a small fraction of
the total dust luminosity, therefore, small changes in the dust SED can greatly change the
850 µm luminosity without significantly affecting the total dust luminosity. Perhaps the
bright sub-mm sources have colder dust SEDs than the Arp 220 dust SED used here. The
bottom line is that on an individual source level there are many uncertainties that can lead
to erroneous sub-mm predictions in the present analysis. It should be noted, however, that
the sub-mm predictions are only intended to serve as a check on the total sub-mm power
not as a individual source predictor.
In assessing the effect of underpredicting the bright sub-mm sources it should be noted
that the correction factor used in Barger et al. (2000) to account for the fainter undetected
sources is a factor of 11,therefore, the bright sub-mm sources account for less than 10%
of the SFR. Based on the size of the correction it might be argued that the majority of
star formation is also missed by the sub-mm observations. The correction to the sub-mm
observations depends on the assumed luminosity function whereas the correction to the
optical and near IR observations depends on the assumed extinction law. Even if all of the
observed sub-mm sources are missed by optical and near-IR studies, the error in the SFR is
small. It can be argued that at the depth of the HST observations only the most obscured
sources are missed and that the bulk of the galaxies that produce the sub-mm emission and
the majority of the SFR are detected and on the average properly accounted for. Within
the caveats discussed above it appears that although some of the extreme sub-mm sources
are missed by the optical and near infrared observations, the total SFR derived from optical,
-- 45 --
optical and near infrared, and sub-mm observations are in agreement indicating that the
majority of the SFR is accounted for in the three types of observations. All of these methods
however, have correction factors on the order of a factor of 10 based on assumptions that
still need to be confirmed.
12. Summary
Analysis of HST archival data from WFPC2 and NICMOS in the NHDF yields photo-
metric redshifts, extinctions and SEDs for almost 2000 galaxies. From this data the extinc-
tion corrected SFR for each galaxy was determined. After correction for surface brightness
dimming and the variable sensitivity over the NICMOS detector arrays, the star formation
history of the NHDF from a redshift 1 to 6 shows a roughly constant SFR. Optical studies
of Lyman break galaxies and sub-mm observations produce SFRs that are consistent with
the results in this study. The sub-mm background predicted from our analysis is consistent
with the observed background, indicating that the analysis has not missed a significant star
forming component due to high extinction. The inability to accurately predict the fluxes
and locations of the observed NHDF sub-mm sources does indicate that some bright sub-mm
sources are certainly missed by this work but that these are not the sources that contribute
the majority of star formation in the field.
13. Acknowledgements
The author would like to acknowledge the significant contributions to this work by
Ray Weymann and Lisa Storrie-Lombardi in the initial data reduction and analysis as well
as their contributions in establishing the methodology which was carried over from TWS.
The author would also like to thank Mark Dickinson and all of his group who planned
and executed the NICMOS observations of the entire Northern HDF in a General Observer
program. This work is supported in part by NASA grant NAG 5-10843. This article is
based on observations with the NASA/ESA Hubble Space Telescope, obtained at the Space
Telescope Science Institute, which is operated by the Association of Universities for Research
in Astronomy under NASA contract NAS5-26555.
-- 46 --
REFERENCES
Adelberger, K. L., Steidel, C. C., Giavalisco, M., Dickinson, M., Pettini, M., & Kellogg, M.
1998, ApJ, 505, 18
Adelberger, K. L. & Steidel, C. C. 2000, ApJ, 544, 218
Aussel, H., et al. 1999, a, 342, 313
Barger, A.J., Cowie, L.L., & Richards, E.A. 2000, AJ, 119, 2092
Bertin, E., & Arnouts, S. 1996, A&A, 117, 393
Blain et al. 1999, ApJ, 512, L87
Brandt, W. M., et al. 2001, AJ, 112, 1
Brandt, W. M., et al. 20001, AJ, 122, 2801
Bruzual, G. and Charlot, S. 1996, GISSEL96
Calzetti, D., Kinney, A. L., & Storchi-Bergmann, T. 1994, ApJ, 429, 582
Calzetti, D. 1999 Private communication of digital data.
Calzetti, D. 1997, AJ, 113, 162
Calzetti, D., Armus, L., Bohlin, R.C., Kinney, A.L., Koornneef, J., & Storchi-Bergmann, T.
1999,ApJ, 533, 682
Calzetti et al. 2000, ApJ, 533, 682
Cambresy, L. et al. 2001, ApJ, 555, 563
Cohen, J.G., Hogg, D.W., Blandford, R.,Cowie, L.L., Hu, E., Songaila, A., Shopbell, P., &
Richberg, K. 2000, ApJ, 538, 29
Cohen, J.G. 2001, AJ, 121, 2895
Coleman, G.D., Wu, C-C, & Weedman, D.W., 1980 ApJS, 43, 393
Dickinson, M. 2000, Phil.Trans. Royal Soc. Lond. A., 358, 2001
Dickinson M. et al. 2000, ApJ, 531, 624
Downes, D. et al. 1999, A&A, 347, 809
-- 47 --
Dunlop, J. S. et al. 2002, astro-ph/0205489
Fern´andez-Soto, A., Lanzetta, K.M., & Yahil, A. 1999, ApJ, 513, 34
Fixsen et al. 1998, ApJ, 508, 123
Hopkins, A. M., Connolly, A.J., & Szalay, A. S. 2000, AJ, 120, 2843
Hornschemeier, A. E., et al. 2000, ApJ, 541, 49
Hughes, D., et al. 1998, Nature, 394, 241
Kauffmann, G. & Charlot, S. 1998, MNRAS, 297, L23
Lanzetta, K.M., Yahata, N., Pascarelle, S., Chen, H.-W. & Fern´andez-Soto, A. 2002, ApJ,
570, 492
Lanzetta, K.M., Chen, H-W., Fernandez-Soto, A., Pascarelle, S., Puetter, R., Yahata, N.,
and Yahil, A. 1999, ASP Conf. Ser. 191, Photometric Redshifts and High Redshift
Galaxies, ed. R. J. Weymann, L.J. Storrie-Lombardi, M. Sawicki, and R. J. Brunner,
(San Francisco ASP), 223.
Lanzetta, K.M., Yahil,A., & Fern´andez-Soto, A. 1996, Nature, 381, 759
Lilly, S. J., Le F`evre O., Hammer, F., & Crampton, D. 1996, ApJ, 460, L1
Lytle, D., Stobie, E., Ferro, A. & Barg, I. 1999, in ASP Conf. Ser. 172, Astronomical Data
Analysis Software and Systems VIII, ed. D. Mehringer, R. Plante, & D. Roberts, (San
Francisco; ASP), 445
Madau, P., Ferguson, H.C., Dickinson, M. E.,Giavalisco, M., Steidel, C. C. & Fruchter, A.
1996, MNRAS, 283, 1388
Madau, P., Pozzetti, L., & Dickinson, M. 1998, ApJ, 498, 106
Madau, P. 1999, AIP Conference Proceedings 470, After the Dark Ages: When Galaxies were
Young (The Universe at 2 < z < 5), ed. Stephen S. Holt & Eric P. Smith, (Woodbury,
New York AIP), 299
Papovich, C., Dickinson, M. & Ferguson, H. C. 2001, ApJ, 559, 620
Peebles, P.J.E. 1980 The Large Scale Structure of the Universe, Princeton University Press,
Princeton New Jersey
Rigopoulou, D., et al. 1999, AJ, 118, 2645
-- 48 --
Rowan-Robinson, M., & Efstathiou, A. 1993, MNRAS, 263, 675
Serjeant, S. et al. 2002, astro-ph/0201502
Steidel, C. C., Adelberger, K. L., Giavalisco, M., Dickinson, M., & Pettini, M. 1999, ApJ,
519, 1
Szalay, A.S., Connolly, A.J., & Szokoly, G.P. 1999, AJ, 117, 68
Thompson et al. 1999, AJ, 117, 7
Thompson, R.I., Weymann, R.J. & Storrie-Lombardi, L.J. 2001, ApJ, 546, 694 (TWS)
Thompson, R. I., ApJ, 581, L85
Wang,Y., Bachall, N., & Turner, E.L. 1998, AJ, 116, 2081
Williams, R.E. et al. 1996, AJ, 112, 1335
Wright, E.L. & Reese, E.D. 2000, ApJ, 545, 43
Yan, L., McCarthy, P.J., Freudling, W., Teplitz, H.I., Malumuth, E.M., Weymann, R. J., &
Malkan, M.A. 1999, ApJ, 519, L47
This preprint was prepared with the AAS LATEX macros v5.0.
-- 49 --
Table 1. Galaxies with Redshift Errors Greater than 0.5.
ID WFPC-ID Phot. Z
Spec. z Template
E(B-V) Qual. a Sp. b
fail c
6
13
84
108
339
369
524
625
1226
1656
1762
4-916
4-928
4-878
· · ·
4-445
· · ·
2-251
2-201
2-531
2-982
· · ·
0.08
0.48
0.0
1.84
1.84
1.92
0.24
1.84
0.56
0.56
1.12
0.904
1.015
0.892
0.584
2.5
2.801
0.962
1.313
1.087
1.147
0.47
6.0
6.0
3.0
5.2
5.8
5.9
4.3
5.0
5.8
5.7
3.5
0.6
1.0
0.04
0.5
0.3
0.3
0.4
0
0.5
0.4
0.0
9
4
4
3
2
3
6
1
3
1
1
A
E
E
A
EA
A
Q
E
A
E
E
P
P
S,P
S
S
· · ·
· · ·
S,P
F,P
F
P
aThis is the quality of the spectrum as noted in Cohen et al.
(2000) where 1 is the
highest 11 is the lowest cited in table 2b of that work.
bSp is the spectral type listed in Cohen et al.
(2000). E is an emission line galaxy, A
is an absorption line galaxy, EA is a galaxy displaying both emission and absorption and
Q is a broad absorption line galaxy
cFail is the probable cause of failure. P is hitting a boundary in parameter space, S
is the probable superposition of two galaxies, and F is the problem associated with z =
0.56 discussed in the text.
-- 50 --
The remainder of the 1927 table entries are available in the electronic version of the
Astrophysical Journal publication.
Table 2. Listing of measured quantities
NICMOS WFPC
z
E(B-V)
ID
ID
1
2
3
4
5
6
7
8
9
10
4-951.0
· · ·
· · ·
· · ·
· · ·
4-916.0
4-922.0
4-952.0
4-930.0
4-953.0
0.00
2.88
2.56
2.56
2.80
0.90
3.20
3.52
0.48
2.88
0.50
0.00
0.00
0.10
0.00
0.60
0.20
0.20
0.30
0.00
SFR
M⊙
yr−1
2.831
0.258
0.131
0.691
1.984
0.182
2.937
5.092
0.662
0.587
Lum.
L⊙
frac.a
0.00E+00
3.64E+09
2.15E+09
3.94E+09
7.36E+09
6.34E+08
1.18E+10
2.15E+10
2.46E+09
2.62E+09
0.86
0.00
0.00
0.39
0.00
0.95
0.53
0.51
0.75
0.00
ISO6
6 µm
flux
mJy
0.00E+00
0.00E+00
0.00E+00
5.16E-06
0.00E+00
1.90E-05
0.00E+00
0.00E+00
2.63E-04
0.00E+00
ISO15
15 µm
flux
mJy
0.00E+00
0.00E+00
0.00E+00
7.82E-06
0.00E+00
1.61E-04
2.10E-05
3.00E-05
1.76E-03
0.00E+00
SCUBA
850 µm
flux
mJy
0.00E+00
0.00E+00
0.00E+00
1.61E-03
0.00E+00
8.75E-04
6.76E-03
1.29E-02
4.70E-03
0.00E+00
T b
2
χ
Tot.c
mag
Ap.d
mag
5.6
3.5
3.3
5.1
5.8
6.0
5.6
5.5
5.8
5.4
3.0
18.
19.
5.6
19.
11.
2.6
4.6
8.1
2.2
26.4
28.2
28.2
28.7
27.1
22.5
27.8
27.4
24.2
27.5
27.4
27.2
27.3
27.2
27.1
24.6
27.7
27.3
25.7
27.9
RA
h
12.
m
36.
38.23
38.24
38.32
38.32
38.49
38.59
38.60
38.77
38.79
38.83
DEC
◦
+62.
12:28.3
12:31.7
12:34.2
12:32.2
12:35.5
12:33.8
12:29.1
12:18.8
12:25.9
12:17.9
aThis is the fraction of the luminosity removed by extinction and re-emitted in the mid and far infrared
bThe selected template number between 1.0 (early-cold) and 6.0 (late-hot).
cThe total AB magnitude in the F160W filter.
dThe F160W magnitude in an 0.6′′ diameter aperture.
--
5
1
--
-- 52 --
Table 3. Properties of galaxies with F160W magnitudes significantly brighter than
predicted L* magnitudes.
ID WFPC-ID Redshift
F160W AB mag. E(B-V) Template
Lum.
1636.0
398.0
410.0
81.0
1604.0
1630.0
1409.0
1179.0
812.0
3-839.0
4-555.11
4-555.1
4-713
3-82
3-367
2-591
2-578
4-169
5.2
3.04
2.799
4.08
4.32
4.48
2.96
3.92
5.04
23.8
24.4
23.8
23.6
24.3
24.8
24.4
25.0
25.2
0.02
0.08
0.20
0.00
0.10
0.06
0.00
0.00
0.10
5.1
5.6
5.7
3.9
3.6
3.5
5.2
5.2
3.9
5.72E11
2.7E11
6.09E11
4.96E11
6.5E11
3.66E11
8.84E10
1.09E11
3.38E11
-- 53 --
Table 4. SFR corrections for extinction and surface brightness dimming
z Uncor. Ext. Cor. Ext.& Dim. % Lum. a
Missing
Cor. SFR
SFR
SFR
1
2
3
4
5
6
6.8E-3
3.7E-2
5.0E-2
1.0E-1
1.8E-2
3.5E-2
2.8E-1
1.8E-1
8.9E-2
6.1E-1
6.5E-2
8.1E-2
2.8E-1
3.7E-1
9.6E-2
1.2E-1
2.3E-1
3.0E-1
0%
50%
8%
50%
72%
73%
aThis is the percentage of missed luminosity be-
tween the extinction corrected SFR and the extinc-
tion corrected SFR also corrected for surface bright-
ness dimming.
-- 54 --
Table 5. Sources of Numerical Variances
Statistic
z = 1
z = 2
z = 3
z = 4
z = 5
z = 6
Number of Sources
1/√N
I2
σN /N
16% Confidence fraction
84% Confidence fraction
495
0.045
0.16
0.09
0.47
0.37
370
0.052
0.11
0.11
0.48
0.21
339
0.058
0.13
0.13
0.21
0.69
160
0.079
0.16
0.16
0.27
1.2
45
0.15
0.19
0.19
0.64
1.1
55
0.13
0.21
0.21
0.82
0.13
|
astro-ph/0412097 | 1 | 0412 | 2004-12-03T22:36:30 | Stars in the Hubble Ultra Deep Field | [
"astro-ph"
] | We identified 46 unresolved source candidates in the Hubble Ultra Deep Field, down to i775 = 29.5. Unresolved objects were identified using a parameter S, which measures the deviation from the curve-of-growth of a point source. Extensive testing of this parameter was carried out, including the effects of decreasing signal-to-noise and of the apparent motions of stars, which demonstrated that stars brighter than i775 = 27.0 could be robustly identified. Low resolution grism spectra of the 28 objects brighter than i775 = 27.0 identify 18 M and later stellar type dwarfs, 2 candidate L-dwarfs, 2 QSOs, and 4 white dwarfs. Using the observed population of dwarfs with spectral type M4 or later, we derive a Galactic disk scale height of 400 \pm 100 pc for M and L stars. The local white dwarf density is computed to be as high as (1.1 \pm 0.3) x10^(-2) stars/pc^3. Based on observations taken 73 days apart, we determined that no object in the field has a proper motion larger than 0.027"/year (3 sigma detection limit). No high velocity white dwarfs were identified in the HUDF, and all four candidates appear more likely to be part of the Galactic thick disk. The lack of detected halo white dwarfs implies that, if the dark matter halo is 12 Gyr old, white dwarfs account for less than 10% of the dark matter halo mass. | astro-ph | astro-ph |
Stars in the Hubble Ultra Deep Field
N. Pirzkal1, K. C. Sahu1, A. Burgasser2, L. A. Moustakas1, C. Xu1, S. Malhotra1, J. E.
Rhoads1, A. M. Koekemoer1, E. P. Nelan1, R. A. Windhorst3, N. Panagia1, C. Gronwall4,
A. Pasquali5, J. R. Walsh6
ABSTRACT
We identified 46 unresolved source candidates in the Hubble Ultra Deep Field,
down to i775 = 29.5. Unresolved objects were identified using a parameter S,
which measures the deviation from the curve-of-growth of a point source. Exten-
sive testing of this parameter was carried out, including the effects of decreasing
signal-to-noise and of the apparent motions of stars, which demonstrated that
stars brighter than i775 = 27.0 could be robustly identified. Low resolution grism
spectra of the 28 objects brighter than i775 = 27.0 identify 18 M and later stellar
type dwarfs, 2 candidate L-dwarfs, 2 QSOs, and 4 white dwarfs. Using the ob-
served population of dwarfs with spectral type M4 or later, we derive a Galactic
disk scale height of 400 ±100 pc for M and L stars. The local white dwarf density
is computed to be as high as (1.1 ± 0.3) × 10−2 stars/pc3. Based on observations
taken 73 days apart, we determined that no object in the field has a proper mo-
tion larger than 0.027"/year (3σ detection limit). No high velocity white dwarfs
were identified in the HUDF, and all four candidates appear more likely to be
part of the Galactic thick disk. The lack of detected halo white dwarfs implies
that, if the dark matter halo is 12 Gyr old, white dwarfs account for less than
10% of the dark matter halo mass.
Subject headings: Galaxy: stellar content, Galaxy: structure, Galaxy: disk,
white dwarfs, stars: late-type
1Space Telescope Science Institute, 3700 San Martin Drive, Baltimore, MD21218, USA
2Department of Astrophysics, Division of Physical Sciences, American Museum of Natural History, Cen-
tral Park West at 79th Street, New York, NY 10024, USA; [email protected]
3Dept. of Physics & Astronomy, Arizona State University, Street: Tyler Mall PSF-470, P.O. Box 871504,
Tempe, AZ 85287-1504, USA
4Department of Astronomy & Astrophysics, Pennsylvania State University, 525 Davey Laboratory, Uni-
versity Park, PA 16802
5Institute of Astronomy, ETH Honggerberg, 8093 Zurich, Switzerland
6ESO/ST-ECF, Karl-Schwarschild-Strasse 2, D-85748, Garching bei Munchen, Germany
-- 2 --
1.
Introduction
Studying the stellar content of our galaxy down to the faintest possible magnitudes is
necessary to study the structures of the Galactic disk and population II halo (referred to
as Galactic halo thereafter) and stellar population (M´endez et al. 1996). The nature of the
Galactic dark matter halo has in addition been the subject of some debate since the MACHO
project (Alcock et al. 2000) observed lensing events in the direction of the Large Magellanic
Cloud (LMC) and claimed that 20% of the dark matter halo mass could be in the form of
low mass objects with masses smaller than 1.0M⊙ (Alcock et al. 2000). While there are
alternative explanations for the observed lensing events (Sahu 1994; Sahu 2003; Gates &
Gyuk 2001, and references therein), the detection of faint, blue unresolved objects with high
proper motion has been used as evidence for the existence of a large population of old white
dwarfs in the Galactic dark matter halo. Oppenheimer et al. (2001) estimated that about
2% of the dark matter halo mass could be accounted for by old halo white dwarfs. There
has since been some debate as to whether the former result might have been caused by a
dynamically warmer thick disk component (Reid, Sahu & Hawley 2001) and the latter might
have been affected by the misidentification of faint blue extra-galactic sources (Kilic et al.
(2004), also see Reid (2004) for a complete review). The existence of ultra cool white dwarfs
(T 6 4000◦ K) has however been confirmed in recent SLOAN data (Gates et al. 2004).
A search for distant white dwarfs using deeper observations provides another opportunity to
firmly examine the existence of a significant population of old white dwarfs in the Galactic
halo and the dark matter halo. As shown by Kawaler (1996), deep broad band imaging can
probe a significant portion of the Galactic dark matter halo and should identify Galactic
dark halo white dwarfs in the process. Ideally, one would like to identify white dwarf candi-
dates spectroscopically, instead of relying on broadband colors only, and to determine their
Galactic disk or halo memberships using sensitive proper motion measurements.
The Advanced Camera for Survey (ACS) HST observations of the Hubble Ultra Deep Field
(HUDF, Beckwith et al. 2004) and their unprecedented depths offer a new opportunity to de-
tect faint stars, such as white dwarfs, at high galactic latitude. Additionally, the GRAPES
survey (GRism ACS Program for Extragalactic Science. PI: Malhotra, see description in
Pirzkal et al. 2004) provides low resolution spectra of most sources in the HUDF. These
spectra can be used to confidently determine spectral types of stars in the HUDF, down to
i775 = 27.0 mag. More importantly, these spectra allow one to differentiate unresolved, blue,
faint extra-galactic objects (e.g. QSO) from white dwarfs and other stars.
Section 3 outlines how unresolved sources were identified in the HUDF, while the spectro-
scopic identification of unresolved sources is discussed in Section 4. Section 5 describes an
analysis of the cumulative number distribution of the late-type dwarfs identified in the HUDF
and its implication for the Galactic disk structure (Section 5.1). Section 5 also includes a
description of the white dwarf content of the Galactic disk and halo (Section 5.2).
-- 3 --
2. Description of the data
The HUDF was imaged using the ACS and the F435W (B435), F606W (V606), F775W
(i775) and F850LP (z850) bands at different epochs. The observations were carried out by
cycling through each filter while using a small dither pattern to mask out the ACS inter-chip
gap and other cosmetic problems. The HUDF was observed during 4 different visits, rotating
the field of view to a different position angle each time (0◦, 8◦, 85◦, 91◦ with respect to the
first visit, Beckwith et al. 2004). The individual ACS images had an integration time of
1200 seconds, a size of 4096x4096 pixels and a pixel scale of 0.050"/pixel. These images were
combined together by drizzling them onto a finer pixel grid with a pixel scale of 0.030"/pixel,
while geometrically correcting these images. The limiting AB magnitudes (S/N=10) of the
i775 and z850 images are 29.0 and 28.7 mag, respectively. In this paper, images taken from
09/24/2003 to 10/28/2003 and from 12/04/2003 to 01/15/2004 are referred to as Epoch 1
and 2 images, respectively. The elapsed time between the mean observing dates of the two
epochs is 73 days. Images from Epoch 1 and Epoch 2 were taken at position angles that are
nearly 90◦ apart.
While the deep, combined ACS HUDF images described above were used to determine which
objects are unresolved (Section 3), eight special partial image stacks were also assembled to
take advantage of the fact that the data were taken over a period of several months. Four
partial image stacks were created for each HUDF Epoch, using the exact same method used
to create the original combined HUDF images. These partial image stacks can be used to
determine the variability (Cohen et al. 2004) as well as measure the proper and parallax
motions of the objects in the field. The eight partial image stacks have the same physical
properties as the combined ACS HUDF images but with one eighth of the integration time.
Note that the cosmic ray rejection step was done separately while assembling each partial
image stack so that any strongly varying or moving object would not be removed by the
cosmic ray rejection process. The astrometric registration of the combined HUDF images
and the eight partial image stacks is estimated to be better than 0.1 ACS pixel (0.005" or
0.16 HUDF image pixel) (Beckwith et al. 2004).
Spectroscopic observations of the UDF using the ACS slitless grism mode (G800L) were
carried out as part of the GRAPES project. The reader is referred to Pirzkal et al. (2004) for
an in-depth description of these spectra. These spectra are low resolution spectra (R=100),
span a wavelength range of 5500A 6 λ 6 10500A, and can be used to spectroscopically
identify objects in the field.
The HUDF Public Catalog 1.0, containing 10181 objects, was used to identify objects in the
-- 4 --
field and all objects are referred to by their ID in this catalog (UID). The magnitudes of
these objects were re-measured using matched aperture photometry with apertures defined
using Sextractor (Bertin & Arnouts 1996) and the i775 band HUDF image. The photometric
ACS magnitude zero-points (25.65288, 26.49341, 25.64053, 24.84315 in the B435, V606, i775,
z850 bands respectively) were taken from the GOODS survey (Giavalisco et al. 2004) .
3.
Identification of point sources in the HUDF
We identified unresolved objects in the HUDF using a single-index (S) method that is
based on a simple analysis of the curve of growth of the light distribution of an object. The
IRAF task RADPROF was used to produce these light curves of growth. This task first
computes the intensity weighted center of the object (in a 3x3 pixel box), and then performs
aperture photometry using increasing aperture sizes (up to a radius of 10 pixels). These
photometric measurements are then fit (using eight 3 σ rejection cycles) to a cubic spline
of order 5. RADPROF was applied to stamp images of objects which were masked using
Sextractor segmentation maps. These segmentation maps were used to avoid contamination
from neighboring objects. The RADPROF parameters described here were selected because,
following extensive initial testing, they were found to result in curves of growth that are not
too sensitive to the presence of faint nearby objects while consistently remaining sensitive to
the presence of faint, diffuse extended emission around objects. The stellarity S of an object
is defined as
rmax
S 2 =
1
rrmax
(Fobj(r) − Fpsf (r))2
X
r=0
where Fobj(r) and Fpsf (r) denote the curve of growth of light, normalized to 1.0 at a distance
of rmax, of the object and of a point source. A value of 10. was used for rmax.
The index S is then simply the square root of the sum of the squared differences between
the two curves. Unresolved objects have S values that are close to 0.0 while increasingly
resolved objects produce increasingly larger (unbound) values of S. The detailed calibration
of S was done in several stages, each time refining the reference curve of growth that was
used. This is described in some detail below.
First, a few obvious, bright, unresolved and unsaturated sources were pre-selected for further
study. This was done in both the i775 and z850 band images separately. The GRAPES spectra
of these sources were examined and a main-sequence star was selected (UID 911). The curves
of growth of this object were measured separately in the i775 and z850 bands and were used
to compute Si775 and Sz850 values for all 10181 objects in the HUDF. A new set of 25 bright
point sources, well distributed over the whole ACS field, were then carefully selected based on
-- 5 --
their low Si775 values (arbitrarily chosen to be Sz850 6 0.04). The GRAPES spectra of these
25 objects were visually inspected to ensure that these objects were stars. Stamp images of
these objects were individually drizzled (Fruchter & Hook 2002) and carefully re-centered
to produce a stack of images which were then combined together to produce empirical i775
and z850 PSF images. A sigma-based rejection algorithm was used to avoid the effect of
any faint neighbor in the final combined images. Final i775 and z850 band reference curves
of growth were generated using these empirical PSF estimates and all of the Si775 and Sz850
values were re-computed. Figure 1 shows the distributions of the S values in both the i775
and z850 bands, as a function of magnitude. There are clearly two distinct groups of object
in the HUDF images: one group consists of object that are completely or partially resolved,
forming a large distribution around the S value of 0.2 and i775 = 28.0 mag. A second group
consists of unresolved objects with low S values at all magnitudes. Bright stars in Figure 1
are observed to have S values which are consistently lower than 0.05. The exception to this
are bright stars near i775 = 18 mag. which are saturated and have artificially high values of
S (≈ 0.1) . The effect of bright star saturation on the measured values of S were investigated
and values of 0.1 6 S 6 0.15 are expected if the center of a star is saturated or missing in
the HUDF images. Nevertheless, the populations of resolved and unresolved objects appear
to be well separated at brighter magnitudes in Figure 1.
Unresolved sources in the HUDF were selected using the criterion that Si775 6 0.05 mag. Four
bright saturated stars were added to the final object list manually. The final list of unresolved
UDF sources contains 46 objects, spanning the magnitude ranges of 19.0 6 i775 6 29.5 mag
and 18.3 6 z850 6 29.2 mag. Table 1 summarizes some of the properties of these objects.
3.1. Testing the selection process
The ability to distinguish between resolved and unresolved objects using S values was
extensively tested, using a subset of the HUDF object catalog containing the 369 brightest
(z850 6 25.0 mag) objects in the field. Simulated i775 and z850 band images of these objects
-- made 0.0, 0.75, 2.0, 3.0, 4.0, 5.0, 6.0, 7.0, 8.0, and 9.0 magnitudes fainter -- were created
and S values measured using each simulated image, exactly as described in Section 3. The
S value of each of these 369 test objects was thus measured 10 times, each measurement
corresponding to a different apparent magnitude. Figure 2 shows how the S value remains
largely unchanged as an object is made fainter and noisier, all the way down to i775 = 29.0
mag. The vertical pattern seen in this figure is a result of the 10 magnitude steps used in this
simulation. This test demonstrates a very small fraction of misclassification: not a single
extended object was misclassified as a star (e.g. no object with S > 0.05 originally in the
HUDF images was measured to have S 6 0.05 as the object was made fainter), and fewer
-- 6 --
than 5% of stars were misclassified down to i775 = 29.0 mag. While some objects shown in
Table 1 are fainter than this, only objects significantly brighter than this limit (i775 6 27.0
mag) are actually included in the analysis described below and these are expected to be bona
fide unresolved objects.
3.2. Testing the effect of proper motion
Some of the stars in the HUDF objects could have a high tangential-velocity (Vt) and a
correspondingly large proper motion. Such objects would be blurred in the combined HUDF
images and would appear to be resolved. Similarly, nearby objects would suffer from a
significant amount of parallax, again causing these objects to be blurred and mis-identified.
The index S proved relatively insensitive to a small amount of object motion. This is not
unexpected since S was tuned to the detection of faint extended components, The curves of
growth produced by RADPROF were computed by summing, using several sigma rejection
steps, the pixel intensities inside annuli of increasing radii. The effect of a nearby source
is strongly diminished as its separation from the main source increases and the number
of unaffected pixels increases.
In addition, the use of a fifth order spline imposes some
smoothness on the curve of growth which further lowers the effect of a nearby object. Still,
the effect of object motion on measured values of S was carefully investigated by generating
images of the 46 unresolved HUDF objects that were selected as described in Section 3.
These were appropriately shift-and-added, accounting for the orientation and pointing used
for each individual HUDF observation while allowing objects to move as a function of time,
to simulate the effect of proper motion. As larger values of proper motion were introduced,
the S values increased only slightly and S was found to be mostly unaffected by shifts smaller
than 2.5 HUDF pixel (µ = 0.3 ′′/year). 70% of objects were still properly identified as point
sources when shifts 2.5 and 5.0 HUDF pixels were applied (0.3 6 µ 6 0.6).
4.
Identifying stars
4.1. Using Grism spectra to identify stars
The spectral type of 28 unresolved objects brighter than i775 = 27.0 could be determined
robustly using template fitting of their GRAPES spectra (Pirzkal et al. 2004). The template
set included Pickles (1998) main-sequence templates, L and T dwarf spectra (Kirkpatrick
et al. 2000; Burgasser et al. 2003), DA and DB white dwarf spectra (Harris et al. 2003)
obtained from the SDSS archive, and some low temperature (3000◦K, 5000◦K, and 8000◦K)
-- 7 --
model spectra of white dwarfs (Harris et al. 2001). The quality of each spectral fit was
evaluated visually and the results of these fits are given in Table 2. Objects too faint to have
reliable fits are indicated. The GRAPES spectra and the two best-fits listed in Table 2 are
shown in Figures 3, 4, and 5. Unresolved extra-galactic sources are labelled as GAL in Table
2 and are discussed in more details in Section 4.3.
Of the 46 point sources detected in the HUDF, 26 stars, all brighter than i775 = 27.0 mag
were identified. As shown in Table 2, the majority of stars in the HUDF (18 out of 26 )
are found to be M dwarfs. Figures 6 and 7 show these objects in B435 vs. B435-V606 and
i775vs.
i775-V606 plots. Such plots have in the past shown a lack of faint red objects in the
region V606 > 27.0 mag and (B435 − V606) > 1.0 (M´endez et al. 1996). Objects found in this
region of the plot, appear to be either extra-galactic or L-type dwarf spectra (UID 443, 366).
Further NIR observations will be required to verify the nature of these 2 L-dwarf candidates
by confirming that they both have the expected blue NIR colors of this type of object.
The HUDF field is at high galactic latitude (b=223◦, l= -- 54◦) and the effects of extinction
and reddening are small (AV = 0.026, Schlegel et al. 1998). The distances to the HUDF stars,
which are shown in Table 2, were determined using their observed V606 magnitudes, their
estimated stellar types, and the dwarf star main-sequence absolute Johnson V magnitudes
from Allen (2000). The latter were converted to the V606 filter bandpass and AB magnitudes
to be consistent with the data at hand. Distance estimates obtained under the assumption
that these objects are white dwarfs with intrinsic magnitudes of Mmv = 14.0 and 17.0 are
also given.
4.2. Proper motion measurements
The Galactic halo is isotropic and has no or little overall rotation. Halo objects are
therefore expected to lag behind the rapidly rotating Galactic disk component (and the local
standard of rest) with an average velocity of ≈ 200 ± 100km/s . As done in Oppenheimer
et al. (2001), it is assumed here that Galactic halo objects have a velocity that is more than
2σ above the velocity expected for an object that is in the Galactic disk (i.e. > 100 km/s,
Chiba et al. 2000).
The HUDF image stacks (Section 2) were subtracted from one another and the results were
examined to see if any object, whether classified to be extended or not based on its Sz850
value, had moved significantly (e.g. more than 2.5 UDF pixel) from one image to the next.
No such object was found.
The motion of objects in the HUDF was further investigated by carefully measuring the
position of objects separately in each of the eight available i775 band HUDF partial image
stacks. These positions were measured using the IDL task MPFIT2DPEAK 7, using a fixed
-- 8 --
width (FWHM=2.5 pixel) Gaussian fit.
Initially, the positions of the 360 most compact
sources (S 6 0.1 and i775 6 28.5 mag) were examined. An average position and a standard
deviation of the mean were computed for each object in the field using Epoch 1 and Epoch
2 HUDF image stacks separately. Average Epoch 1 and Epoch 2 positions were subtracted
from one another to compute the observed shift of each object. These are shown (scaled up
by a factor of 500) in the left panel of Figure 8. Repeated measurements of the position
of sources in the field using HUDF image stacks taken within the same Epoch were always
within 0.1 HUDF pixel, or 0.003". Still, and as shown in Figure 8, there is a systematic,
field dependent disagreement between the Epoch 1 and Epoch 2 measurements. This is likely
caused by Epoch 1 and Epoch 2 images having been taken at position angles that are 90◦
apart, and/or the effect of telescope breathing, and/or very small amount of residual image
distortion that is not taken into account in the current model of the ACS distortion maps.
To produce a 3σ detection with these data, an object must have a measured proper motion
larger than 0.54 HUDF pixel or µ > 0.081 ′′/year. No object in Table 1 was observed to
have such a large proper motion.
The situation was however improved using a simple third order two dimensional polynomial
fit to the observed distribution shown in the left panel of Figure 8. This fit, when applied as
a correction, results in significantly lower systematics across the field, as shown in the right
panel of Figure 8. The accuracy of the corrected shifts (of compact objects) between Epoch
1 and Epoch 2 is improved by a factor of 3, with a 3σ detection now corresponding to a shift
of 0.18 HUDF pixel or µ > 0.027 ′′/year.
The search for high proper motion objects in the HUDF was extended to all 10181 objects in
the field by measuring and correcting (using the correction determined above) the observed
shifts of these objects. Only four sources were found to have moved significantly (but with
a significance slightly lower than 3σ) from Epoch 1 to Epoch 2. These objects are UID 443,
9020, 911, and 7525 and they are all unresolved objects that are listed in Table 1. Their
measured shifts are 0.17, 0.17, 0.14, and 0.15 HUDF pixels respectively, or approximately
0.02"/year, which is smaller than our imposed 3σ limit.
4.3. Extragalactic Unresolved Sources
Two unresolved, extra-galactic sources with i775 6 27.0 were identified (UID 6732, 9397).
The GRAPES spectrum of these objects show prominent emission lines and the redshift of
these objects is estimated to be z=3.2 and z=3.0, respectively (Xu 2004). Identification of
such objects using GRAPES spectra could easily be carried out down to magnitudes fainter
7http://astrog.physics.wisc.edu/ craigm/idl/
-- 9 --
than i775=27.0 mag, as demonstrated by the fact that several fainter extragalactic objects
with 27.0 6 i775 6 29.5 were easily identified (e.g. UID 4120, 8157). Without spectral
information, these objects would have been misidentified as stars (e.g. UID 9397 would be a
good candidate for a cool white dwarf at a distance of ≈ 1 Kpc). The issue of contamination
by faint, blue extra-galactic objects in previous studies (Ibata et al. 1999; M´endez & Minniti
2000) has been discussed in the past (Kilic et al. 2004) and Figures 6 and 7 illustrate how
these objects could be mistaken for a variety of stellar objects ranging from M and L dwarfs
to white dwarfs. Spectroscopic confirmation is therefore crucial to accurately determine
stellar counts in deep fields such as the HUDF. There are 2 extra-galactic objects amongst
the 28 point sources with i775 6 27.0 mag, which corresponds to a 7% contamination level
at these Galactic coordinates.
The 2 extra-galactic objects (UID 6732, 9397) were used to check the accuracy of the proper
motion measurements described in Section 3.2. The measured shifts of these objects between
Epoch 1 and Epoch 2 are smaller than 0.04 HUDF pixel (0.0012"), which is what one would
expect based on the proper motion error estimate quoted in Section 3.2).
There are a total of 14 faint (27.0 6 i775 6 29.5), blue ((B435 −V606) 6 1.8, (V606 −i775) 6 0.8)
objects listed in Table 1 that were not included in the analysis above because they are either
too faint to have useable GRAPES spectra or robust S based classifications. Some of these
sources might well be faint white dwarfs in either the Galactic disk or halo but are hard
to distinguish from extra-galactic sources without higher signal-to-noise spectra or more
accurate proper motion measurements.
5. Discussion
5.1. M and L Dwarfs and the Galactic disk scale height
18 objects were identified to be M (or early L) dwarf main-sequence stars. Two objects
were found to be best fit by early L dwarf templates (Kirkpatrick et al. 2000). Both of these
objects are red enough to be outside of the color range shown in Figures 9 and 10 with
(V606 − i775) = 2.21 and (V606 − i775) = 3.47. Deep infrared observations of these two objects
will be required to confirm that they are L dwarfs. NICMOS observations of the HUDF did
not detect these objects in either the J or H bands. All other very red objects are best fit
using M dwarf templates. The broad spectral features of M dwarfs listed in Table 2 were
well fitted using M dwarf Pickles (1998) templates, and for objects brighter than i775 = 27.0
mag we believe the identification to be secure.
Because these late-type sources can be seen to great distances, they can be used to probe
the structure of the Galactic disk and halo as traced by the lowest-mass stellar components
-- 10 --
of these populations. This was explored using Monte Carlo mass function (MF) simulations
based on those developed by Burgasser et al. (2004). Assuming power-law representations
of the MF, dN/dM ∝ M−α, for masses 0.005 to 0.2 M⊙, luminosity (LF) and effective
temperature (Tef f ) distributions were created using evolutionary models from Burrows et
al. (1997) over the range 100 . Tef f . 4000 K. For all simulations, α was fixed at 1.13 for
0.1 . M . 0.2 M⊙ (Reid et al. 1999) but allowed to vary between 0.5 -- 1.5 for M . 0.1 M⊙;
number densities are normalized to the empirical value at 0.1 M⊙ from the 8 pc sample (Reid
et al. 1999).
Two separate populations were simulated: a disk population with a flat age distribution
spanning 10 Myr to 10 Gyr (the majority of sources have age greater than 1 Gyr), and
a halo population with uniformly sampled ages between 9-10 Gyr and a relative number
density of 0.25% (Digby et al. 2003). The resulting bolometric LFs were converted to I
(Cousins) LF's using a polynomial fit to empirical data (spanning 12.4 . MI . 22.2; i.e., M4
to T8) from Dahn et al. (2002) and Golimowski et al. (2004). Note that this fit is based on
measurements for Solar metallicity field dwarfs, and may not be appropriate for a subsolar
metallicity halo subdwarf population. Apparent I-band distributions were determined by
first assuming a constant density population out to a limiting magnitude mI = 27.5 (taking
into account the difference between HUDF AB-magnitudes and Vega magnitudes from the
empirical data) and then applying a correction for the vertical distribution of sources. For
the disk population, we assumed a density distribution n ∝ sech(z/h0)2 (Reid & Hawley
2000), where z = d sin β, d is the source distance, β = 54.5◦ is the Galactic latitude of the
HUDF, and h0 is the 1/e disk scale height, assumed to range between 200 and 500 pc. For
the halo population, a Galactocentric oblate spheroid distribution as given in Digby et al.
(2003) was assumed, using the values derived there and an axial ratio q = c/a = 0.7.
Figure 11 presents the results of these simulations in the form of cumulative number
distributions as a function of i775 magnitude down to the limiting magnitude of our sample.
These distributions show that variations in the disk scale height are far more pronounced
than those from the different MF's assumed, and this analysis is limited to the former
parameter. The cumulative distribution of the six dwarfs in Table 2 with spectral types
M4 and later (consistent with the mass constraints of the simulations) is also shown. The
observed distribution matches that of the h0 = 400 pc disk simulation very well, particularly
out to i775 . 23.5 mag where the halo population makes negligible contribution. Because the
simulated distributions are rather sensitive to the disk scale height, particularly at fainter
apparent magnitudes, one can conservatively constrain h0 to ±100 pc assuming Poisson
uncertainties. This is in good agreement with the estimate of 340 ± 84pc from Ryan et al.
(2004) which was computed using ACS HUDF parallel fields i-drops. For i775 > 25 mag, the
observed distribution is slightly greater than the disk population alone but does not increase
-- 11 --
as sharply as the combined disk+halo distribution. Indeed, the shape of the observed number
distribution implies few if any halo stars in the HUDF down to i775 = 27.0 mag, consistent
with the lack of significant proper motion sources in this sample. This suggests that either
the number density ratio of halo to disk stars, or the adopted axial ratio for the halo density
distribution, or both, may be smaller than assumed here. With no late-type halo subdwarf
detections in this sample, we cannot usefully constrain these possibilities.
5.2. White Dwarfs
5.2.1. White Dwarf candidates
Objects brighter than i775 = 27.0 were individually examined and distances were esti-
mated under various assumptions of what the exact nature of each object might be: (1) a
main sequence star; (2) a young white dwarf; (3) an old white dwarf. The reddest objects,
which are all well-fitted by M-and-later type templates, as well as the few extra-galactic
objects identified in Section 4.3 were excluded from this analysis. The remaining 8 objects
were all initially considered to potentially be white dwarfs.
Assuming that an object is a white dwarf, one can use the measured V606 − i775 color
of that object (Table 1) and the cooling curves of Richer et al. (2000) (after accounting for
passbands and zeropoints differences), to derive an absolute V606 band magnitude, an age,
and a distance to that object. As shown in figure 2 of Richer et al. (2000), blue white dwarfs
can either be young, hot objects (Age 6 5 Gyr, T ≈ 104K, MV606 6 16.0 mag) or older,
cooler, intrinsically dimmer ones (Age > 10 Gyr, T ≈ 3 × 103 K, MV606 > 16.0 mag). This
results in several white dwarf distance estimates for each object.
Proceeding via a process of elimination, the distinction between main sequence stars, young
white dwarfs, old white dwarfs, and disk or halo white dwarfs is possible. Assuming that a
star is a main sequence star, which is intrinsically much more luminous than a white dwarf,
can cause the distance estimate for that object to be unreasonably large. Assuming that a
star is a young white dwarf (Age 6 5.0 Gyr) implies that this object is less likely to be part
of the Galactic halo since the latter is composed of much older objects. Also, and based on
the lack of proper motion detection discussed in Section 4.2, one can define an upper limit to
the tangential velocity (VT max) of each object once its distance is determined. The motion
of an object moving by more than 0.18 HUDF pixel (0.027 "/ yr) should have been detected
at the 3σ significance level as discussed in Section 3.2. As discussed above, objects in the
Galactic halo are expected to have velocities around 100-200 km/s. Even if projection effects
should result in lower values of VT , one would expect at least some of them to have VT values
-- 12 --
larger than 30-60 km/s. A final clue to help narrow down the nature of a particular object
is provided by the fact that an object that is nearby would suffer from a significant, easily
measurable, parallax during the 73 days interval between Epoch 1 and Epoch 2 observations.
Calculating the parallax vector for the HUDF exposures shows that an object 30 pc away
would produce a parallax of about 0.060" or 2 HUDF pixel between Epoch 1 and Epoch
2 observations. Objects that are closer than 200 pc are therefore expected to produce a
parallax that would be detected at more than 3σ level.
Examining Table 3, objects 4322, 4839, 7768, 9020 are unlikely to be main-sequence
stars as this would make them very distant objects. The Magellanic Stream (Mathewson
& Ford 1984) is too far away from the HUDF to possibly explain the existence of main-
sequence stars at such large distances. Objects 1147, 3166, and 5921 which do not (this
is true of all objects listed in Table 3) have measurable parallax or proper motion, are not
likely to be white dwarfs. Being either young or old white dwarfs would place these two
objects less than 200 pc away and would have resulted in measurable parallax in the HUDF
image stacks. Another object unlikely to be a white dwarf is object 9230 which was ob-
served to have a fainter, unresolved companion 0.5" away. This companion, for which we
have no GRAPES spectrum, has a V606 band magnitude of V606 = 23.7 mag and the colors
B435 − V606 = 1.6, V606 − i775 = 0.9, i775 − z850 = 0.4 mag. Based on these colors, this object
should be an early M dwarf that is (4.9 ± 1.2) × 103 pc away. This places both object 9230
and its companion at the same distance (within error bars). The likelihood of any two stars
being within 0.5" in the HUDF, and at nearly the same distance, is very small (6 5 × 10−3)
and these two objects are likely to be part of a binary system, with object 9230 being a K
type main-sequence star. Independently of this, object 9230 cannot be an old white dwarf
since this would place this object at distance of 34 pc where its parallax motion would have
been very easily detected.
5.2.2. Disk or halo white dwarfs?
Four objects (4322, 4839, 7768, 9020) remain as white dwarf candidates. Using a simple
1/Vmax (Schmidt 1968; Tinney et al. 1993; M´endez 2000) analysis, the white dwarf number
density in the direction of the HUDF can be computed. The detection of 4 white dwarfs in
the HUDF implies a local density of (3.5 ± 1.5) × 10−5 6 stars/pc3 6 (1.1 ± 0.3) × 10−2.
The upper and lower limits on the density are computed using old or young white dwarfs
respectively, while the errors reflect the large uncertainties in the intrinsic luminosities of
-- 13 --
young and old white dwarfs (∼ 1.0 mag). Note that this estimate is actually not affected
by whether or not object 9230 was included as a HUDF white dwarf. The contribution
of this object to the 1/Vmax is negligible. For comparison, Liebert, Dahn & Monet (1988)
previously determined the local white dwarf density to be 3.0 × 10−3 stars/pc3, while Reid,
Sahu & Hawley (2001) found a value of (3.26 ± 1.23) × 10−3 stars/pc3, and M´endez (2000)
derived a density of white dwarfs in the thick disk of (2.610 ± 0.59) × 10−2 stars/pc3. The
white dwarf density derived here is consistent with most of the previous density estimates.
Properly determining the disk and halo membership of the 4 white dwarfs identified in the
HUDF will be needed before a stronger conclusion can be made. Re-observing the HUDF
field in the i775 band in 1.5 years would allow (assuming that a 0.18 HUDF pixel shift corre-
sponds to a 3σ level detection) one to unambiguously detect the proper motion of halo white
dwarfs with VT > 100 km/s at distances up to 1000 pc.
The above density estimate assumed that all four HUDF white dwarfs were Galactic disk
objects. Could some of these objects be in the Galactic halo? At least two of these objects
(4839, 7768) have VT max values that are lower than 60 km/s. The probability that this is
caused by a projection effect is under 10% (assuming a relatively low velocity of ≈ 100 km/s
for halo objects). The last two white dwarfs listed in Table 3 (4322, 9020) are at slightly
larger distances which average to 583 pc, and more importantly have larger values of VT max.
The average distance of these objects still seems a bit low for objects which would be part
of a Galactic halo whose density increases all the way up to 7 kpc (Reid 1993), while white
dwarfs with absolute magnitudes of MV606 = 17 mag (Harris et al. 2003) should be detected
all the way out to 1000 pc. Unless these objects are intrinsically much dimmer, one would
expect the average distance to these two objects to be larger than 583 pc and much closer to
our limit of 1000pc. The VT max values inferred for objects 4322 and 9020, while higher than
those of objects 4839 and 7768, fall substantially short of expected typical halo velocities.
The probability of observing each of the objects with the VT max values listed in Table 3 is
between 14 -- 20%. As a group, the probability that objects 4322 and 9020 are halo objects is
under 3%. Similarly, the hypothesis that all 4 detected white dwarfs in the HUDF are halo
rather than disk objects is excluded at the 99.9986 % level. Overall, the distances to the four
white dwarf candidates identified here is more consistent with them being part of a Galactic
thick disk. Majewski & Seigel (2002) derived a disk scale height of 400 -- 600 pc using a study
of low velocity white dwarfs. This value was somewhat larger than previous estimates of
250 -- 350 pc but it is interesting to note that the M-dwarf disk scale height derived in Section
5.1, as well as the distance estimates of the white dwarfs listed in Table 3 all appear to agree
with this value.
It would be interesting to attempt to set an upper limit on the halo white dwarf density based
on the finding of up to two white dwarfs in the HUDF. Under this assumption, the upper
limit to the Galactic halo white dwarf density is computed to be (6.53 ± 3) × 10−3 stars/pc3.
-- 14 --
This result is consistent with the previous work of M´endez & Minniti (2000) using the com-
bined observations of the Hubble Deep Fields North and South (7.73 × 10−3stars/pc3).
A population II halo white dwarf population that is older than 12-13 Gyr would have re-
mained undetected here (An intrinsically fainter than MV606 = 17.0 mag white dwarf pop-
ulation would also help explaining the low average distances of the white dwarfs listed in
Table 3). As discussed above, the white dwarfs identified in this work are more likely to be
within the Galactic thick disk. It is to be noted that Kilic et al. (2004) recently observed
a disproportionately large fraction of disk to halo white dwarfs in the Hubble Deep Field
North (HDF-N). Kilic et al. (2004) found no evidence of any white dwarfs with tangential
velocities larger than 30 km/s down to i775 = 27.5 mag and the authors concluded that the
blue objects they saw were all part of the Galactic disk. Similarly, the HUDF observations
presented here successfully identified 4 white dwarf candidates with i775 ≤ 27.0 mag in the
HUDF, none very likely to be in the Galactic halo while 2 -- 3 detection would have been
expected to be consistent with previous studies.
5.2.3. Dark halo white dwarfs?
While dynamical studies of the Galaxy predict the dark matter white dwarf density as
high as 1.26 × 10−2 stars/pc3 (assuming that 100% of the dark matter halo mass is in the
form of 0.6M ⊙ white dwarfs), the failure to detect a significant population of high velocity
white dwarfs in the HUDF points to a dark matter halo devoid of a significant white dwarf
population. Following the methodology of Kawaler
(1996), one can estimate the amount
of dark matter halo mass that is probed by the HUDF images, and directly compute the
fraction of the dark matter halo mass that could be contained in a white dwarf population.
Assuming a white dwarf absolute Mi775 magnitude of 17.0 mag(Harris et al. 2003) and a
limiting magnitude of i775 = 27.0 mag, the maximum probed distance is 1000 pc. In the
corresponding volume, and in the direction of the HUDF, the included dark matter halo
mass is 3 M⊙. A population of dark matter halo white dwarfs could possibly have remained
undetected in this study for several reasons: (1) The method used in this paper may be
inefficient at properly identifying halo white dwarfs in Table 3. In this case, if one assumes
that the HUDF contains 4 halo white dwarfs (i775 6 27.0), these would account for 2M⊙,
a significant fraction of the expected dark matter halo mass. This would however increase
the ratio of halo to disk white dwarfs to unrealistically high levels (i.e. ≫ 2%), and would
assume that all the white dwarfs seen in the HUDF are from the dark matter halo and
not from the pop II halo, which is highly unrealistic. The lack of proper motion detections
also makes this scenario unlikely; (2) White dwarfs do not contribute significantly to the
-- 15 --
Galactic dark matter halo mass. From microlensing experiements towards the Magellanic
clouds, the contribution of MACHOs to dark matter has been estimated to be anywhere
between 20% to less than 2% (Alcock et al. 2000; Afonso et al. 2003; Sahu 2003; Evans
& Belokurov 2004, and references therein). Furthermore, Majewski & Seigel (2002) showed
how an increased thick disk scale height of 400-600 pc makes it unlikely that white dwarfs
could be the MACHO objects; (3) White dwarfs could contribute significantly to the mass
of the Galactic dark matter halo but are too faint to be detected in this study. Richer et al.
(2000) showed that an increase in the age of the white dwarf halo population from 10 Gyr
to 16 Gyr is expected to reduce the number of white dwarf detection (down to V606 = 28.0)
by a factor of 7. The lack of high velocity white dwarf detections in the HUDF puts upper
limits on the contribution of an hypothetical white dwarf population to the Galactic dark
matter halo. Based on the brightest objects in the HUDF for which we have spectra (i.e.
i775 6 27.0), if the dark matter halo is as old as about 12 Gyr (Hansen 2002; Charboyer et
al. 1998) then dark matter halo white dwarfs with MV = 17.5 (Mi775 = 17.0) contribute less
than 20% to the dark matter. If the age of the dark matter halo is 10 Gyr, white dwarfs
(Mi775 = 15.9) contribute less than 2% to the dark matter.
The possibility that there is a large population of faint white dwarfs that has remained
unidentified in this study and which would account for a significant fraction of the dark
matter halo can be investigated a little further. Recall that in Section 4.3, 14 sources with
(27.0 6 i775 6 29.5) and blue ((B435 − V606) 6 1.8, (V606 − i775) 6 0.8) colors were identified.
Not all of these sources were expected to be bona-fide stars and some are likely to be extra-
galactic sources, but as discussed in Section 3.1, it is estimated that 95% of the stars in
the HUDF were properly identified, so that this number should be considered to be an
upper limit on the number of faint starts in the HUDF. Following the same methodology
used above, but now using a limiting magnitude of i775 6 29.5, the HUDF images probe
through 79 M⊙ of the dark matter halo. Even under the extreme assumption that all 14
unresolved candidates in the UDF are high velocities white dwarfs, this implies that, when
reaching down to i775=29.5, faint white dwarfs in the Galactic dark matter halo contributing
less than 10% to the total dark matter halo mass for a dark matter halo age of 12 Gyr.
Future observations of the HUDF would allow more sensitive measurements of the proper
motion of these faint objects (Section 5.2.2). Setting tighter constraints on the maximum
tangential velocities of these faint HUDF unresolved objects would allow to exclude some of
these objects from high velocity white dwarfs and would allow to further lower the maximum
fraction of the dark matter halo mass that could be accounted for by white dwarfs.
-- 16 --
6. Conclusions
A systematic search for unresolved objects in the HUDF identified 46 objects with
i775 6 29.5 mag. Using the GRAPES spectra of these objects, 28 objects with i775 6 27.0 mag
were spectroscopically identified, including 18 M and later dwarfs (including 2 unconfirmed L
dwarf candidates), and 2 QSOs. The M dwarf luminosity function of M4 and later-type stars
was computed and compared to predictions based on Monte-Carlo simulations. Assuming a
simple analytical model of the Galactic disk, the number of detected M dwarfs was shown
to be large enough to constraint the scale height of the M and L dwarf Galactic disk to be
h0 = 400 ± 100pc. Out of 8 stars that were found to not be M or later-type stars, four
stars were determined to be old white dwarf candidates. Not a single object was found
to have a proper motion that is larger than 0.027"/year making them likely Galactic thick
disk objects. Further imaging of the HUDF, using a time span larger than 1.5 year would
provide more sensitive proper motion measurements. More sensitive proper measurements
are required to positively place unresolved objects in the Galactic disk or halo. It would also
allow to search for high tangential velocity objects amongst the 14 fainter (27.0 6 i775 6 29.5)
unresolved sources identified in the HUDF. Excluding halo membership of objects down to
i775 = 29.5 based on proper motion measurements would further constraint the maximum
white dwarf contribution to the dark matter halo mass. The currently available observations
of the HUDF, spanning 73 days, show the absence of a significant population of high velocity
white dwarfs down to i775 = 27.0, and a relatively small number of unresolved, faint blue
sources in the field down to i775 = 29.5. This is interpreted as a consequence of white dwarfs
accounting for less than 10% of the dark matter halo, assuming that the dark matter halo
is 12 Gyr old.
We would like to thank I. N. Reid for helpful discussions. This work was supported
by grant GO -09793.01-A from the Space Telescope Science Institute, which is operated by
AURA under NASA contract NAS5-26555.
REFERENCES
Afonso C. et al. 2003, A&A, 404, 145
Alcock, C., et al. 2000, ApJ, 542, 281
Allen, C. W. 2000, Astrophysl Quantities 4th ed. (Athlone Press, London), p388
Beckwith S., et al. 2004, in preparation
-- 17 --
Bertin, E. Arnouts, S. 1996, A&AS, 117, 393
Burgasser, A. J., Kirkpatrick, J. D., Liebert, J. & Burrows, A. 2003, ApJ, 594, 510
Burgasser, A. J. 2004, ApJS, in press (astro-ph/0407624)
Burrows, A. et al. 1997, ApJ, 491, 856
Chaboyer, B., Demarque, P., Kernan, P.J., Krauss, L.M. 199. Astrophys. J. 494:96-110
Chiba, M. & Beers, T. C., AJ, 199, 2843
Cohen, S. H., Ryan, R. E., Hathi, N., Windhorst, R. A., Koekemoer, A., Xu, C., Pirzkal,
N., Malhotra, S. & Strolger, L. 2004, ApJ in preparation
Dahn, C. C. et al. 2002, AJ, 124, 1170
Digby, A. P., Hambly, N. C., Cooke, J. A., Reid, I. N., & Cannon, R. D. 2003, MNRAS, 344,
583
Evans, N. W., Belokurov, V. 2004, in press (astroph/0411222)
Fruchter, A. S. Hook, R. N. 2002, PASP, 114, 144F
Gates, I. E. & Gyuk, G. 2001, ApJ, 547, 786
Gates, I. E. et al. 2004, in press (astroph/0405566)
Giavalisco, H. C. et al. 2004, ApJ, L93
Golimowski, D. A. et al. 2004, AJ, 127, 3516
Harris, H. C. et al. 2001, ApJ, 549, L109
Harris, H. C. et al. 2003, AJ, 126, 1023
Hansen, B. M. S., et al. 2002, AJ, 574, L155
Ibata, R. A., Richer, H. B., Gilliland, R. L. & Scott, D. 1999, ApJ, 524, L95
Ibata, R. A., Irwin, M., Bienaym´e, O., Scholz, R. & Guibert, J. 2000, ApJ, 532, L41
Kawaler, S., D. 1996, AJ, L61
Kilic, M., von Hippel, T., M`endez, R. A. & Winget, D. E. 2004, ApJ, 609, 766
Kirkpatrick et al. 2000, AJ, 120, 447
-- 18 --
Liebert, J., Dahn, C., C. & Monet, D. G. 1988, ApJ, 332, 891
Majewski, S. R., Siegel, M. H., 2002, ApJ, 569, 432
Mathewson, D., S. & Ford, V., L. 1984, IAUS, 108, 125
M´endez, R. A., Minniti, D., De Marchi, G., Baker, A. & Couch, W. J. 1996, MNRAS, 283,
666
M´endez, R. A. & Minniti, D. 2000, ApJ, 529, 911
M´endez, R. A. 2002, A&A, 395, 779
Oppenheimer, B. R., Hambly, N. C., Digby, A. P., Hodgkin, S. T. & and Saumon, D. 2001,
Science, 292, 698
Pirzkal N., et al. 2004, ApJS, 154
Pickles, A. J. 1998, PASP, 110, 863
Reid, I. N. & Majewski, S. R. 1993, ApJ, 409, 635
Reid, I. N. et al. 1999, ApJ, 521, 613
Reid, N. & Hawley, S. L. 2000, New Light on Dark Stars (Chichester: Praxis)
Reid, I. N., Sahu, K., C. & Hawley, S. L. 2001, ApJ, 559,942
Reid, I. N., ARA&A, 43, in press
Richer, H. B., Hansen, B., Limongi, M., Chieffi, A., Straniero, O. & Fahlman, G. C. 2000,
ApJ, 529, 318
Robin, A. C., Reyl´e, C., Derri`ere, S. & Picaud, S. 2003, A&A, 409, 523
Ryan, R. E., Hathi, N. P., Cohen, S. H., & Windhorst, R. A., ApJ in preparation
Sahu, K. C. 2003, Space Telescope Science Institute symposium series, Vol. 15, Cambridge
University Press, p. 14 - 23 (astro-ph/0302325)
Sahu, K. C. 1994, Nature, 370, 275
Schlegel, D. J., Finkbeiner, D. P. & Davis, M. 1998, ApJ, 500, 525
Schmidt, M. 1968, ApJ, 151, 393
-- 19 --
Tinney, C. G., Reid, I. N. & Mould, J. R. 1993, ApJ, 414, 254
Xu, C. et al. 2004, ApJS, in preparation
This preprint was prepared with the AAS LATEX macros v5.0.
-- 20 --
Table 1. Unresolved objects in the HUDF, as selected using the criterion Sz850 60.05
(F850LP ACS band).
UID
RA (J2000) Dec (J2000)
B435
B435-V606
V606-i775
i775-z850
Si775
Sz850
19
366
443
834
911
1147
1343
2150
2368
2457
2977
3166
3561
3794
3940
4120
4322
4643
4839
4945
5317
5441
5921
5992
6334
6442
6461
6620
6732
7113
7194
7357
7525
7768
7894
8081
8157
8186
53.1623248
53.1753062
53.1583988
53.1648156
53.1460739
53.1834639
53.1398080
53.1767483
53.1676057
53.1616105
53.1726388
53.1583425
53.1483050
53.1474097
53.1498728
53.1839107
53.1349058
53.1848922
53.1883621
53.2003709
53.1257144
53.1625253
53.1322881
53.1498432
53.1782000
53.1416300
53.1984259
53.1428439
53.1784892
53.1707114
53.1860335
53.1692789
53.1318579
53.1472186
53.1831085
53.1580838
53.1649526
53.1867094
-27.8269460
-27.8198904
-27.8189953
-27.8143652
-27.8123237
-27.8067589
-27.8100295
-27.7996684
-27.8037923
-27.8027738
-27.8006672
-27.7949215
-27.7977856
-27.7965561
-27.7961691
-27.7954147
-27.7944480
-27.7933971
-27.7923200
-27.7898451
-27.7904647
-27.7896458
-27.7828516
-27.7874448
-27.7861688
-27.7856726
-27.7848674
-27.7849376
-27.7840395
-27.7826303
-27.7822207
-27.7813943
-27.7820076
-27.7714786
-27.7802451
-27.7701503
-27.7736713
-27.7735490
26.58
29.99
31.25
27.67
25.21
19.58
27.36
22.60
29.07
26.98
27.94
20.80
29.24
28.89
27.06
28.04
26.83
31.73
27.22
25.06
29.06
27.81
20.57
27.52
27.71
28.16
26.80
31.27
25.73
30.83
26.91
27.93
28.70
27.12
27.46
27.92
30.85
27.72
1.91
2.20
0.86
2.35
2.09
-0.16
-0.23
1.99
1.18
1.91
-0.18
0.96
1.98
2.14
-0.48
0.03
0.06
1.52
1.20
2.06
0.49
1.75
0.77
1.63
-0.27
-0.19
1.77
3.19
1.13
2.49
-0.12
0.10
2.22
1.66
0.03
-0.32
2.16
-0.13
1.01
3.05
3.47
2.00
1.80
0.52
-0.14
1.43
0.08
0.98
-0.21
0.80
0.91
0.98
-0.21
0.12
-0.06
0.74
0.40
1.74
-0.00
1.27
1.19
0.69
-0.16
0.04
1.00
0.25
-0.04
0.65
-0.13
-0.21
1.08
0.68
-0.15
-0.40
0.40
-0.14
0.30
1.27
1.51
0.76
0.65
0.13
-0.04
0.80
-0.20
0.30
0.18
0.37
0.24
0.29
-0.18
1.31
-0.07
0.25
0.07
0.67
-0.30
0.39
0.51
0.16
-0.18
-0.07
0.30
-0.14
0.01
0.20
-0.06
-0.16
0.30
0.20
-0.12
-0.30
0.07
-0.34
0.015
0.009
0.016
0.013
0.007
0.091
0.045
0.147
0.046
0.004
0.042
0.086
0.018
0.010
0.033
0.023
0.067
0.019
0.018
0.020
0.045
0.016
0.094
0.014
0.049
0.044
0.010
0.041
0.010
0.026
0.048
0.024
0.011
0.014
0.047
0.038
0.042
0.044
0.004
0.030
0.028
0.011
0.002
0.002
0.046
0.003
0.050
0.008
0.049
0.007
0.012
0.007
0.038
0.012
0.043
0.050
0.015
0.013
0.039
0.019
0.070
0.005
0.049
0.030
0.010
0.049
0.022
0.024
0.045
0.031
0.007
0.005
0.039
0.020
0.039
0.029
-- 21 --
Table 1 -- Continued
UID
RA (J2000) Dec (J2000)
B435
B435-V606
V606-i775
i775-z850
Si775
Sz850
9006
9020
9212
9230
9331
9351
9397
9959
53.1853179
53.1685421
53.1485421
53.1580164
53.1638484
53.1781506
53.162852
53.1611698
-27.7799976
-27.7805214
-27.7701387
-27.7691869
-27.767128
-27.7691241
-27.7671662
-27.7555109
27.46
27.00
25.50
21.18
27.95
26.21
21.44
29.12
-0.05
0.28
1.82
0.76
2.20
2.03
0.26
2.42
-0.07
-0.01
0.96
0.38
1.84
1.32
0.11
0.91
0.24
-0.16
0.29
0.10
0.67
0.42
0.02
0.29
0.039
0.025
0.011
0.037
0.025
0.015
0.043
0.022
0.024
0.034
0.003
0.007
0.005
0.004
0.033
0.021
-- 22 --
Table 2. Spectral types of the 46 unresolved objects in the HUDF (28 brighter than
i775 = 27.0 mag) and their distance estimates assuming that they are main-sequence objects
or white dwarfs with MV = 14 or MV = 17. Two additional, fainter extra-galactic sources
are included (UID 4120, 8157) The spectral types were determined by fitting the GRAPES
(Pirzkal et al. 2004) spectra of these sources to a set of templates (See Section 4.1).
UID
Stype
i775 > 27.0
MV606
Log(D)
MS
WD MV = 14 WD MV = 17
19 M1-M2
366
L0-L1
443
L0-M6
834 M4-M5
911 M4-M5
1147
F6-F8
1343 A0-K7
2150 M3-M4
2368 O9-M6
2457 M0-M1
2977 A0-M4
3166 K4-K5
3561 M0-M1
3794 M0-M1
3940 O5-M3
4120
GAL
4322 A7-K5
4643 A0-M6
4839 G5-K2
4945 M4-M5
5317 A0-M6
5441 M2-M3
5921 K4-K5
5992 K7-M0
6334
F8-L0
6442 A0-G5
6461 M0-M1
6620 B8-M4
6732
GAL
7113 G8-K0
7194 O5-K5
7357 O9-M5
7525 M1-M2
7768 K5-K7
7894
B8-K7
8081 O5-A3
8157
GAL
no
no
no
no
no
no
yes
no
yes
no
yes
no
no
no
yes
yes
no
yes
no
no
yes
no
no
no
yes
yes
no
yes
no
yes
yes
yes
no
no
yes
yes
yes
9.30 ± 0.25
18.9 ± 0.50
15.6 ± 2.9
11.4 ± 0.39
11.4 ± 0.39
3.82 ± 0.17
4.32 ± 3.58
11.3 ± 0.23
4.19 ± 8.55
8.77 ± 0.28
11.28 ± 0.23
6.96 ± 0.14
8.77 ± 0.28
8.77 ± 0.28
2.37 ± 7.94
4.70 ± 2.4
6.75 ± 5.96
5.65 ± 0.61
11.4 ± 0.38
6.71 ± 6.0
9.93 ± 0.38
6.96 ± 0.14
8.07 ± 0.43
8.34 ± 4.4
2.89 ± 2.2
8.77 ± 0.28
5.45 ± 5.6
5.60 ± 0.19
0.76 ± 6.3
3.73 ± 8.1
9.33 ± 0.23
7.37 ± 0.27
3.73 ± 3.9
−1.98 ± 3.6
4.072 ± 0.050
2.780 ± 0.100
4.261 ± 0.570
3.777 ± 0.077
3.338 ± 0.077
4.180 ± 0.033
6.080 ± 0.716
2.862 ± 0.045
7.145 ± 1.710
4.258 ± 0.056
4.365 ± 0.045
3.573 ± 0.027
4.697 ± 0.056
4.594 ± 0.056
7.315 ± 1.588
5.632 ± 0.479
6.578 ± 1.191
5.087 ± 0.122
3.315 ± 0.077
6.267 ± 1.199
4.228 ± 0.076
3.565 ± 0.027
4.568 ± 0.085
5.505 ± 0.875
6.271 ± 0.429
4.251 ± 0.056
6.345 ± 1.121
5.543 ± 0.038
7.215 ± 1.266
7.131 ± 1.618
4.429 ± 0.045
4.616 ± 0.055
6.223 ± 0.780
7.471 ± 0.716
3.129
3.750
4.274
3.258
2.818
2.142
3.713
2.315
3.774
3.209
3.818
2.164
3.647
3.544
3.702
3.548
4.237
3.400
2.796
3.910
3.407
2.155
3.373
3.791
3.865
3.202
3.812
3.861
3.601
3.760
3.491
3.286
3.680
3.843
2.529
3.150
3.674
2.658
2.218
1.542
3.113
1.715
3.174
2.609
3.218
1.564
3.047
2.944
3.102
2.948
3.637
2.800
2.196
3.310
2.807
1.555
2.773
3.191
3.265
2.602
3.212
3.261
3.001
3.160
2.892
2.686
3.080
3.243
-- 23 --
Table 2 -- Continued
UID
Stype
i775 > 27.0
MV606
Log(D)
MS
WD MV = 14 WD MV = 17
B0-K7
8186
B9-A3
9006
9020
F0-F2
9212 M0-M1
9230 K2-K3
9331 M5-M6
9351 M1-M2
9397
9959 M0-M1
GAL
yes
yes
no
no
no
no
no
no
no
1.89 ± 5.8
0.94 ± 0.65
3.16 ± 0.44
8.77 ± 0.28
6.40 ± 0.14
12.3 ± 0.45
9.30 ± 0.25
7.038 ± 1.151
6.328 ± 0.131
5.716 ± 0.089
3.979 ± 0.056
3.800 ± 0.028
3.701 ± 0.089
3.974 ± 0.050
3.764
3.697
3.540
2.929
2.279
3.345
3.031
8.77 ± 0.28
4.585 ± 0.056
3.535
3.164
3.097
2.940
2.329
1.679
2.745
2.431
2.935
Table 3. White dwarf candidates in the HUDF (i775 627.0 mag). We list the derived
distance to each object under the assumptions that it is a main sequence star or white
dwarfs (Richer et al. 2000) with the appropriate V − i colors (See Table 1). VT max is the
upper limit of the tangential velocity (VT ) of each object. The seventh column lists the
acceptable nature of the object while the last column gives the 1/Vmax density for that
particular object/scenario. Section 5.2 explains how these quantities were derived in more
detail.
UID
Type
MV606
D
(pc)
Age
(Gyr)
VT max
(km/s)
Possible
Type
1/Vmax
(star/pc−3)
3.7-4.0
Old WD
1147 MS F6-F8
1147 Young WD 14.0-15.0
1147
17.4-18.3
3166 MS K4-K5
6.8-7.1
3166 Young WD 14.8-15.8
Old WD
3166
17.2-18.2
4322 MS A7-K5
2.3-7.1
4322 Young WD 12.1-13.1
17.5-18.4
4322
Old WD
4839 MS G5-K2
5.0-6.3
4839 Young WD 12.5-13.5
17.4-18.3
4839
5921 MS K4-K5
6.8-7.1
5921 Young WD 14.5-15.5
17.3-18.2
5921
7768 MS K5-K7
7.1-7.6
7768 Young WD 14.5-15.5
7768
17.3-18.2
Old WD
9020 MS F0-F2
2.7-3.6
9020 Young WD 13.3-13.3
17.5-18.4
9020
9230 MS K2-K3
6.3-6.7
9230 Young WD 13.6-14.6
9230
17.5-18.4
Old WD
Old WD
Old WD
Old WD
--
2
4
--
1.5 ± 1.1 × 104
110 ± 24
24 ± 5
1.4-3.9
10.3-12.1
3.7 ± 0.23 × 103
80± 18
27 ± 6
2.5-5.7
9.9-11.5
14.1
3.0
10.2
3.5
F6-F8
K4-K5
4.29 ± 3.4 × 105
6.9 ± 1.6 × 103
590 ± 126
1.2 ± 0.3 × 105
2.4 ± 0.5 × 103
417 ± 88
3.7 ± 0.2 × 103
93 ± 20
26 ± 5
4.1 ± 0.5 × 104
1.3 ± 0.3 × 103
345 ± 70
5.2 ± 1.0 × 105
6.2 ± 1.4 × 103
576 ± 123
6.3 ± 0.4 × 103
0.4-1.4
10.7-12.9
1545.6
75.5
Disk or Halo
3.37 × 10−3
1.1-3.3
10.4-12.3
1.9-4.7
10.1-11.7
1.9-4.7
10.1-11.7
0.5-1.6
10.7-12.9
885.
53.4
11.9
3.3
160.2
44.1
793.9
73.7
Disk
Disk or Halo
K4-K5
9.45 × 10−6
1.80 × 10−2
Disk
Disk or Halo
2.1 × 10−5
9.9 × 10−4
Disk or Halo
3.16 × 10−3
K2-K3
187 ± 41
32 ± 7
1.1-3.2
10.4-12.3
23.9
4.0
-- 25 --
0.6
0.4
0.2
0
20
25
30
Fig. 1. -- The distribution of Si775 (green) and Sz850 (red) values for the 10181 HUDF objects.
The stellarity index S was defined (see Section 3) so that higher values correspond to objects
which are increasingly resolved in the images. We selected unresolved objects fainter than
i775 > 20.0 mag to be unresolved if Sz850 60.05, and Sz850 60.15 for brighter objects which
are saturated.
-- 26 --
Fig. 2. -- The distribution of Sz850 values for the brightest 369 objects in the field. Each
object appears 10 times in this figure, with increasingly lower signal to noise (increasingly
darker dots). The observed increase in S values as a function of magnitude at i775 > 29.0 mag
is similar to the one observed in Figure 1 where the real distribution of the S values from the
10181 HUDF objects is shown. This demonstrates the robustness of S to distinguish between
point sources and extended objects down to faint magnitudes. At fainter magnitudes, images
of stars become increasingly dominated by noise which tends to increases the measured values
of S . The same test was performed in the i775 band and yielded an identical behavior.
-- 27 --
4
4
2
2
0
0
10
10
8
8
6
6
4
4
2
2
600
600
500
500
400
400
300
300
200
200
3
3
2.5
2.5
0.8
0.8
0.6
0.6
0.4
0.4
0.2
0.2
0
0
5
5
4.5
4.5
4
4
3.5
3.5
3
3
1
1
0.5
0.5
0
0
50
50
40
40
30
30
20
20
10
10
500
500
400
400
300
300
200
200
550
550
500
500
450
450
400
400
350
350
6000
6000
7000
7000
8000
8000
9000
9000
6000
6000
7000
7000
8000
8000
9000
9000
Fig. 3. -- GRAPES spectra of unresolved objects in the HUDF and the two best matching
templates (solid lines) listed in Table 2
-- 28 --
1
1
0.5
0.5
0
0
60
60
50
50
40
40
30
30
20
20
600
600
500
500
3.5
3.5
3
3
2.5
2.5
2
2
1.8
1.8
1.6
1.6
1.4
1.4
1.2
1.2
0.8
0.8
0.7
0.7
0.6
0.6
0.5
0.5
0.4
0.4
1.4
1.4
1.2
1.2
1
1
0.8
0.8
0.6
0.6
1.5
1.5
1
1
1.4
1.4
1.2
1.2
1
1
0.8
0.8
1.4
1.4
1.2
1.2
1
1
0.8
0.8
0.6
0.6
6000
6000
7000
7000
8000
8000
9000
9000
6000
6000
7000
7000
8000
8000
9000
9000
Fig. 4. -- GRAPES spectra of unresolved objects in the HUDF and the two best matching
templates (solid lines) listed in Table 2
-- 29 --
12
12
11
11
10
10
9
9
6
6
4
4
2
2
0.6
0.6
0.4
0.4
0.2
0.2
8
8
6
6
4
4
2
2
100
100
50
50
0.4
0.4
0.2
0.2
0
0
200
200
180
180
160
160
140
140
120
120
10
10
9
9
8
8
7
7
6
6
1
1
0.5
0.5
0
0
0.4
0.4
0.3
0.3
0.2
0.2
0.1
0.1
0
0
6000
6000
7000
7000
8000
8000
9000
9000
6000
6000
7000
7000
8000
8000
9000
9000
Fig. 5. -- GRAPES spectra of unresolved objects in the HUDF and the two best matching
templates (solid lines) listed in Table 2. The bottom 4 objects are the extra-galactic sources
4120, 6732, 8157, 9397 estimated to be at redshifts of z = 2.1, z = 3.2, z > 3.0, z = 3.0,
respectively (Xu 2004).
-- 30 --
20
25
30
0
1
2
Fig. 6. -- The location of the 28 spectroscopically identified sources in a magnitude color plot.
The small, dark points are the entire set of objects in the HUDF catalog with i775 6 29.0
mag and z850 6 28.2 mag. Solid red squares are M dwarfs. Empty red squares are the two L
dwarf candidates. Solid green triangles represent the non M or L dwarfs stars which could
be main-sequence stars. The empty blue circles are white dwarf candidates (Table 3). The
solid yellow circles represent the QSOs.
-- 31 --
18
20
22
24
26
28
30
0
1
2
3
4
Fig. 7. -- The location of the 28 spectroscopically identified sources in a magnitude color
plot. The symbols are the same as in Figure 6.
-- 32 --
Fig. 8. -- Left panel: observed shifts (scaled by a factor of 500) of the 360 most compact
sources in the HUDF between Epoch 1 to Epoch 2, i775 band, images; Right panel: cor-
rected shifts, obtained after applying a correction based on a fit of the raw measurements
(right panel, see Section 3.2). North is up on these panels. The axes units are UDF pixels
(0.030"/pixel).
-- 33 --
O5V
A0V
F0V
G0V K0V
M0V
4.9
T4000K
T5000K
4.0
2.0
1.0
3.0
T3000K
0.1
0
1
2
2
1.5
1
0.5
0
-0.5
Fig. 9. -- (V606-i775) vs. (B435-V606) color-color plot of i775 6 27.0, unresolved, spectroscop-
ically identified objects in the HUDF. M and L dwarfs are shown using red squares with
photometric error bars. Main-sequence stars are shown in green thick circles. Extra-galactic
objects, including those dimmer than i775 = 27.0 mag, are shown using large blue stars. The
5 white dwarf candidates are shown using large blue crosses. Pickles main-sequence objects
are also shown using small black circles. The (B435-V606) locations of the O, A , F, G, K, and
M stellar type are shown for reference. Hot white dwarfs are shown in large blue squares.
Cool white dwarfs ((3000◦K, 4000◦K, 5000◦K), Harris et al. 2001) are shown using labeled,
empty circles. Finally, the solid black line shows the QSO track, from redshift of z=4.0 (top
right) to z=0.1 (bottom left)
-- 34 --
O5V
A0V F0V
K0V
M0V
0.5
0
-0.5
T3000K
T5000K
T4000K
4.0
1.0 2.0
3.0
4.9
0.1
0
1
2
Fig. 10. -- (i775-z850) vs. (V606-i775) color-color plot of i775 6 27.0 mag, unresolved, spectro-
scopically identified objects in the HUDF. The labels and symbols used are the same as in
Figure 9.
-- 35 --
Fig. 11. -- Predicted and observed cumulative luminosity functions of stars with spectral
types M4 and later. This plot shows the expected distribution from the Galactic disk and
halo. Four different disk scale heights are plotted with h0 values of 200, 300, 400 and
500 pc (bottom curves to top curves respectively). For each value of h0, Mass Functions
corresponding to α = 0.5 and 1.5 are shown (solid and dashed lines respectively). The
observed HUDF distribution is shown using a solid black line which is best fit by a disk with
a scale height of h0 = 400pc while reasonably excluding values of h0 = 200,300, and 500.
|
astro-ph/0511574 | 1 | 0511 | 2005-11-18T18:20:17 | Metallicity and colours in galaxy pairs in chemical hydrodynamical simulations | [
"astro-ph"
] | Using chemical hydrodynamical simulations consistent with a Lambda-CDM model, we study the role played by mergers and interactions in the regulation of the star formation activity, colours and the chemical properties of galaxies in pairs. A statistical analysis of the orbital parameters in galaxy pairs (r <100 kpc/h) shows that the star formation (SF) activity correlates strongly with the relative separation and weakly with the relative velocity, indicating that close encounters (r <30 kpc/h) can increase the SF activity to levels higher than that exhibit in galaxies without a close companion. Analysing the internal properties of interacting systems, we find that their stability properties also play a role in the regulation the SF activity (Perez et al 2005a). Particularly, we find that the passive star forming galaxies in pairs are statistically more stable with deeper potential wells and less leftover gas than active star forming pairs. In order to compare our results with observations, we also build a projected catalog of galaxy pairs (2D-GP: rp <100 kpc/h and Vr <350 km/s), constructed by projecting the 3D sample in different random directions. In good agreement with observations (Lambas et al 2003), our results indicate that galaxies with rp < 25 kpc/h (close pairs) show an enhancement of the SF activity with respect to galaxies without a close companion. (Abridged.) | astro-ph | astro-ph |
Asociaci´on Argentina de Astronom´ıa
BAAA, Vol. 48, 2005
E.M. Arnal, A. Brunini, J.J. Clari´a, J.C. Forte, D.A. G´omez, D. Garc´ıa Lambas,
Z. L´opez Garc´ıa, Stella M. Malaroda, & G.E. Romero, eds.
COMUNICACI ´ON DE TRABAJO -- CONTRIBUTED PAPER
Metallicity and colours in galaxy pairs in chemical
hydrodynamical simulations.
Josefa P´erez
Facultad de Ciencias Astron´omicas y Geof´ısicas, UNLP, Argentina,
[email protected]
Patricia Tissera
Instituto de Astronom´ıa y F´ısica del Espacio, Buenos Aires, Argentina,
[email protected]
Diego Garcia Lambas
Observatorio Astron´omico de la Universidad Nacional de C´ordoba,
Argentina, [email protected]
Cecilia Scannapieco
Instituto de Astronom´ıa y F´ısica del Espacio, Buenos Aires, Argentina,
[email protected]
Abstract. Using chemical hydrodynamical simulations consistent with
a Λ-CDM model, we study the role played by mergers and interactions
in the regulation of the star formation activity, colours and the chemical
properties of galaxies in pairs. A statistical analysis of the orbital param-
eters in galaxy pairs (r < 100 kpc h−1) shows that the star formation
(SF) activity correlates strongly with the relative separation and weakly
with the relative velocity, indicating that close encounters (r < 30 kpc
h−1) can increase the SF activity to levels higher than that exhibit in
galaxies without a close companion. Analysing the internal properties
of interacting systems, we find that their stability properties also play a
role in the regulation the SF activity (P´erez et al 2005a). Particularly,
we find that the passive star forming galaxies in pairs are statistically
more stable with deeper potential wells and less leftover gas than active
star forming pairs.
In order to compare our results with observations,
we also build a projected catalog of galaxy pairs (2D-GP: rp < 100 kpc
h−1 and ∆V < 350 km s−1), constructed by projecting the 3D sam-
ple in different random directions. In good agreement with observations
(Lambas et al 2003), our results indicate that galaxies with rp < 25 kpc
h−1 (close pairs) show an enhancement of the SF activity with respect
to galaxies without a close companion. All the properties studied for
galaxy pairs are analysed in the 2D and 3D simulated catalogs, allowing
us to assess the contamination level introduced by spurious pairs. Con-
sistently with observational estimations (Nikolic et al 2004), we find that
1
2
J. P´erez, P. Tissera, D.G. Lambas & C. Scannapieco
the percentage of spurious pairs decreases with the relative separation,
representing almost a 30% for the 2D-GP and a 19% for close pairs. We
also analyse the environmental effect on the star formation (SF) activity
for both, pairs and isolated galaxy samples, finding the expected SFR-
local density relation (Gomez et al 2003), with a significantly stronger
dependence for close pairs. Finally, we analyse the colour and chemical
properties of galaxies in pairs in order to investigate the effect of inter-
actions on the bimodal colour distribution observed in galaxies (Balogh
et al. 2004; Tissera et al. 2005) and the mass-metallicity relation for the
stellar population (Tremonti et al. 2004).
Resumen. Usando simulaciones qu´ımicas hidrodin´amicas consistentes
con un modelo Λ-CDM, estudiamos el rol de las interacciones y colisiones
de galaxias sobre la actividad de formaci´on estelar (SF), los colores y las
propiedades qu´ımicas de las mismas. El an´alisis estad´ıstico de los efectos
de las interacciones entre galaxias sobre la actividad de formaci´on este-
lar como funci´on de los par´ametros orbitales, muestra que los encuentros
cercanos (r < 30 kpc h−1) pueden inducir una actividad de formaci´on
estelar superior a la encontrada en sistemas aislados. Sin embargo, en-
contramos que la estabilidad de los sistemas gal´acticos tambi´en juega un
rol fundamental en la regulaci´on de este proceso (P´erez et al 2005). As´ı,
las galaxias en pares con baja actividad de SF tienden a ser m´as estables
(con pozos de potencial m´as profundos) que las galaxias en pares con alta
taza de formaci´on estelar. A fin de comparar nuestros resultados con los
observacionales, construimos un cat´alogo proyectado de pares de galaxias
(2D-GP), siguiendo los criterios de selecci´on utilizados por Lambas et al
(2003): rp < 100 kpc h−1 and ∆V < 350 km s−1. Consistentemente con
las observaciones, encontramos que los pares cercanos (rp < 25 kpc h−1)
muestran niveles de SF superiores a aquellos correspondientes a galaxias
aisladas. Comparando los cat´alogos 2D-GP y 3D-GP, estimamos los posi-
bles efectos de proyecci´on sobre los resultados observacionales, analizando
la contribuci´on de los pares espurios en las simulaciones. De acuerdo
con estimaciones observacionales (Nikolic et al 2004), nuestros resultados
muestran que el porcentaje de pares espurios disminuye con la distan-
cia relativa, encontrando un 30% para 2D-GP y un 19% para los pares
cercanos. Por otro lado, analizamos los efectos ambientales sobre la ac-
tividad de formaci´on estelar, tanto para las galaxias en pares como para
aquellas sin un vecino cercano. Encotramos, en buen acuerdo con las
observaciones (Gomez et al 2003), que existe una relaci´on entre la SF y
la densidad local para ambas muestras con una mayor dependencia para
los pares cercanos. Finalmente, investigamos los efectos que producen las
interacciones de galaxias sobre el color y las propiedades qu´ımicas de las
mismas, con el fin de evaluar su rol en la determinaci´on de la distribuci´on
bimodal de colores observada (Balogh et al. 2004; Tissera et al. 2005)
y la relaci´on masa-metalicidad de la poblaci´on estelar (Tremonti et al.
2004).
Metallicity and colours in galaxy pairs.
3
1.
Introduction
Observations show that mergers and interactions can induce star formation (SF)
activity in galaxies (e.g. Larson and Tinsley 1978). Barton et al. (2000) and
Lambas et al. (2003, LTAC03) analysed a sample of pair galaxies in the field
finding a clear correlation between the proximity in projected distance and ra-
dial velocity of two galaxies and their SF activity. On the other hand, numerical
simulations of pre-prepared mergers showed that interactions between axisym-
metrical systems without bulges might induce gas inflows to the central region of
the systems, triggering starburst episodes (Mihos and Hernquist 1996). Results
from the study of the effects of mergers in the SF history of galactic objects in
cosmological hydrodynamical simulations (Tissera et al. 2002), indicate that,
during some mergers events, gaseous disks could experience two starbursts de-
pending on the characteristic of the potential well.
Recently, observational results obtained by Balogh et al (2004) confirm a bi-
modal colour distribution, which segregates galaxies into blue and red popula-
tions. They find that this bimodality is well fitted using two Gaussian distribu-
tions. Analysing this colour distributions for an important range of magnitudes
and densities, they find that while the characteristics of the Gaussians depends
strongly with luminosity, seems not to change with the local density. According
to Balogh et al (2004) this invariability with the environment suggest that the
process of transforming blue into red galaxies has to be very efficient and to
overcome the effects of environment.
In this work, we will investigate how effective interactions and mergers of galaxies
are in the regulation of SF activity and if they are responsible in determining
the other internal properties of galaxies like colours and chemical abundances.
If the Universe is consistent with a hierarchical scenario, then, interactions and
mergers are of utmost importance to understand the effects and efficiency that
these physical processes have on the life of galaxies. Thus, we will focus on
a statistical analysis of galaxy pairs in a hierarchical scenario with the aim
at confronting this scenario with recent observational results of galaxy pairs.
Details can be found in P´erez et al. (2005a, 2005b).
2. Results
We have analysed a Λ-CDM simulation (Λ = 0.7, Ω = 0.3, H = 100h kms−1Mpc−1
with h = 0.7) run with the chemical cosmological GADGET-2 (Scannapieco et
al. 2005). From this simulation we constructed the 2D-GP and 3D-GP cata-
logs. In order to unveil the effects of interactions, we also build the respective
galaxy control samples defined by galaxies without a close companion within
the corresponding thresholds in relative separation and velocity used to define
each GP catalogs (P´erez et al 2005a). For each simulated galaxy, we estimated
the stellar birthrate parameter b, defined as the present level of SF activity of
a galaxy normalized to its mean past SF rate and the absolute magnitudes in
different wavelenghts (De Rossi et al 2005).
The analysis of the SF activity for galaxies in the 3D-GP catalog as a function
of their orbital parameters shows that close encounters (pairs with a relative
separation less 30 ± 10 kpc h−1) can enhance the SF activity at higher levels
4
J. P´erez, P. Tissera, D.G. Lambas & C. Scannapieco
than those measured for galactic systems without a close companion. On the
other hand, the SF activity seems to correlate more weakly with the relative
velocities. However, we also find almost a 50% of passive star forming galaxies
in close pairs, suggesting that the internal properties of these systems are also
present in the SF process. Effectively, we find that the triggering of SF by tidal
interactions is also regulated by the stability of the systems and the gas reservoir.
The analysis of the passive SF close pairs show that part of these systems have
experienced recent star formation activity and the rest shows deeper potential
well and are less leftover gas than galaxies in pairs with strong SF activity.
When the projected catalog of pair is analysed, the same global trends detected
in the 3D-GP sample are found, although a shrinking in the enhancement thresh-
old in projected distance is observed. It value drops with respect to that found in
3D to ∼ 25 ± 5 kpc h−1 in good agreement with observational results (LTAG03).
This shrinking in the threshold is produced by both geometrical projection ef-
fects and spurious pairs. In order to separate these both effects, we have removed
spurious pairs from the 2D-GP sample by checking their 3D relative separations.
Consistently with previous works (Alonso et al. 2004), we found that 30% of
the pairs in 2D-GP sample are spurious. This percentage reduces to 19% for 2D
close pairs (rp < 25 kpc h−1 and ∆cz < 100 km s−1).
Analysing the dependence of the SF on environment with the projected local
density parameter (defined like in the observational analysis as Σ = 6/(πd2
6),
with d6 the projected distance to the 6th neighbor brighter than Mr = −20.5),
we detect the expected SFR-local density relation (Gomez et al 2003) for both
galaxies in pairs and without a near companion, with a stronger dependence
for close pairs. The important decrease in the SF activity and the significant
increase of the fraction of passive SF members from low to high density regions
for galaxies in close pairs suggest that interactions might a relevant role in the
origin of the SF-density relation.
In order to infer the effect of interactions on the colour distribution of galaxies,
we also evaluate the u−r colour for the simulated galaxies in pairs compared with
that found for galaxies without a near companion. According results obtained
by Balogh et al (2004), we adopt the value u − r = 1.8 to segregate red and
blue galaxies. We found that the mean colours of the blue and red peaks for
pairs are at < u − r >≈ 1.60 and < u − r >≈ 2.06, with similar values for the
control sample. However, while pairs exhibit a clear bimodal colour distribution
with a 58 per cent of galaxies in the blue peak, the control sample is more
consistent with an unimodal distribution with an excess of red systems (26 per
cent of galaxies in the blue peak). Comparing the simulated control sample with
observations (Balogh et al 2004), we also find an excess of the fraction of red
galaxies. Part of this red excess might be produced by the high efficiency in
the transformation of gas into stars of our simulations which is not regulated
by supernova energy feedback. However, tidal torques generated by interactions
can compress the leftover gas in a short-time producing strarbursts which might
be the responsible of producing the bimodal distribution found for galaxies in
pairs. To gain insight in this analysis, we divide the galaxy pair catalog into
merging (r < 30 kpc h−1) and interacting (30 kpc h−1 < r < 25 kpc h−1)
pairs. Although the bimodal distribution is present in both subsamples, the
fraction of blue galaxies is higher than the red one for the merging systems,
Metallicity and colours in galaxy pairs.
5
with the opposite result for the interacting pairs. This results is consistent with
the present level of SF activity found for merging and interacting systems. The
former has a 36% of active SF galaxies (SF activity higher than for the control
sample), while the latter has only a 10% of active SF galaxies. The SF activity
in the recent past for the currently passive SF systems is also responsible for the
colour distribution of galaxies. As shown in P´erez et al (2005b), we find that the
fraction of currently passive SF systems which have experienced strong activity
in the recent 0.5 Gyr (F∗) anticorrelates with the u − r colour. The contribution
to the blue colours comes from galaxies that independently of their current SF
activity, have experienced an strong SF activity in the recent last 0.5 Gyr.
Finally, we analyse the chemical properties of the interstellar medium (ISM) and
the stellar population (SP) of galaxies in pairs. While for SP exhibits a clear
excess of their chemical abundance respect to that found for galaxies without
a near companion, the ISM enriched by the new stars, has different levels of
chemical contamination depending on the SF activity of the galaxy. Particularly,
if we segregate galaxies into active and passive according their current SF level,
we find that SP of passive SF galaxies are more enriched than the active ones. On
the contrary, ISM of passive galaxies are less enriched than the ISM of the active
ones. Interesting, we also find that the distribution of the chemical abundance
of the ISM of these systems has a similar behaviour with the relative separation
between the members of the pair than that found for the SF activity. In other
words that the ISM of currently passive SF pairs shows an enhancement of their
chemical enrichment respect to that measured in the control sample for very
close systems (rp < 25 kpc h−1). The analysis of the recently past SF activity
of these galaxy pairs shows that, although these systems are currently passive
SF ones, they have experienced strong SF activity in the recent past which has
contributed to enhance the chemical abundance in their ISM.
3. Conclusions
From the analysis of the 3D-simulated galaxy pair catalog, we conclude that close
galaxy interactions (r < 30 kpc h−1) can be correlated with an enhancement of
SF activity at higher levels than those measured for galactic systems without a
close companion. We also found that the internal dynamical stability of galactic
systems plays an important role as it can be deduced from the presence of an
anticorrelation signal between the deepness of the potential well and the star
formation activity.
The construction of a projected galaxy pair catalog allowed us, firstly, to make
a suitable comparison with observations and then, to analyse the contamination
effects introduced by spurious pairs. In a good agreement with observational
results, we find that all trends observed for the 3D-GP catalog are reproduced
in a similar way for the 2D-GP one. On the other hand, the analysis of the
spurious pairs shows that the contamination level increases with the relative
separation. Consistently with previous works (Alonso et al. 2004), we find that
almost a 30% of the pairs in 2D-GP sample are spurious.
The environmental effects on the interacting pairs were analysed, defining the
local density using the projected distance to the 6th. We find a dependence
of the SF on local density which is consistent with the observed SF-density
6
J. P´erez, P. Tissera, D.G. Lambas & C. Scannapieco
relation (Gomez et al 2003). It yields a decrease in the level of SF activity and
an increase in the fraction of passive SF systems with increasing local density.
Although, this relation is found for both pairs and control samples, we find
that it is significantly stronger for close pairs suggesting the fundamental role of
interaction in driving the SF-density relation.
The analysis of the colours and chemical properties of galaxies also seem to
be regulated by the interactions. We find a clear bimodal colour distribution
for galaxies in pairs with a blue peak populated basically by galaxies with an
strong, currently or recently past SF activity. Finally, analysing the chemical
properties of galaxy pairs, we find that while SP seems to store information
about the history of mergers and interactions, the chemical abundance of the
ISM seems to be contaminated by the new stars, reflecting in this way the effects
of the present interactions. Our results are in very good agreement with recent
observational works. Regarding the bimodal colour distribution, we found that
although interactions and merger are able to explain the bimodal colour distri-
bution, supernova energy feedback seems to be required to prevent an excessive
transformation of gas into stars and to obtain the detail characteristics of this
distribution.
Acknowledgments. This work was partially supported by the Consejo Na-
cional de Investigaciones Cient´ıficas y T´ecnicas and Fundaci´on Antorchas. Sim-
ulations were run on Ingeld PC Cluster funded by Fundaci´on Antorchas. We
thank the LOC of this meeting for their help made our participation possible.
Patricia B. Tissera thanks the Aspen Center for Astrophysics for the hospitality
during the Summer Workshop 2004.
References
Alonso, M. S., Tissera, P. B., Coldwell, G., Lambas D. G. 2004,MNRAS, 352,1081.
Balogh M. L., Eke V., Miller C., et al., 2004, MNRAS, 348, 1355.
Barton E. J., Geller M. J., Kenyon S. J., 2000, ApJ, 530, 660.
Lambas, D. G., Tissera, P. B., Alonso, M. S. Coldwell, G. 2003, MNRAS,
346,1189 (LTAG03).
Larson, R. B., Tinsley, B. M., 1978, ApJ, 219, 46.
Mihos, J. C., Hernquist, L., 1996,ApJ, 464, 641.
Navarro J.F. & White S.D.M., 1994, MNRAS 267, 401.
Nikolic B., Cullen H., Alexander P., 2004, MNRAS 355, 874.
P´erez, M.J., Tissera P.B., Lambas, D. G., Scannapieco, C. 2005, A&A accepted
(astro-ph/0510327)
P´erez, M.J., Tissera P.B., Lambas, D. G., Scannapieco, C. 2005,
Scannapieco C., Tissera P.B., White S.D.M. & Springel V., 2005, MNRAS ac-
cepted (astro-ph/0505440)
Tissera P.B, Dom´inguez-Teneiro R., Scannapieco C. & S´aiz A., 2002, MNRAS,
333, 327.
|
0804.0589 | 1 | 0804 | 2008-04-03T16:38:46 | Comment on 'Discreteness Effects in Lambda Cold Dark Matter Simulations: A Wavelet-Statistical View' by Romeo et al | [
"astro-ph"
] | Essential equivalence of the conclusions of Romeo et al. (arXiv:0804.0294v1) with earlier work is pointed out. A possible general explanation for these conclusions is suggested. | astro-ph | astro-ph | Comment on “Discreteness Effects in Lambda Cold
Dark Matter Simulations: A Wavelet-Statistical View”
by Romeo et al.
Adrian L. Melott
Department of Physics and Astronomy
University of Kansas
[email protected]
Abstract: Essential equivalence of the conclusions of Romeo et al.
with earlier work is pointed out. A possible general explanation for
these conclusions is suggested.
Introduction and Conclusions
Romeo et al. (2008) performed simulations with truncated power spectra in order
to investigate the effect of small—scale power in initial conditions in numerical
simulations. This approach was introduced in Melott & Shandarin (1989); see
also Melott & Shandarin (1990), Beacom et al. (1991), Melott & Shandarin
(1993).
They conclude that discreteness effects are visible in the simulations on all
scales ε < 2d where ε is the force resolution (sometimes called “softening”), and
d is the mean interparticle separation. Although there is no sharp and sudden
disappearance of discreteness effects, this conclusion is in disagreement with
the usual claims of viability down to the scale ε. Furthermore, although not noted
by Romeo et al., this is essentially the same as conclusions previously discussed
by Peebles et al. (1989), Melott & Shandarin (1989), Melott (1990), Melott &
Shandarin (1990), Beacom et al. (1991), Melott & Shandarin (1991), Kuhlman et
al. (1996), Melott et al. (1997), and Splinter et al. (1998). Specific numerical tests
of discreteness are described in the first of these, and are the primary topic of the
last three; the others present visual comparisons and use truncated power
spectra for comparison purposes.
The suppression of discreteness is the basis for the choice of PM codes over
direct N-body and P3M methods, since it maximizes reliable resolution for given
computer capacity compared with those approaches. Such suppression is
particularly important where discreteness can easily spoil results (e.g. Melott
1982). This, for example, made it possible to discern for the first time filamentary
superclusters in CDM (Melott et al. 1983). As noted by Romeo et al., adaptive
mesh refinement may make it possible to improve upon this, but AMR needs
careful, critical tests, and will still exclude initial fluctuations below the comoving
initial particle Nyquist scale.
Discussion
A possible explanation for this limitation based on the idea of chaotic trajectories
was proposed by Melott et al. (1997). Chaotic trajectories were analyzed by Suto
(1991) as a function of time and the ratio ε/d. He found that the Lyapunov
exponent ~ε-1.2. The criterion that uncertainties in trajectory on scale ε not grow
larger than scale d in a Hubble Time leads to the condition ε ≥ d; for more details
see these two papers.
References
Beacom, J.F., K.G. Dominik, A.L. Melott, S.P. Perkins, and S.F. Shandarin,
“Gravitational Clustering in the Expanding Universe: Controlled High-Resolution
Studies in Two Dimensions”, Astrophysical Journal 372, 351 (1991)
Kuhlman, B., A.L. Melott, and S.F. Shandarin, ”A Test of the Particle Paradigm in
N-Body Simulations”, Astrophysical Journal Letters 470, L41 (1996)
Melott, A.L. “The Formation of Galactic Halos in the Neutrino-adiabatic Theory”
Nature 296, 721 (1982)
Melott, A.L., J. Einasto, E. Saar, I. Suisalu, A. Klypin and S. Shandarin “Cluster
Analysis of the Nonlinear Evolution of Large Scale Structure in an
Axion/Gravitino/Photino Dominated Universe” Physical Review Letters 51, 935
(1983)
Melott, A.L, and S.F. Shandarin, “Gravitational Instability with High Resolution”
Astrophysical Journal 343, 26 (1989)
Melott, A.L. “More Resolution Isn't Always Better Resolution”, Comments on
Astrophysics 15, 1 (1990)
Melott, A.L,, and S.F. Shandarin, “Generation of Large-Scale Cosmological
Structures by Gravitational Clustering” Nature 346, 633 (1990)
Melott, A.L. and S.F. Shandarin, “Controlled Experiments in Cosmological
Gravitational Clustering”, Astrophysical Journal 410, 469 (1993)
Melott, A.L., S.F. Shandarin, R.J. Splinter, and Y. Suto, “Demonstrating
Discreteness and Collision Error in Cosmological N-body Simulations of Dark
Matter Gravitational Clustering”, Astrophysical Journal Letters 479, L79 (1997)
Peebles, P.J.E., A.L. Melott, M.R. Holmes, and L.R. Jiang, “A Model for the
Formation of the Local Group”, Astrophysical Journal 345, 108 (1989)
Romeo, A.B., O. Agertz, B. Moore, and J. Steidel, “Discreteness Effects in
Lambda Cold Dark Matter Simulations: A Wavelet-Statistical View”, preprint
(2008) arXiv:0804.0294v1
Splinter, R.J., A.L. Melott, S.F. Shandarin, and Y. Suto, “Fundamental
Discreteness Limitations of Cosmological N-Body Clustering Simulations”,
Astrophysical Journal 497, 38 (1998)
Suto, Y. “On the stochasticity of gravitational many-body systems in cosmology”,
Publ. Astron. Soc. Japan. 43, L9 (1991)
|
astro-ph/0207434 | 1 | 0207 | 2002-07-19T21:04:29 | Carbon Isotopic Abundances in the Red Giants of Omega Centauri (NGC5139) | [
"astro-ph"
] | Carbon-12 and carbon-13 abundances are measured in eleven red-giant members of the globular cluster Omega Centauri via observations of first-overtone CO bands near 2.3 microns. The mean value for the entire sample is <12C/13C>= 4.3 +/-0.4, with nine giants equal, within the errors, to the equilibrium ration of 12C/13C= 3.5. No correlation is found within Omega Cen between 12C/13C and the abundance of iron. The relation between 12C/13C and other abundance ratios, such as [O/Fe], {Na/Fe], or [Al/Fe] are also discussed. | astro-ph | astro-ph |
Carbon Isotopic Abundances in the Red Giants of ω Centauri (NGC 5139)
Department of Physics, University of Texas at El Paso, El Paso, TX 79968
Verne V. Smith
Donald M. Terndrup1
Department of Astronomy, The Ohio State University, Columbus, OH 43210
and
Nicholas B. Suntzeff
Cerro Tololo Inter-American Observatory, National Optical Astronomy Observatories, Casilla
603, La Serena, Chile
ABSTRACT
Carbon-12 and carbon-13 abundances have been measured in eleven bright giant
members of the globular cluster ω Centauri via observations of the first-overtone CO
bands near 2.3 µm. The stars in this sample were selected to span a substantial fraction
of the range of iron abundances found in this cluster. In addition, the sample covers
a range of [O/Fe], [Na/Fe] and [Al/Fe] abundance ratios derived in previous studies.
In all ω Cen giants the 12C/13C abundance ratio is found to be quite low, indicating
deep mixing in these red giants. The mean value for the entire sample is h12C/13Ci =
4.3 ± 0.4 (σ = 1.3), with nine stars equal, within the errors, to the equilibrium ratio
12C/13C = 3.5 and two stars having slightly higher values. There is no correlation
between the 12C/13C and the abundance of iron. In addition, no correlation of 12C/13C
with [12C/Fe] is found (all giants are deeply mixed), although the derived abundances
of [12C/Fe] show a positive correlation with [O/Fe], and an anticorrelation with [Na/Fe]
(with the oxygen and sodium abundances taken from previous studies in the literature).
A comparison of the isotopic carbon ratios in ω Cen with those from other globular
clusters (M4, M71, NGC6752, and 47 Tuc), and with literature oxygen abundances,
may reveal a slight trend of decreasing 12C/13C ratios with decreasing [O/Fe] in the
entire globular cluster sample of red giants. A comparison between 12C/13C and both
[Na/Fe] and [Al/Fe], however, reveals no trend.
Subject headings: globular clusters:
osynthesis, abundances -- stars; abundances -- stars: late-type -- stars: Population II
individual (ω Centauri) -- nuclear reactions, nucle-
1Visiting Astronomer, Cerro Tololo Inter-American Observatory, National Optical Astronomy Observatories,
which are operated by the Association of Universities for Research in Astronomy, Inc., under cooperative agree-
ment with the National Science Foundation.
-- 2 --
1.
Introduction
In the last two decades, it has become clear that evolved giants in metal-poor globular clusters
often exhibit a wide range in the abundances of light elements (Paltoglou & Norris 1995; Norris &
Da Costa 1995b; Pilachowski, Sneden, & Kraft 1996; Shetrone 1996a, b; Kraft et al. 1997; Sneden
et al. 1997; Gonzalez & Wallerstein 1998; Ivans et al. 1999; Smith et al. 2000). These variations
are generally agreed to arise from proton capture chains that convert C and O into N, Ne into Na,
and Mg to Al in the hydrogen-burning layers of red giants. There has been considerable discussion,
however, as to the origin of these abundance patterns (e.g., Cottrell & Da Costa 1981; Kraft 1994;
Norris & Da Costa 1995a; Sneden et al. 1997).
In the "primordial" scenario, the abundance
patterns reflect nucleosynthesis in a prior generation of massive stars; support for this comes from
variations in the C, N, O, Na, Mg, and Al abundances among both main-sequence and turn-off
stars in various globular clusters, as found by Bell, Hesser, & Cannon (1983), Briley, Hesser, &
Bell (1991), Suntzeff & Smith (1991), Briley et al. (1996), Cannon et al. (1998) or Gratton et
al. (2001). In contrast, the "evolutionary" scenario envisions that the products of internal proton
capture are brought to the stellar surface through mixing in the giants now observed in the clusters.
Evidence in favor of this comes from abundance patterns that depend on the evolutionary state of
the stars in a cluster. For example, several systems show a decline in the overall carbon abundance
with increasing stellar luminosity and a corresponding rise in the abundance of nitrogen, a sharp
decline in the ratio of 12C/13C from the presumed primordial value near 90 (i.e., the solar value) to
values as low as the nuclear equilibrium value of 3.5. This equilibrium ratio of 12C/13C= 3.5, which
is set largely by the ratio of reaction rates 12C(p,γ)13N(β+,ν)13C and 13C(p,γ)14N, is insensitive
to temperature and is a well-determined astrophysical limit (see the review by Wallerstein et al.
1997). So-called "standard" stellar models evolving up the first ascent of the red giant branch, e.g.,
Iben (1964) or Charbonnel (1994), predict values to go only as low as 12C/13C∼ 20: the differences
between the observed (low) isotopic carbon ratios and the higher values predicted by the standard
models of stellar evolution has been a longstanding piece of evidence used to invoke extra mixing
(or deeper mixing) in red giants. Quite possibly, in the globular cluster stars, both primordial and
evolutionary chemical evolution occurs, whereby deep mixing in the current giants is superimposed
on preexisting abundance patterns (Briley et al. 1994; Ivans et al. 1999).
Mixing signatures are seen also in metal-poor field giants (Gratton et al. 2000, and references
therein), with some interesting differences found between the field stars and globular cluster giants of
the same overall metallicity. The O -- Na and Mg -- Al anticorrelations found in some of the globular
clusters are not observed in any of the field red giants. One effect noted in the field giants by
Gratton et al.
(2000), and also predicted by current non-standard stellar models (Charbonnel
1994; Wasserburg, Boothroyd, & Sackmann 1995; Charbonnel, Brown, & Wallerstein 1998), is that
the efficacy of first giant branch mixing will increase as the metallicity decreases. The globular
cluster ω Centauri is a site where the various mixing trends (as a function of [Fe/H] and as a
function of Na and Al variations) can be studied in one object.
-- 3 --
The most massive globular in the Milky Way, ω Cen has a wide spread of [Fe/H]2 (Suntzeff
& Kraft 1996; Norris, Freeman, & Mighell 1996, and references therein), unique among globulars,
with the possible exception of M22 (which may show an abundance spread, but much smaller
than in ω Cen).
In ω Cen the distribution of [Fe/H] has a floor near -- 2.0 to -- 1.8, presumably
representing the initial metallicity of the gas out of which the cluster was formed, and shows an
extended tail to higher metal abundances. The abundances of most elements increase with [Fe/H]
and exhibit relatively small scatter, with the exception of O, Na, Al and some other light elements
(Brown & Wallerstein 1993; Norris & Da Costa 1995b; Smith et al. 2000). As in other globular
clusters, the giants of ω Cen have [Na/Fe] and [Al/Fe] abundances which are positively correlated
with each other, but anticorrelated with [O/Fe] in a way that is most easily (but not definitively!)
explained as a product of deep mixing. The abundance patterns of elements heavier than Fe suggest
nucleosynthesis from AGB stars between 1 and 3M⊙, but whether this occurred during formation
of the cluster that involved self-enrichment, or from mergers of fragments with different chemical
abundances, is not yet determined.
To date, only a few measurements of the 12C/13C isotopic ratio exist for ω Cen giants (3 from
Brown & Wallerstein 1993). Their 3 values were all quite low, with 12C/13C= 6, 4, and 4; however,
the metallicities sampled by these 3 giants was limited ([Fe/H]= -1.36, -1.34, and -1.26). The goal
of this work is to derive additional values of 12C/13C in a larger sample of ω Cen giants, in order to
investigate if there are any trends with [Fe/H], or if this somewhat unique globular cluster differs
in its 12C/13C ratios when compared to other clusters.
2. Observations and Data Reduction
The program stars were selected from the high-resolution study of Norris & Da Costa (1995b),
and were chosen to cover a wide range of [Fe/H] at magnitudes where spectroscopy at 2.35µm was
practical with a 4m telescope. Most of these stars also have other metallicity determinations from
various studies employing low- to high-resolution spectra. Table 1 summarizes the available data
for our targets. The first column of that table lists the name of each star from Woolley (1966).
The second and third columns list photometry from several sources as assembled by Suntzeff &
Kraft (1996); for three stars not tabulated in that paper (ROA 150, 155, and 161), we took the
photometry from Cannon & Stobie (1973) and applied a reddening correction of E(B − V ) = 0.11
(Butler, Dickens, & Epps 1978). Infrared photometry from Persson et al. (1980) follows in column
4, while the remaining columns show determinations of [Fe/H] from papers listed in the table, all
of which are evidently on the same metallicity scale (an exception are the values of Francois, Spite,
& Spite (1988), not tabulated here, which are about 0.2 dex lower than in these other studies). We
will use the Norris & Da Costa (1995b) values of [Fe/H] throughout the remainder of this paper.
2Throughout this paper, we will use the normal spectroscopic notation that [A/B] ≡ log10(NA/NB)star −
log10(NA/NB)⊙, for elements A and B.
-- 4 --
In Figure 1 we present a color-magnitude diagram of ω Cen. The small points show cluster
members (≥ 50% membership probability) from van Leeuwen et al. (2000), where the photometry
comes from sources referenced in that paper. The large symbols show photometry for our targets.
Most of our stars are brighter than the magnitudes where the RGB and AGB are clearly separated,
but most should be on the RGB since the evolutionary lifetimes there are much longer than on the
AGB.
An additional selection criteria was to observe stars that spanned the range of Na and Al
abundances, which in ω Cen and other globular clusters are often anticorrelated with oxygen.
Spectra of the first-overtone bands of CO near 2.3 µm were obtained on June 2 -- 4, 1996, and in a
second run on June 10-12, 1998 with the Blanco 4m Telescope and the CTIO infrared spectrometer
(DePoy et al. 1990). The 1996 run used an engineering-grade 256 × 256 InSb detector from SBRC
which had many defective pixels and an appreciable dark current. The 1998 run employed a similar
detector of higher quality, and also was obtained with a tip-tilt secondary. The length of the slit was
16′′, and the slit width was set to 0.7′′. The pixel size for both runs was 24 µm for a resolution of
λ/∆λ ≈ 9900 pixel−1. The effective resolution was obtained from observations of telluric emission
lines, which had an average FWHM of 2.7 pixels. A K filter was used for order separation.
The 1996 observations were obtained in three overlapping wavelength intervals, which are
shown in Table 2. In 1998, we obtained spectra at two wavelength settings. These covered the
(2 -- 0), (3 -- 1), and (4 -- 2) bands of 12CO and the (3 -- 1) band of 13CO and also included features of
Ca and Mg (Kleinman & Hall 1986); the latter are, however, very weak in our spectra of the ω Cen
targets and will not be discussed further. Figure 2 shows two sample calibrated spectra and the
identification of the more prominent features. These two giants have similar effective temperatures
and metallicities, but note that ROA 155 has noticably stronger CO bands than ROA 139. Norris &
Da Costa (1995b) classify ROA 139 as CN-strong and ROA 155 as CN-weak; the difference between
the CN bandstrengths represents differences in the nitrogen abundance (Norris & Da Costa find
[N/Fe]= +1.05 for ROA 139 and +0.40 for ROA 155). The differing CO bandstrengths observable
in Figure 2 fit in with the C -- N anticorrelation expected from differing degrees of CN-cycle mixing.
The observing procedure in both years was to obtain repeated spectra with the star placed at
various locations along the slit. Exposure times ranged from 30 to 60 seconds at each position, and
the motion of the star along the slit between exposures was sufficient to completely separate the
spectra on the detector. In both runs, observations of a hot star (HD 116717, spectral type A0V)
were interleaved with observations of the ω Cen targets. The observing procedure was identical
to that used for the program stars, except that since HD 116717 was brighter, the exposure times
were shorter and fewer positions along the slit were needed in order to achieve the same, or higher,
S/N than in the ω Cen stars. Spectra were obtained for several stars at one wavelength interval,
then repeated at the other wavelength settings.
-- 5 --
Data reduction was performed using scripts written for the IRAF3 and VISTA packages. The
images were first corrected for the nonlinear response of the detectors. The correction multiplied
the raw counts by a quadratic polynomial in intensity. The correction was almost always less
than 0.5%, but was 2 -- 3% at the center of the spectrum for the brightest stars. Next, the several
images for each star were combined via a median, which removed the individual spectra. This
was subtracted from the images, thereby correcting for the sky and the dark current. Finally, the
images were divided by a flat field, which was obtained by combining many images of a white spot
on the telescope dome, obtained in two passes with the dome lights on and off then subtracted.
The resulting spectra were then traced, extracted in an aperture of width twice the FWHM
of the spectra, and co-added after multiplicatively scaling to the average mean exposure. The co-
adding process computed an average with rejection of pixels that were > 4σ away from the average.
The deviation about the mean gave a measure of the S/N per spectrum, which was typically 75 to
120 for the ω Cen spectra.
The final step was to correct for the many telluric absorption lines in the spectra using the
observations of the spectra of the hot star. When hot star spectra were taken both before and
after one of the ω Cen targets, we interpolated the spectra linearly in time to the midpoint of the
target star's observation. If spectra were not available that bracketed an observation in ω Cen , we
always had one adjacent to the observation. We measured the shift in the wavelength direction (in
pixels) between the interpolated spectrum and the combined target spectrum to ensure that they
were aligned correctly. The resulting pixel shifts were almost always less than 0.3 pixel. Then the
object spectrum was divided by the interpolated spectrum.
Because there are few lamp arc lines in the spectral intervals we observed, the wavelength
calibration was achieved in several steps. We first computed the linear wavelength solution for the
bluest observed wavelength interval in the 1996 spectra, in which there are several OH airglow lines
that have wavelengths compiled by Oliva & Origlia (1992). The dispersion (µm pixel−1) from this
solution was assumed to be the same for the other two wavelength intervals. We determined the
zero point for the central wavelength interval by computing the average pixel shift between the two
spectrum sections via cross correlation. Finally, the zero point for the red spectral interval was
adjusted to match the the 3 -- 1, 2 -- 0 and 4 -- 2 bandheads of 12CO, where the wavelengths of these
were taken from Kleinman & Hall (1986). A similar procedure was followed for the 1998 data,
using the airglow lines in the blue spectra to set the dispersion for both wavelength intervals, then
adjusting the zero point for the red spectra to match the 12CO bandheads. The r.m.s. residuals in
the solutions using either the telluric lines or the CO bandheads were always less than 0.4 pixels.
3The IRAF software is distributed by the National Optical Astronomy Observatories under contract with the
National Science Foundation
-- 6 --
3. The Carbon Isotopic Abundance Analysis
An abundance analysis of the carbon isotopes, using the CO features, was carried out for
the program stars using the most recent version of the LTE spectrum synthesis code MOOG
(Sneden 1973). Model atmospheres were computed using a version of the MARCS code described
in Gustafsson et al. (1975), and are plane parallel, hydrostatic equilibrium models. For the ω Cen
giants spanning the range in temperature, gravity, and metallicity found here, the use of MARCS --
type models is appropriate, as discussed in a number of abundance studies, e.g., Smith et al. (2000),
Ramirez et al. (2001), or Cunha et al. (2002). In addition, an analysis by Hinkle & Lambert (1975)
of the formation of the CO vibration-rotation (V -- R) lines in the types of red-giant atmospheres
studied here finds that these lines form to a high degree in LTE.
As mentioned in the Introduction, all of the ω Cen program stars observed here were analyzed
by Norris & Da Costa (1995b), thus detailed abundances from a number of elements are already
available. The stellar parameters of effective temperature (Teff ), surface gravity (log g), micro-
turbulence (ξ), and metallicity (taken as [Fe/H]) were derived by a combination of photometric
and spectroscopic indicators, as discussed by Norris & Da Costa (1995b). A comparison of those
parameters, with those derived independently by Smith et al. (2000) for a subset of the Norris &
Da Costa (1995b) sample, is described in both Smith et al. (2000) and Cunha et al. (2002). No
unexpected systematic errors or uncertainties are found and Teff , log g, ξ, and [Fe/H], as derived by
Norris & Da Costa (1995b) are adopted here. Estimates of uncertainties in the stellar parameters,
as discussed in the various abundance analysis papers mentioned above, are about ±100K in Teff ,
±0.2 dex in log g, ±0.2 km s−1 in ξ, and ±0.05 dex in [Fe/H]. The adopted stellar parameters are
shown in the beginning columns of Table 3.
Spectral -- line lists covering the regions observed were constructed using the atomic linelist
from the Kurucz & Bell (1995) dataset and CO lines from Goorvitch (1994). Lines due to CN
in this spectral region were also included and were kindly provided by B. Plez (2001, provate
communication). For giants with the temperatures and metallicities of our program stars, the CN
lines are a minor contribution, even for very large nitrogen enhancements. In all of the program
stars the C/O abundance ratio is less than 1.0 and CO formation is controlled primarily by the
carbon abundance. Although the derived carbon abundance from CO depends slighty on the
oxygen abundance, this dependence is not large for the temperature regime covered here and
oxygen abundances are available from Norris & Da Costa (1995b), who used the [O I] λ6300A line.
With a given O abundance and model atmosphere, the CO bands were synthesized as a function
of carbon abundance. Synthetic spectra were calculated with wavelength increments of 0.01A (or
λ/∆λ= 2.3×106) and smoothed with a Gaussian broadening function of 6.4A, corresponding to the
measured widths of intrinsically narrow telluric emission lines. Typical internal stellar broadening
mechanisms in these giants (thermal, microturbulence, and macroturbulence) are ∼ 0.8-1.0A at
this wavelength, so instrumental broadening is by far the dominant term. The Gaussian smoothing
function was found to provide excellent fits to the observed spectra. The 12C abundance was
derived first, giving most weight to the regions covering the 12CO (3 -- 1) and (4 -- 2) bandheads, as
-- 7 --
these are free from 13CO contamination and were found to be most sensitive to the carbon-12
abundance. A sample comparison and fit between observed and synthetic spectra is shown in
Figure 3 for the star ROA155 ([Fe/H]= -1.64). The top panel shows the 12CO (3 -- 1) bandhead
for three different 12C abundances. Straightforward least-squares residuals were used to select the
best -- fit carbon abundance, and this abundance was rounded to the nearest 0.1 dex. Residuals
were computed for regions covering ∼23225-23260A for 12CO (3 -- 1) and 23515-23550A for 12CO
(4 -- 2). Each region contains about 17 data points, so the minimum sum of squared residuals was
not affected significantly by a dew deviant points. The bottom panel illustrates the fit for the
13CO (2 -- 0) bandhead in this star, using the already derived 12C abundance; the 13C abundance
is parameterized by the 12C/13C ratio and synthetic spectra computed for three different isotopic
ratios are shown, with 12C/13C= 4.0 being the best fit. The interval for calculating residuals
was ∼23447-23480A containing 16 or 17 data points: the resulting minima were not influenced
significantly by the occasional noisy point. Isotopic ratios were rounded to the nearest 0.5.
The primary uncertainties in the derived carbon isotopic abundances arise from uncertainties
in the input stellar parameters Teff, log g, and microturbulence (the model metallicity has negligible
effect on the derived abundances, unless the change in metallicity is substantial, i.e. factors of 10).
Tests were done for parameter changes of 100K in effective temperature, 0.2 dex in log g, and 0.2 km
s−1 in ξ, with the following results for the change in carbon-12 abundances: ∆Teff= +100K yields
∆C= +0.05 dex, ∆(log g)= +0.2 dex yields ∆C= +0.05 dex, and ∆ξ= +0.2 km s−1 yields ∆C=
-0.04 dex. An additional uncertainty arises from the S/N of the spectra, and based on the depth
and width of residual minima as a function of carbon abundance, a typical 1σ uncertainty is ±0.04
dex in ∆C. Compared to all of these errors, broadening adds an insignificant uncertainty as the
total integrated absorption is conserved and red giants with these temperatures and abundances
hasve flux points close to the continuum blueward of the 12CO (3-1) bandhead. Adding these
abundance errors in quadrature provides an abundance uncertainty of ±0.09 dex for carbon. The
isotopic ratios are very insensitive to the model parameters, but a conservative estimate based on
testing various fits indicates that the ratio itself is uncertain by ±1.0 in 12C/13C. For all program
stars here, Norris & Da Costa (1995b) derived 12C abundances from the violet CH bands, and a
comparison is shown (with both our estimated errors and their estimated errors) in Figure 4. The
comparison is quite good, with the mean difference (Us -- Norris & Da Costa) being +0.14 ±0.24
dex. One star, ROA 139, is the only giant to show a discrepancy (we find [C/Fe]= -1.13 while
Norris & Da Costa derive -0.50). It is worth noting that this star has the most extreme nitrogen
abundance found by Norris & Da Costa, with [N/Fe]=+1.05. In a plot of [N/Fe] versus [C/Fe],
assembled from the Norris & Da Costa abundances, ROA 139 stands out as falling well above their
mean relation (i.e., quite a large [N/Fe] ratio for their value of [C/Fe]). Simply extrapolating their
[N/Fe] -- [C/Fe] relation to a large value of [N/Fe] would suggest that ROA 139 should have a very
low [C/Fe] ratio (≤ -1.0), as we find. It may be that at very low carbon abundances, the violet
CH-band becomes quite weak.
The Norris & Da Costa estimated uncertainties of ±0.21 dex, coupled with our uncertainties
-- 8 --
of ±0.10 dex (in [C/Fe], assuming an uncertainty of 0.05 dex in Fe), yield an expected scatter of
0.23 dex, close to the observed value of 0.24 dex. The small offset of +0.14 dex in the absolute
abundance scale probably arises from small differences in the gf-value scales between CH and CO,
as well as a possible small difference in the adopted solar carbon abundance. A recent value for
the solar carbon abundance is by Holweger (2001), who finds log ǫ(C)= 8.59. This abundance can
be checked from a solar analysis using the same CO lines as those used in our ω Cen analysis. We
generated a MARCS solar model (Teff = 5777K, log g= 4.438) with a microturbulent velocity of 1.0
km s−1 (this value of ξ yields our preferred iron abundance of log ǫ(Fe)=7.50) from an analysis of
both Fe I and Fe II lines with well-defined laboratory oscillator strengths, and is the same solar Fe
abundance adopted by Norris & Da Costa (1995b), whose ω Cen Fe-abundance scale is used here).
The solar flux spectrum used is that from Wallace & Hinkle (1996), which has a spectral resolution
of λ/∆λ= 261,000. Syntheses of the same regions used in the ω Cen analysis (23225-23260A and
23515-23550A) were examined and the best solar 12C abundance derived from the 12CO lines (using
a MARCS model) is logǫ(C)= 8.55: this is close enough to the Holweger (2001) value that we will
adopt 8.59 as the solar 12C abundance. the two respective solar Fe abundances are the same, with
log ǫ(Fe)= 7.50). This comparison indicates that the carbon isotopic abundances presented here
do not suffer from significant systematic effects. The carbon-12 abundances, as well as the 12C/13C
ratios are listed in Table 3. The [Fe/H] values from Norris & Da Costa (1995b) are used to also
tabulate values of [12C/Fe].
4. Results and Discussion
An initial investigation into the isotopic carbon abundance ratios in these ω Cen giants is
illustrated in Figure 5. In the top panel, 12C/13C ratios are plotted as a function of [Fe/H] for
various samples of metal-poor giants. A number of globular clusters have been analyzed for isotopic
carbon ratios and we include in this discussion the results from Smith & Suntzeff (1989) and Suntzeff
& Smith (1991) for M4 (39 giants), Suntzeff & Smith (1991) for NGC6752 (12 giants), Bell, Briley,
& Smith (1991) for 47 Tuc (4 giants), and Briley et al. (1997) for M71 (10 giants). In addition,
samples of field stars have been analyzed and new results, as well as a summary of earlier work,
have been presented most recently by Keller, Pilachowski, & Sneden (2001): these results are also
included in the comparisons. Concentrating first on the ω Cen giants, there is no significant trend
of 12C/13C with [Fe/H], and all member giants show evidence of deep mixing. Two of the more
metal-poor ω Cen giants have slightly larger ratios (6 for ROA 252 and 7 for ROA 58), however, the
overall carbon abundances in these particular stars are quite low and the CO bands weaker. The
13CO bands in these two giants are weaker than in the other program stars, thus the uncertainties
are probably somewhat larger: before a trend could be claimed, a larger sample of ω Cen giants
would need to be studied. With the current results, no real trend with [Fe/H] is found for the ω
Cen members.
In general, the ω Cen results look very similar to what is found for the other globular clusters,
-- 9 --
with the following note: the two fairly metal-rich clusters, M71 and 47 Tuc, appear to have sub-
stantial fractions of their members with both slightly high (12C/13C ∼ 7-8) and low (12C/13C ∼ 4)
ratios. Briley et al. (1997) have shown that these ratios are anticorrelated with CN band strengths
(the CN-strong giants have lower 12C/13C ratios and the CN-weak giants have higher ratios). This
dichotomy is not seen in the somewhat more metal-poor clusters M4 and NGC 6752 (although
this cluster does have 1 giant in the sample of 12 with a high ratio and it is one of the CN-weak
stars). No measurable effect in the C-isotopic ratios as a function of CN strength is found for ω
Cen: 12C/13C = 4.9 ± 0.8 and 3.5 ± 0.5 (m.e.) for the CN-weak and CN-strong stars, respectively:
a difference of only 1.5σ.
An additional observation that should be noted about the top panel of Figure 5 is that the
field-star sample contains a number of quite metal-poor giants with 12C/13C ratios of about 8-10,
somewhat higher than most of the globular-cluster giants. This point is addressed in the bottom
panel of Figure 5, where 12C/13C is plotted versus absolute visual magnitude. The field giants with
the slightly larger C-isotopic ratios tend to be somewhat less luminous than the globular-cluster
sample. The luminosity dependence of 12C/13C is discussed by Keller et al. (2001) and the isotopic
ratios in the more luminous field giants overlap perfectly with the globular-cluster giants. Again,
the ω Cen member ROA 58 (with 12C/13C= 7.0) stands out somewhat in this plot; however, as
discussed above, this giant has a low carbon abundance and the CO bands are weak: better spectra
would need to be obtained, or more ω Cen giants observed, in order to decide whether this result
for ROA 58 is real or significant. The conclusion from this study is that for these rather luminous
ω Cen giants, with MV≤ -1.8, there is no measurable change in 12C/13C versus [Fe/H].
As discussed in Section 3, the absolute carbon abundances can also be derived fairly accurately
for these giants from the CO bands. In Figure 6, a number of isotopic and elemental abundance
ratios are plotted versus [12C/Fe] (hereafter [C/Fe]): the carbon-12 abundance is predicted to
decrease as the envelope fraction of CN-processed material increases.
In the mixing scenario a
lower [C/Fe] corresponds to deeper mixing, while in the primordial scenario it results from incresed
pollution from material processed through presumed progenitor stars. The top left panel in Figure
6 shows 12C/13C versus [C/Fe] for ω Cen and a number of other samples. Concentrating on the ω
Cen members for the moment, no trend is found, with all ω Cen giants having low 12C/13C ratios,
indicative of envelope material heavily exposed to the CN-cycle, and 12C to Fe ratios spanning
about a factor of 10. Correlations between 12C, O, and Na are explored in the bottom left and top
right panels of Figure 6. In these panels there is clear correlation between [O/Fe] versus [C/Fe] and
an anti-correlation between [Na/Fe] versus [C/Fe] (the Na and O abundances for ω Cen are from
Norris & Da Costa 1995b). As discussed in the Introduction, these are the trends expected for
material exposed to H-burning. Anticorrelations between [Na/Fe] and [O/Fe], are found in many
globular clusters, e.g., Langer, Hoffman, & Sneden (1993), Kraft et al. (1993), Pilachowski et al.
(1996), or Kraft et al. (1998). One interpretation of these Na -- O anticorrelations is deep mixing
in the currently observed giants, although Briley et al (1996) find Na-CN anticorrelations in 47
Tuc turnoff stars, while Cannon et al. (1998) find C-N anticorrelations in 47 Tuc main-sequence
-- 10 --
stars, and Gratton et al.
(2001) have found Na -- O anticorrelations in main-sequence members
of NGC6752. The ultimate origins of these various abundance correlations and anticorrelations
observed in a number of globular clusters are still not understood completely.
Other globular clusters are plotted in Figure 6 as comparisons to ω Cen. In the top left panel,
results are available for five other clusters, as well as field stars. Omega Cen does not obviously
stand out relative to the other globular clusters. In the bottom left panel of Figure 6 results for the
clusters M4 and M71 are compared to those from ω Cen; the [O/Fe] values for M4 are from Ivans
et al. (1999) and for M71 from Briley et al. (1997). Oxygen abundances are not available for the
samples we are using as comparisons for NGC6752 and 47 Tuc. All three globular clusters plotted
show a large degree of overlap in the trend of oxygen versus carbon; however, ω Cen displays a
larger range in values of both [C/Fe] and [O/Fe]. Along a similar line, the top right panel of Figure
6 shows [Na/Fe] versus [C/Fe] for the ω Cen stars, along with the M4 sample (with the [Na/Fe]
coming from Ivans et al. 1999). An anticorrelation of Na and C is found in both clusters and the
trends are, within the uncertainties, identical. Again, though, ω Cen displays a larger range in
values of [Na/Fe] relative to M4. Greater differences in [C/Fe], [O/Fe], and [Na/Fe] could indicate
either deeper mixing (in the mixing scenario) or, on the other hand, larger amounts of pollution
from material processed through H-burning (in the primordial scenario). In a recent study of M71,
Ramirez & Cohen (2002) have summarized [Na/Fe] and [O/Fe] trends for 11 globular clusters (their
Figure 12) and these different clusters display a range in their respective values of both [Na/Fe] and
[O/Fe]. If this range is quantified simply as the difference between the largest and smallest values,
a weak trend is found between a larger range in [Na/Fe] and [O/Fe] with decreasing [Fe/H]. It is
not clear how ω Cen fits into this possible trend, as these other clusters have no spread in [Fe/H],
whereas the spread in ω Cen exceeds a factor of 20; however the range in O and Na to Fe ratios
displayed by ω Cen members is as large as any found in the other globular clusters, with M13 and
M92 showing the largest ranges.
A somewhat different type of chemical behavior is probed in the bottom-right panel of Figure
6, where a slow neutron-capture (s-process) element (La) is plotted versus [C/Fe] for both ω Cen
and M4. Omega Cen is peculiar in that its member stars display a large increase, in proportion
to Fe, of s-process elements as [Fe/H] increases: such behavior is not observed in any other stellar
population. The increase in s-process abundances originates from thermally-pulsing asymptotic
giant branch (TP-AGB) stars, and the cause of their apparently large contribution to the chemical
evolution in ω Cen (Lloyd Evans 1983) remains a mystery, although speculation as to the reason
can be found in Smith et al. (2000) and Cunha et al. (2002).
In this panel no trend is found
between [La/Fe] and [C/Fe], indicating that the sources of La enrichment and C depletion are not
related. The more La-poor giants in ω Cen again overlap perfectly with their counterparts in M4.
Possible correlations of 12C/13C with abundances of three elements either destroyed (O) or
produced (Na and Al) by various proton capture cycles are explored in Figure 7. The top panel
plots the isotopic carbon ratios versus [O/Fe] for ω Cen, M4, and M71. In the regions of overlap
between these three globular clusters, there is excellent agreement between 12C/13C ratios and
-- 11 --
[O/Fe] abundance ratios. When taken together, the three cluster samples hint of a correlation
between 12C/13C and [O/Fe]. This possible trend needs to be verified by future studies of larger
samples, but it may point towards a deep-mixing signature in the current globular cluster giants.
The bottom two panels, with 12C/13C versus [Na/Fe] and [Al/Fe] show no hint of any trend. This
is somewhat puzzling in light of the possible trend of the carbon isotopic ratio with [O/Fe]. The
deeper and more extensive mixing necessary to enhance the [Na/Fe] and [Al/Fe] ratios, would be
expected to drive the envelope C-isotope ratios to their CN-equilibrium value of 3.5: this is not
observed. The lack of any correlation between 12C/13C and Na or Al points to the origins of some
of these abundance variations as arising in some other site, and not in the current cluster giants, as
concluded earlier by Briley et al. (1997). A larger sample of 12C, 13C, O, Na, and Al abundances
from globular cluster giants would be able to shed more light on this issue.
5. Conclusions
The 12C/13C ratios in all of the ω Cen giants are quite low, with a mean h12C/13Ci = 4.3±0.4:
this indicates extensive and deep mixing in these low-mass, low-metallicity giants. This result is in
general agreement with previous studies of globular cluster giants, as well as low-metallicity field
giants. In particular for the ω Cen giants, all of which have MV≤ -1.8, there is no meaurable change
in 12C/13C over the metallicity range of [Fe/H]= -1.7 to -0.7. It is found that [12C/Fe] correlates
well with [O/Fe] and anticorrelates with [Na/Fe], and that the overall trends among these various
ratios agrees very well between ω Cen and other globular clusters. The 12C/13C ratios themselves
exhibit no measurable trends with [Na/Fe] or [Al/Fe], but may hint at a positive correlation with
[O/Fe].
We wish to thank the staff of the Cerro Tololo Inter-American Observatory, especially M.
Fern´andez and H. Tirado, for their excellent assistance with the observations. Partial support
for this project came from the National Science Foundation under grants AST-9157038 and INT-
9215844 to The Ohio State University Research Foundation. VVS acknowledges support from the
National Science Foundation through grant AST-9987374.
REFERENCES
Bell, R. A., Briley, M. M., and Smith, G. H. 1990, AJ, 100, 187
Bell, R. A., Hesser, J. E., & Cannon, R. D. 1983, ApJ, 269, 580
Briley, M. M., Hesser, J. E., & Bell, R. A. 1991, ApJ, 373, 482
Briley, M. M., Bell, R. A., Hesser, J. E., & Smith, G. H. 1994, Canadian J. Phys. 72, 772
-- 12 --
Briley, M. M., Smith, V. V., Suntzeff, N. B., Lambert, D. L., Bell, R. A., & Hesser, J. E. 1996,
Nature, 383, 604
Briley, M. M., Smith, V. V., King, J., & Lambert, D. L. 1997, AJ, 113, 306
Brown, J. A., & Wallerstein, G. 1993, AJ, 106, 133
Butler, D., Dickens, R. J., & Epps, E. 1978, ApJ, 225, 148
Cannon, R. D., & Stobie, R. S. 1973, MNRAS, 162, 207
Cannon, R. D., Croke, B. F. W., Bell, R. A., Hesser, J. E., & Stathakis, R. A. 1998, MNRAS, 298,
601
Charbonnel, C. 1994, A&A, 282, 811
Charbonnel, C., Brown, J. A., & Wallerstein, G. 1998, A&A, 332, 204
Cottrell, P. L., & Da Costa, G. S. 1981, ApJ, 245, L79
Cunha, K., Smith, V. V., Suntzeff, N. B., Norris, J. E., Da Costa, G. S., & Plez, B. 2002, AJ,
submitted
DePoy, D. L., Gregory, B., Elias, J., Montan´e, A., P´erez, G., & Smith, R. M. 1990, PASP, 102,
1433
Francois, P., Spite, M., & Spite, F. A&A, 191, 267
Gonzalez, G., & Wallerstein, G. 1998, AJ, 116, 765
Goorvitch, D. 1994, ApJS, 92, 311
Gratton, R. G., Sneden, C., Carretta, E., & Bragaglia, A. 2000, A&A, 354, 169
Gratton, R. G., Bonifacio, P., Bragaglia, A., Carretta, E., Castellani, V., Centurion, M., Chieffi, A.,
Claudi, R., Clementini, G., D'Antona, F., Desidera, S., Francois, P., Grundahl, F., Lucatello,
S., Molaro, P., Pasquini, L., Sneden, C., Spite, F., Straniero, O. 2001, a, 369, 87
Gustafsson, B. E., Bell, R. A., Eriksson, K., & Nordlund, A. 1975, a, 42, 407
Hinkle, K. H., & Lambert, D. L. 1975, MNRAS, 170, 447
Holweger, H. 2001, in Solar and Galactic Composition, ed. R. F. Wimmer-Schweingruber (American
Institute of Physics: New York), p.23
Iben, Jr., I. 1964, ApJ, 140, 1631
Ivans, I. I., Sneden, C., Kraft, R. P., Suntzeff, N. B., Smith, V. V., Langer, G. E., & Fulbright, J.
P. 1999, AJ, 118, 1273
-- 13 --
Keller, L. D., Pilachowski, C. A., & Sneden, C. 2001, AJ, 122, 2554
Kleinmann, S. G., & Hall, D. N. B. 1986, ApJS, 62, 501
Kraft, R. P., Sneden, C., Langer, G. E., & Shetrone, M. D. 1993, AJ, 106, 1490
Kraft, R. P. 1994, PASP, 106, 553
Kraft, R. P., Sneden, C., Smith, G. G., Shetrone, M. D., Langer, G. E., & Pilachowski, C. A. 1997,
AJ, 113, 279
Kraft, R. P., Sneden, C., Smith, G. H., Shetrone, M. D., & Fulbright, J. 1998, AJ, 115, 1500
Kurucz, R. L., & Bell, B. 1995, CD-ROM 23 (Cambridge: SAO)
Langer, G. E., Hoffman, R., & Sneden, C. 1993, PASP, 105, 301
Lloyd Evans, T. 1983, MNRAS, 204, 975
Norris, J. E., & Da Costa, G. S. 1995a, ApJ, 441, L81
Norris, J. E., & Da Costa, G. S. 1995b, ApJ, 447, 680
Norris, J. E., Freeman, K. C., & Mighell, K. J. 1996, ApJ, 462, 241
Oliva, E., & Origlia, L. 1992, A&A, 254, 466
Paltoglou, G., & Norris, J. E., 1989, ApJ, 336, 185
Persson, S. E., Cohen, J. G., Matthews, K., Frogel, J. A., & Aaronson, M. 1980, ApJ, 235, 452
Pilachowski, C. A., Sneden, C., & Kraft, R. P. 1996, AJ, 111, 1689
Ramirez, S. V., Sellgren, K., Carr, J. S., Balachandran, S. C., Blum, R., Terndrup, D. M., & Steed,
A. 2000, ApJ, 537, 205
Ramirez, S. V., & Cohen, J. G. 2002, AJ, 123, 3277
Shetrone, M. D. 1996a, AJ, 112, 1517
Shetrone, M. D. 1996b, AJ, 112, 2639
Smith, V. V., Suntzeff, N. B., Cunha, K., Gallino, R., Busso, M., Lambert, D. L., & Straniero, O.
2000, AJ, 119, 1239
Sneden, C. 1973, Ph.D. thesis, Univ. Texas
Sneden, C., Kraft, R. P., Shetrone, M. D., Smith, G. H., Langer, G. E., & Prosser, C. F. 1997, AJ,
114, 1964
-- 14 --
Suntzeff, N. B., & Kraft, R. P. 1996, AJ, 111, 1913
Suntzeff, N. B., & Smith, V. V. 1991, ApJ, 381, 160
van Leeuwen, F., Le Poole, R. S., Reijns, R. A., Freeman, K. C., & de Zeeuw, P. T. 2000, A&A,
360, 472
Wallerstein, G., Iben, Jr., I., Parker, P., Boesgaard, A. M., Hale, G. M., Champagne, A. E., Barnes,
C. A., Kappeler, F., Smith, V. V., Hoffman, R. D., Timmes, F. X., Sneden, C., Boyd, R.
N., Meyer, B. S., Lambert, D. L. 1997, Rev.Mod.Phys., 69, 995
Wasserburg, G. J., Boothroyd, A. I., & Sackmann, I.-J. 1995, ApJ, 447, L37
Woolley, R. v.d.R. 1966, R. Obs. Ann., No. 2
This preprint was prepared with the AAS LATEX macros v5.0.
-- 15 --
Fig. 1. -- Color-magnitude diagram for proper motion members of omega Cen from van Leeuwen
et al. (2000), shown as small symbols. The photometry for the stars in the current sample are
shown as large points.
Fig. 2. -- Sample spectra of two ω Cen giants. Both of these giants have similar effective temper-
atures and metallicities ([Fe/H]). The prominent 12CO and 13CO bands are identified and most of
the structure observed as a function of wavelength is real (being composed of many CO vibration-
rotation lines). The spectra are normalized and shifted in relative flux so both can be viewed
clearly. Note that ROA 139 has weaker CO bands than ROA 155, indicating a lower total carbon
abundance in ROA 139, although both have quite low isotopic ratios (12C/13C=4.5 for ROA 155
and 3.5 for ROA 139).
Fig. 3. -- Observed (solid dots) and synthetic (continuous curves) spectra for the ω Cen giant
ROA155 (Teff = 4200K, log g= 0.8, ξ= 2.0 km s−1, and [Fe/H]= -1.64). The top panel shows
a region near the 12CO (3-1) bandhead, with synthetic spectra calculated for three different 12C
abundances (the best-fit abundance and ±0.1 dex). The abundance notation A(x)= log ǫ(x)=
log(Nx/NH + 12.0 is used. The bottom panel shows the 13CO (2-0) bandhead region with three
synthetic spectra computed with three different isotopic carbon abundance ratios.
Fig. 4. -- A comparison of carbon-12 abundances (plotted as [12C/Fe]) derived here using CO and
derived by Norris & Da Costa (1995b) using violet CH. The solid line illustrates perfect agreement.
Errorbars are taken from the discussions in Norris & Da Costa and here. The agreement between
the two studies is good, with a mean difference of (This Study -- Norris & Da Costa)= +0.14± 0.24
dex: the scatter is that expected given the respective errors in each study.
Fig. 5. -- Isotopic carbon abundance ratios are plotted as a function of [Fe/H] (top panel) and
absolute visual magnitude (bottom panel) for the different samples of metal-poor giants. There are
no strong trends with metallicity, although it should be noted that the scatter in 12C/13C in the
more metal-rich globular clusters M71 and 47 Tuc is anticorrelated with the CN strengths (Briley
et al. 1997). The field sample also contains a number of giants with somewhat higher isotopic
ratios (∼9 -- 10) at lower metallicities; however, the bottom panel reveals these field giants to be of
somewhat lower luminosities than most of the globular cluster giants analyzed. The bottom panel
reveals a modest trend of decreasing 12C/13C as the giant's luminosity increases (as discussed by
Keller et al. 2001): the globular cluster giants fall on the same trend defined by the field-giant
population.
Fig. 6. -- Comparisons of four different abundance ratios versus [12C/Fe]. The top left panel
shows 12C/13C versus [12C/Fe] for ω Cen and a number of different samples of low-metallicity
giants. No strong trends are evident, with all giants mixed extensively. The bottom left panel
shows correlations between [O/Fe] and [12C/Fe] in three globular clusters: ω Cen, M4, and M71.
The respective trends for all three clusters overlap and the correlation of depleted carbon-12 with
oxygen is a signature of material processed through the CNO cycles. The top left panel illustrates
-- 16 --
the anticorrelation between [Na/Fe] and [12C/Fe] in ω Cen and M4, again with substantial overlap
in the two clusters. The decrease in carbon-12 is due to mixing, while the increase in sodium
is due, presumably, to the Na -- Ne cycle. The bottom right panel explores the relation between
lanthanum, an s-process element, and 12C in ω Cen and M4. No trends exist, indicating that the
mixing responsible for 12C depletion has no dependance on the s-process production site.
Fig. 7. -- An investigation of possible trends of 12C/13C with O, Na, and Al. In the top panel
isotopic carbon ratios are plotted versus [O/Fe] for ω Cen, M4, and M71. The three globular
clusters together indicate a possible systematic decrease, towards the CN-equilibrium value, in
12C/13C as [O/Fe] decreases: this might signify deep mixing in the observed giants. The middle
and bottom panels show 12C/13C versus [Na/Fe] and [Al/Fe], respectively, with no measurable
trends between these abundance ratios.
10
11
12
13
14
15
V
0
.5
1
B − V
1.5
2
{ {
Table . Program Star Photometry and Metallicities
Star (ROA)
V
(B (cid:0) V )
(V (cid:0) K )
[Fe/H]
[Fe/H]
[Fe/H]
[Fe/H]
a
b
c
d
.
.
.
{.
{.
(cid:1) (cid:1) (cid:1)
{.
.
.
.
{.
{.
(cid:1) (cid:1) (cid:1)
(cid:1) (cid:1) (cid:1)
.
.
.
{.
{.
(cid:1) (cid:1) (cid:1)
(cid:1) (cid:1) (cid:1)
e
e
.
.
.
(cid:1) (cid:1) (cid:1)
{.
(cid:1) (cid:1) (cid:1)
{.
.
.
.
(cid:1) (cid:1) (cid:1)
{.
(cid:1) (cid:1) (cid:1)
(cid:1) (cid:1) (cid:1)
.
.
.
(cid:1) (cid:1) (cid:1)
{.
(cid:1) (cid:1) (cid:1)
{.
e
.
.
.
{.
{.
(cid:1) (cid:1) (cid:1)
(cid:1) (cid:1) (cid:1)
.
.
(cid:1) (cid:1) (cid:1)
{.
{.
(cid:1) (cid:1) (cid:1)
(cid:1) (cid:1) (cid:1)
.
.
.
{.
{.
{.
(cid:1) (cid:1) (cid:1)
.
.
.
{.
{.
(cid:1) (cid:1) (cid:1)
{.
f
.
.
.
{ .
{ .
(cid:1) (cid:1) (cid:1)
(cid:1) (cid:1) (cid:1)
a
Suntze(cid:11) & Kraft ( ).
Norris & Da Costa ( ).
b
c
Smith et al. ( ).
d
Other determinations of [Fe/H] as follows:
Paltoglou & Norris ( ).
e
f
Brown & Wallerstein ( ).
l
x
u
F
e
v
i
t
a
e
R
l
2
1.5
1
.5
ROA 155
ROA 139
0
23100
23200
23300
23400
23500
23600
{ {
Table . Observed wavelength intervals ((cid:22)m)
Spectrum (cid:21)(min)
(cid:21)(max)
(cid:1)(cid:21)
a
data
blue
.
.
: (cid:2)
central
.
.
(cid:1) (cid:1) (cid:1)
red
.
.
(cid:1) (cid:1) (cid:1)
data
(cid:0)
(cid:0)
blue
.
.
: (cid:2)
red
.
.
(cid:1) (cid:1) (cid:1)
a
Derived for this spectrum interval and assumed
to apply to the others.
l
x
u
F
e
v
i
t
l
a
e
R
l
x
u
F
e
v
i
t
l
a
e
R
ROA 155
1
.8
.6
23200
23220
23240
23260
23280
23300
ROA 155
1.1
1
.9
.8
.7
23400
23420
23440
23460
23480
23500
{ {
Table . Program Star Parameters & Isotopic Carbon Abundances
Star (ROA) T
(K)
Log g (cm s
)
(cid:24) (km s
) A(
C)
[
C/Fe]
C/
C
e(cid:11)
(cid:0)
(cid:0)
a
b
.
.
.
- .
.
.
.
.
- .
.
.
.
.
-.
.
.
.
.
- .
.
.
.
.
- .
.
.
.
.
- .
.
.
.
.
- .
.
.
.
.
- .
.
.
.
.
- .
.
.
.
.
+ .
.
.
.
.
- .
.
a
A(
C)= log (cid:15)(
C)= log[N(
C)/N(H)] + .
b
This ratio is computed using solar abundances of A(
C)= . and A(Fe)= . , with
[Fe/H] for the star taken from Norris & Da Costa ( b)
.5
0
-.5
-1
-1.5
-1.5
-1
-.5
0
.5
M4
M71
47 Tuc
NGC6752
Field
15
10
5
0
-2.5
-2
-1.5
[Fe/H]
-1
-.5
15
10
5
0
1
0
-1
-2
-3
-4
M4
M71
47 Tuc
NGC6752
Field
15
10
5
.5
0
-.5
]
e
F
O
/
[
1
.5
0
1.5
1
.5
0
]
/
e
F
a
N
[
]
e
F
a
L
/
[
-1.5
-1
-.5
0
.5
-1.5
-1
-.5
0
.5
[C/Fe]
[C/Fe]
12
10
8
6
4
2
0
-.8
M4
M71
-.6
-.4
-.2
0
.2
.4
.6
.8
[O/Fe]
12
10
8
6
4
2
0
12
10
8
6
4
2
0
1
.8
.6
.4
.2
0
-.2
-.4
[Na/Fe]
1.4
1.2
1
.8
.6
.4
.2
0
-.2
[Al/Fe]
|
astro-ph/9503089 | 1 | 9503 | 1995-03-24T19:52:59 | High Redshift Lyman Limit and Damped Lyman-Alpha Absorbers | [
"astro-ph"
] | We have obtained high signal:to:noise optical spectroscopy at 5\AA\ resolution of 27 quasars from the APM z$>$4 quasar survey. The spectra have been analyzed to create new samples of high redshift Lyman-limit and damped Lyman-$\alpha$ absorbers. These data have been combined with published data sets in a study of the redshift evolution and the column density distribution function for absorbers with $\log$N(HI)$\ge17.5$, over the redshift range 0.01 $<$ z $<$ 5. The main results are: \begin{itemize} \item Lyman limit systems: The data are well fit by a power law $N(z) = N_0(1 + z)^{\gamma}$ for the number density per unit redshift. For the first time intrinsic evolution is detected in the product of the absorption cross-section and comoving spatial number density for an $\Omega = 1$ Universe. We find $\gamma = 1.55$ ($\gamma = 0.5$ for no evolution) and $N_0 = 0.27$ with $>$99.7\% confidence limits for $\gamma$ of 0.82 \& 2.37. \item Damped \lya systems: The APM QSOs provide a substantial increase in the redshift path available for damped surveys for $z>3$. Eleven candidate and three confirmed damped Ly$\alpha$ absorption systems, have been identified in the APM QSO spectra covering the redshift range $2.8\le z \le 4.4$ (11 with $z>3.5$). Combining the APM survey confirmed and candidate damped \lya absorbers with previous surveys, we find evidence for a turnover at z$\sim$3 or a flattening at z$\sim$2 in the cosmological mass density of neutral gas, $\Omega_g$. \end{itemize} The Lyman limit survey results are published in Storrie-Lombardi, et~al., 1994, ApJ, 427, L13. Here we describe the results for the DLA population of absorbers. | astro-ph | astro-ph | High Redshift Lyman Limit and
Damped Lyman-Alpha Absorbers
L.J. Storrie-Lombardi1,2, R.G. McMahon1,5, M.J. Irwin3, C. Hazard1,4
1 Institute of Astronomy, Madingley Road, Cambridge CB3 0HA UK
2 current address: UCSD-CASS, Mail Code 0111, La Jolla, CA 92093 USA [email protected]
3Royal Greenwich Observatory, Madingley Road, Cambridge CB3 0HE UK [email protected]
4University of Pittsburgh, Pittsburgh, PA USA
5 [email protected]
5
9
9
1
r
a
M
4
2
1
v
9
8
0
3
0
5
9
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
To appear in "ESO Workshop on QSO Absorption Lines"
Preprint: astro-ph/yymmddd
Abstract
We have obtained high signal:to:noise optical spectroscopy at 5A resolution of 27
quasars from the APM z>4 quasar survey. The spectra have been analyzed to create
new samples of high redshift Lyman-limit and damped Lyman-α absorbers. These
data have been combined with published data sets in a study of the redshift evolution
and the column density distribution function for absorbers with logN(HI)≥ 17.5, over
the redshift range 0.01 < z < 5. The main results are:
• Lyman limit systems: The data are well fit by a power law N (z) = N0(1 + z)γ
for the number density per unit redshift. For the first time intrinsic evolution
is detected in the product of the absorption cross-section and comoving spatial
number density for an Ω = 1 Universe. We find γ = 1.55 (γ = 0.5 for no evolution)
and N0 = 0.27 with >99.7% confidence limits for γ of 0.82 & 2.37.
• Damped Lyα systems: The APM QSOs provide a substantial increase in the
redshift path available for damped surveys for z > 3. Eleven candidate and three
confirmed damped Lyα absorption systems, have been identified in the APM QSO
spectra covering the redshift range 2.8 ≤ z ≤ 4.4 (11 with z > 3.5). Combining
the APM survey confirmed and candidate damped Lyα absorbers with previous
surveys, we find evidence for a turnover at z∼3 or a flattening at z∼2 in the
cosmological mass density of neutral gas, Ωg.
The Lyman limit survey results are published in Storrie-Lombardi, et al., 1994, ApJ,
427, L13. Here we describe the results for the DLA population of absorbers.
1
Introduction
How and when galaxies formed are questions at the forefront of work in observational cos-
mology. Absorption systems detected in quasar spectra provide the means to study these
phenomena up to z∼5, back to when the Universe was less than 10% of its present age. While
the baryonic content of spiral galaxies that are observed in the present epoch is concentrated
in stars, in the past this must have been in the form of gas. Damped Lyα absorption (DLA)
systems have neutral hydrogen column densities of N(HI)> 2 × 1020cm−2. They dominate
the baryonic mass contributed by HI. The principal gaseous component in spirals is HI which
1
has led to surveys for absorption systems detected by the DLA they produce (Wolfe, Turn-
shek, Smith & Cohen 1986 [WTSC]; Lanzetta et al. 1991 [LWTLMH]; Lanzetta, Wolfe &
Turnshek 1995 [LWT95]). We extend the earlier work on Lyman limit systems and DLAs to
higher redshifts using observations of QSOs from the APM z>4 QSO survey (Irwin, McMa-
hon & Hazard 1991), These data more than triple the redshift path surveyed at z>3 and
allow the first systematic study up to z=4.5.
2 APM Damped Lyman-Alpha Survey at z∼4
We have obtained 27 high S/N spectra at 5A resolution at the William Herschel Telescope.
The spectra were analyzed starting 3000 km s−1 blueward of z(emission). The analysis
was stopped when the S/N ratio became too low to detect a Lyα line with W(rest)≥5A.
This point was typically caused by the incidence of a Lyman limit system. Features with
∼ 25A were selected with an automated procedure. Most of these are blends
W(observed) >
of the dense Lyα forest features present at high redshift. The equivalent width and FWHM
were measured interactively and N(HI) was estimated for features with W or FWHM >30A.
Of the 34 measured, 15 have estimated N(HI)≥ 2 × 1020 cm−2 covering 2.8 ≤ z ≤ 4.4.
Only one candidate has estimated N(HI)≥ 1021 cm−2. High resolution spectroscopy of 4
candidates has confirmed 3 as damped (log N(HI)≥20.3). The sensitivity of the survey with
redshift was determined using the method developed by Lanzetta (see LWT95). The function
g(z) is calculated, giving the number of lines of sight along which a damped system at a
redshift z could be detected. Figure 1a shows the sensitivity of the APM survey alone and
in combination with the WTSC and LWTLMH surveys.
3 Evolution of the Number Density per Unit Redshift
The candidate and confirmed DLA systems from the APM sample and previous surveys
(WTSC; LWTLMH; LWT95) have been combined to study the evolution of the number
density per unit redshift for 0.01 < z < 4.7. Fit with the customary power law N(z) =
N0(1 + z)γ, a population with no intrinsic evolution in the product of the absorption cross-
section and comoving spatial number density will have γ = 1/2 (Ω = 1) or γ = 1 (Ω = 0).
A maximum likelihood fit to the data with z>1.5 yields N(z) = 0.03(1 + z)1.5±0.6, consistent
with no intrinsic evolution even though the value of γ is similar to that found for the Lyman
limit systems where evolution is detected at a significant level. However, there is redshift
evolution evident in the higher column density systems with an apparent decline in N(z)
for z>3.5. These results are displayed in figure 1(b). The combined data set is plotted as
dashed lines with the above fit. The results for only the absorbers with log N(HI)≥ 21 are
shown as solid lines. The z<1.5 bin is taken from LWT95.
2
4 Evolution of Ωg -- Baryons in Neutral Gas
The mean cosmological mass density contributed by damped Lyα absorbers can be estimated
as
Nf (N)dN
Nmin
hΩgi =
H0µmH
cρcrit Z ∞
as defined in LWTLMH (equations 17-18), giving the current mass density in units of the
current critical density. The errors in Ωg are difficult to estimate because the column density
distribution function, f (N), is not known. LWTLMH utilised the standard error in the
distribution of N(HI) which yields zero error if all the column densities in a bin are the
same. We have estimated the fractional variance in Ωg as Pp
i=1 Ni)2 which yields
√n errors if all the column densities included in a bin are equal. This method yields larger
errors. The results for Ωg are shown in figure 2 for q0=0 and q0=0.5 (H0=50). The z<1.5
bin is taken from LWT95. The z>1.5 solid bins utilise the data from WTSC, LWTLMH,
and the APM survey. The dotted bins exclude the APM data. The inclusion of the APM
survey data for z > 3 lowers the value previously found for 3 < z < 3.5 and indicates a
possible turnover for z > 3.5. The results are also consistent with a relatively constant value
of Ωg for z > 2 as the error bars are still very large at high redshift. Larger samples of bright
z > 4 quasars are needed.
i /(Pp
i=1 N 2
5 Summary
The QSOs from the APM survey more than triple the z > 3 redshift path for DLA surveys.
Fourteen candidate DLA systems have been identified in the APM spectra covering 2.8 ≤
z ≤ 4.4 (11 with z > 3.5), with 3 confirmed. Combining these data with the previous surveys
and fitting a single power law for z>1.5 gives N(z)= .03(1 + z)1.5±0.6, marginally consistent
with no evolution models. Evolution is evident in the highest column density absorbers
∼ 3.5. We find
with the incidence of systems with log N(HI)≥21 apparently decreasing for z >
evidence for a turnover or flattening in the cosmological mass density of neutral gas, Ωg
at high redshift. The more gradual evolution of Ωg than previously found helps alleviate
the 'cosmic G-dwarf problem' (LWT95), i.e.
if a large amount of star formation has taken
place between z=3.5 and z=2, a much larger percentage of low metallicity stars should exist
than is detected. It is also consistent with the suggestion by Pettini et al. (1994) that the
wide range in DLA metallicities measured at the same epoch indicates that at z∼2 they are
observed prior to the bulk of star formation in the disk.
3
References
[1] Gnedin, N.Y. & Ostriker, J.P., 1992, ApJ, 400,1
[2] Irwin, M.J., McMahon, R.G. & Hazard, C., 1991, in ASP Conf. Series, Vol. 21, ed.
Crampton D., (San Francisco: ASP), 117
[3] Lanzetta, K.M., Wolfe, A.M., Turnshek, D.A., Lu, L., McMahon, R.G. & Hazard, C.,
1991, ApJS, 77, 1 (LTWLMH)
[4] Lanzetta, K.M., Wolfe, A.M. & Turnshek, D.A., 1995, ApJ, 440, 435 (LWT95)
[5] Pettini, M., Smith, L.J., Hunstead, R.W. & King, D.L., 1994, ApJ, 426, 79
[6] Rao, S. & Briggs, F., 1993, ApJ, 419, 515
[7] Storrie-Lombardi, L.J., McMahon, R.G., Irwin, M.J. & Hazard 1994, ApJ, 427, L13
[8] Wolfe, A.M., Turnshek, D.A., Smith, H.E. & Cohen R.D., 1986, ApJS, 61, 249 (WTSC)
Figure Captions
Figure 1: (a) The sensitivity function, g(z), of the DLA surveys. This gives the number of
lines of sight along which a damped system at redshift z could be detected. (b) The number
density of DLA per unit redshift, N(z), vs. z(absorption). The dashed bins show N(z) for
all the damped systems and the solid bins for systems with N(HI)≥ 1021 cm−2. A single
power law fit to the sample for z>1.5 gives N(z)= .03(1 + z)1.5±0.6.
Figure 2: The mean cosmological mass density in neutral gas, Ωg, contributed by DLA
absorbers for 0.01≤z≤4.7 for q0=0 and q0=0.5 (H0=50). The z<1.5 bin is taken from
LWT95. The z>1.5 solid bins utilise the WTSC, LWTLMH, and APM survey data. The
dotted bins exclude the APM data. The points at z=0 are Ωstars (star) from Gnedin &
Ostriker (1992) and ΩHI (circle) from Rao & Briggs (1993).
4
|
astro-ph/9609023 | 1 | 9609 | 1996-09-03T22:11:24 | An Overview of Blazar Variability | [
"astro-ph"
] | Blazars are characterized by rapid variability at virtually all wavelengths from radio through TeV gamma-rays. The challenge since their discovery has been to understand the origin of their luminous, apparently nonthermal, nuclear emission. Considerable progress has been made in recent years thanks to a handful of multiwavelength monitoring campaigns with high enough temporal sampling to resolve the most rapid variations. The best data for a few objects have shown a variety of behaviors, for the most part commensurate with synchrotron and Compton-scattered emission from a relativistic jet, though better data for more blazars are still clearly needed. In particular, the origin of the seed photons that are upscattered to gamma-ray energies remains unclear. The latest multiwavelength light curves for the BL Lac object PKS2155-304 appear to rule out synchrotron emission from a homogeneous source. | astro-ph | astro-ph |
An Overview of Blazar Variability
C. Megan Urry
Space Telescope Science Institute, 3700 San Martin Drive, Baltimore,
MD 21218
Abstract. Blazars are characterized by rapid variability at virtually all
wavelengths from radio through TeV gamma-rays. The challenge since
their discovery has been to understand the origin of their luminous, ap-
parently nonthermal, nuclear emission. Considerable progress has been
made in recent years thanks to a handful of multiwavelength monitor-
ing campaigns with high enough temporal sampling to resolve the most
rapid variations. The best data for a few ob jects have shown a vari-
ety of behaviors, for the most part commensurate with synchrotron and
Compton-scattered emission from a relativistic jet, though better data for
more blazars are still clearly needed. In particular, the origin of the seed
photons that are upscattered to gamma-ray energies remains unclear. The
latest multiwavelength light curves for the BL Lac ob ject PKS 2155–304
appear to rule out synchrotron emission from a homogeneous source.
1.
Introduction
Soon after blazars were discovered and identified, they were selected as inter-
esting targets for long-term monitoring programs, by pioneers like Alex Smith
among others, because of their extreme characteristics: they were the most vari-
able, the most luminous, the most polarized, and in some sense the most exciting
type of Active Galactic Nuclei (AGN). Understanding them at first seemed the
key to understanding the AGN phenomenon. Ironically, blazars were eventually
perceived as less interesting for the very same reason — because they were so
unusual, so the reasoning went, they must not be relevant to the greater body of
AGN (i.e., Seyfert galaxies and quasars). Blazar research was seen as a special
field and blazars as arcane oddities.
Now in the mid-90s, we have come full circle. Because blazars are rare ge-
ometric manifestations of a general phenomenon (assuming they are relativistic
jets pointing directly at us; Urry & Padovani 1995), they must be a quite com-
mon kind of AGN, and when they are not pointing at us, we simply call them
radio galaxies. This makes blazars highly relevant to understanding AGN as a
whole. Specifically, the enormous energy of a relativistic jet, its emanation from
the vicinity of the putative central black hole, and its high degree of collimation
over many orders of magnitude in scale, offer direct clues to the extraction of
energy from the black hole. Blazars thus reveal the energetic processes occurring
in the very centers of active galaxies, while in the more common radio galaxies,
jet radiation (and hence information) is beamed away from us. With blazars,
1
then, the goal is to understand black hole physics through understanding the
physics of the jet. This in turn can be deduced from multiwavelength spectral
characteristics, most notably correlated variability across the spectrum.
2. Blazars as Relativistic Jets
Here I summarize the arguments for believing blazars are relativistic jets. First,
they commonly exhibit superluminal motion (Vermeulen & Cohen 1994; see also
Wehrle et al., these Proceedings) which, while it could arise from pattern rather
than bulk relativistic velocity, is at least suggestive. Second, blazars can ex-
hibit extremely high brightness temperatures; in at least some cases, intraday
variations are observed at optical wavelengths, ruling out an extrinsic (scintilla-
tion) explanation for the variability, although the implied bulk Lorentz factors
are uncomfortably high (Wagner & Witzel 1995). Third, the characteristically
high and variable polarization of blazars is explained naturally by an aligned jet
(Smith et al., these Proceedings). Fourth, multiwavelength radio variability is
well explained by shocks in a jet (Aller, these Proceedings).
Fifth and most compelling, the strong and variable gamma-ray emission
observed in many blazars (Hartman, these Proceedings) implies such a high
compactness that the gamma-ray source would inevitably be dominated by pair
production unless the emission is relativistically beamed (Dondi & Ghisellini
1995). While there is some uncertainty about the degree of beaming required
(it depends on the ambient X-ray photon density, as X-gamma interactions are
the most likely pair-production mechanism), the argument for some beaming is
fairly tight.
Blazars can be defined in various ways, via their rapid variability, their
compact flat-spectrum radio emission, their superluminal motion, their polar-
ization, and now their gamma-ray brightnesses, and in fact these characteristics
occur in the same sources. That is, those sources that are superluminal have
flat radio spectra and are highly polarized, and so on.1 An even more direct
link has been seen in at least two cases. Ten years of VLBI maps of 3C 279
(Wehrle et al., these Proceedings) show a new VLBI component being “born”
(extrapolating the observed position backward with the observed velocity) at
the time of the bright gamma-ray flare in June 1991 (Wehrle et al. 1994). The
same phenomenon has been observed in PKS 0528+134 (Pohl et al. 1995). In
these two blazars, the production of (beamed) gamma-rays is directly related to
superluminal motion of the radio source. That these various blazar character-
istics are all closely linked, statistically and in some cases directly, is a strong
argument that the underlying cause is relativistic beaming.
1An exception is that some highly polarized quasars (HPQ), largely radio-quiet, have continuum
emission polarized by scattering rather than intrinsic processes like synchrotron radiation.
These obviously do not have blazar characteristics like superluminal motion or rapid variability.
2
FSRQ, LBL
HBL
u
f
u
g
o
L
Radio
IR
Opt UV EUV X-Ray
GRO
10
15
Log Frequency (Hz)
20
Figure 1.
Schematic broad-band spectra of blazars from radio
through TeV gamma rays. The low-energy component is probably due
to synchrotron radiation and the high-energy component to Compton
scattering of lower-energy seed photons, possibly the synchrotron pho-
tons or ambient UV/X-ray disk or line photons. Two different curves
represent the average spectral shapes (Sambruna et al. 1996) of HBL
(High-frequency peaked BL Lac ob jects; dotted line) and LBL (Low-
frequency peaked BL Lac ob jects; dashed line) as defined by their ratios
of X-ray to radio flux (see footnote 2). Strong emission-line blazars
(i.e., flat-spectrum radio quasars, or FSRQ) have continua like LBL
(Sambruna et al. 1996).
3. Multiwavelength Spectra and Monitoring
Observations in individual wavebands have established the viability of the rel-
ativistic beaming hypothesis for blazars but have led to at best cursory under-
standing of the physical state of the jet. For this, multiwavelength variability
holds the key. Blazar spectra span an extremely broad range of energies, from
radio through GeV gamma-rays and perhaps through TeV gamma-rays. The
emission consists of two distinct spectral components, a low-energy synchrotron
bump and a high-energy Compton-scattered bump (Figure 1). The temporal
evolution of each spectral component and the correlation between them are
critical to understanding the underlying emission mechanisms, which is why a
number of large multiwavelength monitoring campaigns have been carried out
in recent years.
It should be noted that the difficulty of arranging these multiwavelength mo-
nitoring observations, typically coordinated among several satellites and many
more ground-based telescopes, has kept us from obtaining good data on more
3
than a few of the brightest blazars. More importantly, sensitivity limits have
introduced significant target selection effects. Specifically, among BL Lac ob-
jects, “High-frequency peaked BL Lacs” (HBL) and “Low-frequency peaked BL
Lacs” (LBL) have distinctly different continuum shapes.2 These may indeed be
opposite extremes of a continuous distribution because current BL Lac samples
come from radio or X-ray surveys with fairly high flux limits and so are natu-
rally dominated by LBL or HBL, respectively. The strong emission-line blazars
(i.e., flat-radio-spectrum quasars, or FSRQ), also generally radio-selected, have
continua like LBL (Sambruna et al. 1996).
In any case, the spectral energy distributions of HBL and LBL/FSRQ differ
in several ways (Fig. 1). The peak wavelength of the synchrotron component
is in the infrared-optical band for LBL and FSRQ, whereas it peaks in the
extreme ultraviolet to soft X-ray range for HBL. Also, LBL and FSRQ have
a higher ratio of gamma-ray to synchrotron flux than the HBL (Sambruna et
al. 1996), and most of the EGRET blazars are in fact FSRQ and LBL. Note
that the synchrotron emission is most variable above the peak in ν Fν , where
the shortest wavelength component becomes optically thin (Ulrich, Maraschi, &
Urry 1996).
Only two blazars, Mrk 421 and Mrk 501, both HBL, have been detected
at TeV energies (Punch et al. 1992, Quinn et al. 1996). It may be that these
particular ob jects were detected because they are relatively nearby (z ∼ 0.03
in both cases) so that the ultra-high-energy gamma-rays have little path length
along which to produce pairs via scattering of intergalactic microwave photons,
but it is probably also significant that the peak Compton emission in HBL is
likely at much higher energies than in LBL/FSRQ. The luminosity represented
by the extension of the blazar spectrum to GeV/TeV energies is phenomenal;
clearly, understanding the production of this emission is central to understanding
the blazar.
Because of these systematic spectral differences, the observational details of
the multiwavelength study dictate the type of blazar studied. To study blazars
above the synchrotron peak, where they are most variable, means selecting UV-
and X-ray-bright targets, which are inevitably HBL. To study correlated intra-
day variability at radio and optical wavelengths means selecting LBL/FSRQ. To
correlate with GeV gamma-rays, one looks primarily at LBL/FSRQ; to corre-
late with TeV gamma-rays, one looks instead at HBL. With higher sensitivities,
this artificial distinction will disappear, but it is important to remember that
as presently observed, the radio-optical intraday variables are systematically
different ob jects than the highly variable UV/X-ray-bright sources.
With multiwavelength monitoring of blazars, there are two critical questions
we are in the process of addressing. First, where do the gamma-rays come from,
and second, what is the structure of the jet itself ? By figuring out what the
particle density is, what the magnetic fields are, how each varies along the jet,
and what causes flaring behavior, we can ultimately understand what created
2The two sub-classes of BL Lac ob ject are defined by their ratio of X-ray to radio flux, which anti-
correlates with the wavelength of their peak synchrotron emission. A High-frequency peaked
BL Lac (HBL) has αrx (between 5 GHz and 1 keV) less than 0.75, while a Low-frequency
peaked BL Lac (LBL) has αrx > 0.75 (Padovani & Giommi 1995).
4
Figure 2. Multi-epoch broad-band spectra of 3C 279, in a high state
in June 1991 and in a low state in January 1994 (Maraschi et al. 1994).
While the UV decreased by a factor of 3-4, the gamma rays decreased
by a factor of ∼ 10 or more. As is typical for blazars, there is little or no
variability below the peak wavelength of the synchrotron component,
the peak of which is in the unobserved far-infrared (now accessible with
ISO).
the jet, how it was formed, and what is happening down at the center where we
cannot observe directly. For the rest of this paper, I discuss only two ob jects,
3C 279 and PKS 2155–304, which illustrate some of the best available data (see
also Takahashi et al., these Proceedings) and thus the limits of what we can
learn about blazars at this point.
4. Multi-Epoch Flaring in the Superluminal Quasar 3C 279
3C 279 is the brightest gamma-ray blazar in the sky.
[During the week of the
Miami blazar meeting, it was undergoing a ma jor outburst, rising an order of
magnitude above its previous highest gamma-ray state, with substantial vari-
ations at other wavelengths.] Figure 2 shows the broad-band spectra at two
epochs, the high state in June 1991 when it was first discovered and a low
state in January 1994; the greatest change in intensity occurs at gamma-ray
wavelengths, with lesser but still substantial variations at UV and X-ray wave-
lengths, and little change at radio to sub-millimeter wavelengths (Maraschi et al.
1994). Note that the gamma-ray flux can dominate the bolometric luminosity,
particularly in the flare state. Maraschi et al. (1994) suggested that the larger
5
increase in gamma-rays relative to synchrotron emission was consistent with the
synchrotron self-Compton (SSC) model, although a change in Doppler factor of
the jet or concurrent increases in seed photons and scattering electrons could
also cause the observed variability. Clearly, sampling these flares more finely
is critical to identifying the origin of the seed photons and thus the produc-
tion of the dominant spectral component. The multiwavelength observations of
the 1996 flare should help resolve the ambiguity about the identity of the seed
photons (Wehrle et al., in preparation).
The basic blazar models currently under consideration have one common
element, a synchrotron-emitting jet filled with energetic electrons and perhaps
positrons. In addition, there may be ambient UV and X-ray emission from the
vicinity of an accretion disk, and broad-line emission from clouds farther out,
photoionized either by the central UV/X-ray source or by the jet itself. The
debate turns on which photons are Compton-scattered to gamma-ray energies.
Photons impinging on the jet from the side are boosted in the frame of the jet
electrons, and so constitute a very intense flux of seed photons even when the
directly observed non-jet UV/X-ray flux is low.
Ghisellini and Madau (1996) present a nice comparison of the principal
ideas (see references therein as well) and conclude two interesting points. One is
that the inner part of the jet has to be dissipationless. If the energy density of
gamma-rays in an optically thick inner region of the jet were high, there would
probably be enough local UV photons to generate a pair cascade, transferring
much of the gamma-ray energy into X-rays, contrary to what is observed. So
the principal mode by which energy is transferred from the black hole to the
jet must not be via energetic photons. The second point is that, for plausible
numbers, the illumination of the broad-line clouds by the beamed continuum can
be an important contribution to the UV flux impinging on the jet. The bulk of
the gamma-rays could therefore come from scattered broad-line photons which
were photoionized by the jet itself. This might be the dominant mechanism in
FSRQ, say, while SSC emission dominates in the weaker-lined HBL.
It would be extremely interesting to monitor simultaneously the variability
of the broad lines and the synchrotron continuum. As far as I am aware this
has not yet been done, at least with sufficient sampling. We have looked at the
archival UV data for 3C 279 (Koratkar et al. 1996) but even there, with a very
well-observed source, the data are not sufficient to determine the photoionization
source unambiguously. The 1996 campaign on 3C 279 will help and there are
a number of other blazar campaigns planned for the upcoming year which may
contribute to solving this problem. Unfortunately, the loss of IUE, with its long,
sustained monitoring campaigns, is a ma jor blow for this kind of study.
5. Multiwavelength Variability of the HBL PKS 2155–304
PKS 2155–304 is the brightest BL Lac ob ject at ultraviolet wavelengths and one
of the brightest in the X-ray as well. So it is the obvious choice for UV/X-ray
monitoring. It is an HBL, with peak synchrotron emission near 1017 Hz, which is
four orders of magnitude higher in frequency than the peak synchrotron emission
in 3C 279. We have every reason to expect, therefore, that the physics of the
emission from PKS 2155–304 and 3C 279 differ in significant ways.
6
Figure 3.
Normalized X-ray, UV, and optical light curves of PKS
2155–304 from the first intensive multiwavelength monitoring cam-
paign, in mid-November 1991 (Edelson et al. 1995). The emission
is well-correlated, with comparable amplitude independent of wave-
length, and the X-rays appear to lead the UV by ∼ 2-3 hours. Even
with extensive coverage from the ground, the best optical data were
obtained with the FES monitor on IUE, a simple star-tracking device,
which highlights the value of adding modest optical devices to high-
energy satellites.
There have been two intensive multiwavelength campaigns to observe PKS
2155–304.
[There was a third three months after the Miami blazar meeting.]
The first was in November 1991 and lasted for one month. At that time, no
one even knew whether the UV and X-ray emission were related, nor what the
fastest time scale for variability was. We observed PKS 2155–304 with IUE once
per day throughout November 1991 (thinking that was probably overkill) and
then in the middle of the month, at the insistence of Rick Edelson (who rightly
realized it was not overkill), we observed it for nearly 5 days continuously. The
UV variations we detected were indeed fast enough that the daily sampling was
insufficient and only the continuous observations were useful for multiwavelength
cross-correlations (Urry et al. 1993). We also had ∼ 3.5 days of continuous X-
ray observations (Brinkmann et al. 1995) overlapping with most of the intensive
UV coverage, as well as considerable ground-based radio, infrared, and optical
observations (Smith et al. 1992, Courvoisier et al. 1995).
Results from the November 1991 campaign are shown in Figure 3. The
X-ray, UV, and optical light curves are well-correlated, arguing for a common
origin of the optical through X-ray emission, and the X-rays lead the UV by ∼ 2-
3 hours (Edelson et al. 1995). The optical/UV emission can not be produced
by thermal emission from a viscous accretion disk because they should arise at
7
Figure 4.
Normalized X-ray, EUV, and UV light curves of PKS 2155–
304 from the second intensive multiwavelength monitoring campaign,
in May 1994 (Urry et al. 1996). The ASCA data show a strong flare,
echoed one day later by EUVE and two days later by IUE. The ampli-
tude of the flares decreases and the duration increases with increasing
wavelength.
very different radii, implying a large lag between the two. The amplitude of
variation is independent of wavelength, a result contrary to what is expected
from a synchrotron flare caused by an increase in energetic electrons (Celotti,
Maraschi, & Treves 1991).
Several aspects of the 1991 results were intriguing enough to inspire us to
repeat the experiment for a longer period and with more extensive wavelength
coverage. First, the light curves consisted of a series of peaks modulated by an
overall decline in flux; while not strictly periodic (Edelson et al. 1995), they
suggested possible repetition through five cycles. Second, the UV and X-ray
light curves, which tracked very well for most of the observation, appeared to
diverge at the very end of the overlapping period.
We therefore arranged a second intensive monitoring campaign in May 1994,
with IUE for 10 days continuously (Pian et al. 1996), EUVE for 9 days (Marshall
et al., in prep.), and ASCA for 2 days (Kii et al., in prep.), and with additional
Rosat (Urry et al. 1996) and ground-based data (Pesce et al. 1996). Fig-
ure 4 shows the results from this second set of multiwavelength observations of
PKS 2155–304 (Urry et al. 1996). A very sharp flare is seen in X-rays, followed
by an EUV flare one day later and a broader, lower amplitude UV flare two days
later. The X-ray flux doubles in ∼ 0.2 days and is approximately symmetric.
The Rosat data, while sparse, show that large X-ray flares were not unusual but
were occurring throughout the week prior to the ASCA observations. The EUV
flux increased by ∼ 50% in less than a day (the time scale is difficult to estimate
8
Figure 5.
Cross correlation of May 1994 ASCA and IUE light curves
of PKS 2155–304, showing that the X-ray flux leads the UV by ∼ 2 days
(Urry et al. 1996).
given the errors and the untimely end of the data train), and the UV flux rose
by ∼ 35% with a doubling time scale of ∼ 3 days. The duration of the flare
increases from less than a day in the X-rays to nearly 4 days in the UV.
The delays between X-ray, EUV, and UV light curves are easily measured
with cross-correlation functions. Formally, the EUV flux leads the UV by 1.1
days and the X-ray flux leads the UV by 2.0 days. The cross-correlation between
IUE and ASCA light curves is shown in Figure 5; the cross-correlation of X-ray
versus EUV light curves is not well defined due to the small temporal overlap.
(Fig. 5).
The IUE data also reveal a complex and extremely rapid flare at the be-
ginning of the observation, with doubling times as fast as 1 hour, the fastest
ever observed at ultraviolet wavelengths and comparable to the fastest doubling
times seen in the X-ray (Pian et al. 1996). The event is seen in both LWP
and SWP cameras, and is undersampled by both; it has larger amplitude in the
LWP in part because the LWP integration times are less than half the SWP
integration times. It is also possible to see similar structure in the EUV light
curves, although the EUVE data are relatively noisy, ∼ 1-2 days in advance of
the IUE event. Figure 6 shows an expanded view of the EUVE and IUE light
curves, with the EUVE curve shifted forward by 1.25 days.
There are strong differences between the May 1994 and November 1991
light curves of PKS 2155–304. In the second epoch, the amplitude and duration
depend strongly on wavelength and the lags are considerably longer than in the
first. In addition, there is little of the low-amplitude repetitive variability, at
least in the first half of the IUE observation. These differences mean either
that there are two different mechanisms operating or that the relevant physical
9
Figure 6.
Expanded view of the beginning of the IUE light curve
of PKS 2155–304 from May 1994, with the EUVE light curve shifted
forward by 1.25 days. The extremely rapid UV variability on May 16,
undersampled in both LWP and SWP cameras, is similar to earlier,
lower amplitude variations in the EUVE light curve.
parameters have changed considerably. The variability in 1991 was probably
not a synchrotron flare, since the observed time scales were energy independent.
Instead, it is possible that it was caused by microlensing: for a dense star cluster
at the redshift of a known Lyman-alpha absorption system (about half way to the
BL Lac), and assuming relativistic motion of the BL Lac jet, the amplitudes and
time scales are approximately correct, and the achromatic nature is automatic
as long as the source size is independent of wavelength. This is a plausible but
not proven explanation.
The 1994 data are much richer in extent and wavelength coverage. These
data are consistent with a synchrotron flare in a jet but the clear delay between
X-ray and UV flares rules out the homogeneous case (Urry et al. 1996). The
reasoning is as follows. The simplest causes of a flare in the homogeneous case
would be a sudden uniform increase in the injection of energetic electrons or
the instantaneous and uniform enhancement of the magnetic field, perhaps via
compression of a charged plasma. In both cases, the flux at all (optically thin)
wavelengths would rise simultaneously, with amplitude increasing with decreas-
ing wavelength, while the duration of the flare (if due to energy losses only)
would go as λ1/2 . That is, the long-wavelength emission would last longer than
the short-wavelength emission but the (instantaneous) peaks in the light curves
would be simultaneous. The delay in the flare onset and the longer rise time for
the UV flare compared to the X-ray flare clearly rule out the simplest homoge-
neous case, unless ad hoc dependences on time and energy of the injection rate
are postulated.
10
Thus, some energy stratification of the synchrotron-emitting plasma is re-
quired, either behind a shock or in the jet structure itself. The X-ray decay
is faster than the UV but perhaps by less than a factor (λU V /λX )1/2
∼ 20,
suggesting the decay is dominated by geometry rather than synchrotron losses.
The flare duration increases with wavelength, possibly indicating that the size
of the emitting region is increasing with wavelength. The observed lags are also
comparable to the flare durations. Both these results are as expected for a shock
propogating outward in an inhomogeneous jet, successively passing from X-ray-
to EUV- to UV-emitting regions.
6. Future Multiwavelength Monitoring
As this review has shown, the best available data are still insufficient for deter-
mining jet structure but enough to indicate the kind of data needed. We should
repeat the kind of multiwavelength monitoring done for PKS 2155–304 for many
more blazars, including LBL, which will require a new ultraviolet capability. We
also need to study further the correlation of gamma-ray and optical/IR variabil-
ity. In all cases, long and intensive time sampling is critical: two light curves that
are well correlated could, if sampled at few points or for less time than the char-
acteristic lag, appear uncorrelated. Thus the single epoch approach is no longer
valuable for adding information. The requirement for intensive monitoring and
for larger samples of ob jects points to the need for a multiwavelength platform
with modest, very simple, very inexpensive optical/UV telescopes paired with
X-ray and gamma-ray detectors.
I am grateful to my many collaborators for their con-
Acknowledgments.
tributions to our multiwavelength monitoring of blazars, particularly those with
whom I have worked most closely in recent years, Laura Maraschi, Joe Pesce,
Elena Pian, Rita Sambruna, Aldo Treves, and Ann Wehrle, and those who were
PIs for the second campaign on PKS 2155–304, Tsuneo Kii, Herman Marshall,
and Greg Madejski.
Discussion
Al Marscher : In the 1994 PKS 2155–304 flare, you say that the fast rise is from
the time scale for particle acceleration but that the slower decay is geometric.
Light travel effects should limit the shorter time scale, not the longer. Also I dis-
agree that the longer time scales and lags at lower frequencies require gradients
in the underlying jet to be the root cause: shocks should be frequency-stratified
and give the behavior you observe (Marscher, Gear & Travis 1992, in Variability
of Blazars).
Meg Urry : You are right about the geometric limit for the fastest time scale. We
agree that the jet needs to be inhomogeneous. Whether the inhomogeneity arises
from the natural stratification due to energy-dependent losses and diffusion from
a shock, or from some other means, is not known at this point.
11
Hugh Al ler : I am worried about involving different physical processes to explain
two different flare events (1991 versus 1994) in PKS 2155–304.
Isn’t a single
process which permits both events to be explained ultimately simpler?
Meg Urry : I can see how it would seem that way but frankly, the character of
the variations in 1991 and in 1994 is so different that the required change in
parameters of the single process you might prefer is more or less equivalent to a
distinct process. If, for example, we decide that the 1994 flares are the signature
of a shock propagating through an inhomogeneous jet, with the expected depen-
dence of amplitude on wavelength and delays of ∼ 1 day from X-rays to EUV to
UV-emitting regions, then how can this model explain the 1991 variability, with
its achromaticity and negligible lags? In a sense these are opposite behaviors,
which naturally suggest unrelated causes.
References
Brinkmann, W., et al. 1994, A&A, 288, 433
Celotti, A., Maraschi, L., & Treves, A. 1991, ApJ, 377, 403
Courvoisier, T. J.-L., et al. 1995, ApJ, 438, 108
Dondi, L., & Ghisellini, G. 1995, MNRAS, 273, 583
Edelson, R. A., et al. 1995, ApJ, 438, 120
Ghisellini, G., & Madau, P. 1996, MNRAS, 280, 67
Koratkar, A., Pesce, J. E., Urry, C. M., & Pian, E. 1996, in preparation
Maraschi, L., et al. 1994, ApJ, 435, L91
Padovani, P., & Giommi, P. 1995, ApJ, 444, 567
Pesce, J. E., et al. 1996, ApJ, submitted
Pian, E., et al. 1996, ApJ, submitted
Pohl, M. et al. 1995, A&A, 303, 383
Punch M., Akerlof C. W., Cawley M. F., Chantell M. ., Fegan D. J., Fennell S.,
Gaidos J. A., Hagan J., Hillas A. M., & Jiang Y. 1992, Nature, 358, 477
Quinn J., et al. 1996, ApJ, 456, L83
Sambruna R., Maraschi L., & Urry C. M. 1996, ApJ, 463, 444
Smith, P. S., Hall, P. B., Allen, R. G., & Sitko, M. L. 1992, ApJ, 400, 115
Ulrich, M.-H., Maraschi, L., & Urry, C. M. 1996, ARAA, in preparation
Urry, C. M., et al. 1993, ApJ, 411, 614
Urry, C. M., et al. 1996, ApJ, submitted
Urry, C. M., & Padovani, P. 1995, PASP, 107, 803
Vermeulen R., & Cohen M. 1994, ApJ, 430, 467
Wagner S., & Witzel A. 1995, ARA&A, in press
Wehrle, A. E., Zook, A. C., Unwin, S. C., Urry, C. M., & Madejski, G. 1994,
in The Multi-Mission Perspective (High Energy Astrophysics Division
Meeting of the American Astronomical Society, Napa Valley, November
1994), p. 27
12
|
astro-ph/9712216 | 1 | 9712 | 1997-12-17T17:03:13 | Optical-IR Spectral Energy Distributions of z>2 Lyman Break Galaxies | [
"astro-ph"
] | Broadband optical and IR spectral energy distributions are determined for spectroscopically confirmed z>2 Lyman break objects in the Hubble Deep Field (HDF). These photometric data are compared to spectral synthesis models which take into account the effects of metallicity and of internal reddening due to dust. It is found that, on average, Lyman break objects are shrouded in enough dust (typically E(B-V)~0.3) to suppress their UV fluxes by a factor of more than 10. The dust-corrected star formation rate in a typical HDF Lyman break object is ~60 h^-2 M_sun/yr (q_0=0.5). Furthermore, these objects are dominated by very young (<~0.2 Gyr, and a median of ~25 Myr) stellar populations, suggesting that star formation at high redshift is episodic rather than continuous. Typically, these star formation episodes produce ~10^9 h^-2 M_sun of stars, or ~1/20 of the stellar mass of a present-day L* galaxy. | astro-ph | astro-ph |
Astronomical Journal, in press
Optical-IR Spectral Energy Distributions of z > 2 Lyman Break Galaxies 1 2
Marcin Sawicki and H. K. C. Yee
Department of Astronomy, University of Toronto, Toronto, Ontario M5S 3H8, Canada
email: [email protected], [email protected]
ABSTRACT
Broadband optical and IR spectral energy distributions are determined for spectro-
scopically confirmed z > 2 Lyman break objects in the Hubble Deep Field (HDF). These
photometric data are compared to spectral synthesis models which take into account
the effects of metallicity and of internal reddening due to dust. It is found that, on av-
erage, Lyman break objects are shrouded in enough dust (typically E(B − V ) ≈ 0.3) to
suppress their UV fluxes by a factor of more than 10. The dust-corrected star formation
rate in a typical HDF Lyman break object is ∼ 60h−2
100M⊙yr−1 (q0 = 0.5). Furthermore,
these objects are dominated by very young (∼< 0.2 Gyr, and a median of ∼ 25 Myr) stel-
lar populations, suggesting that star formation at high redshift is episodic rather than
continuous. Typically, these star formation episodes produce ∼ 109h−2
100M⊙ of stars, or
∼ 1
20 of the stellar mass of a present-day L∗ galaxy.
1.
INTRODUCTION
Recent observational studies of the high-redshift Universe have begun to yield direct informa-
tion on the nature of the population of "normal" galaxies at z > 2. Color-selected searches have
started to produce growing, spectroscopically-confirmed samples (e.g., Steidel et al. 1996a; Steidel
et al. 1996b; Pascarelle et al. 1996; Lowenthal et al. 1997; Steidel et al. 1997), while larger samples
of objects identified solely on the basis of multicolor photometry have been used to boost statistics
(e.g., Madau et al. 1996; Sawicki, Lin & Yee 1997; Connolly et al. 1997). Whether spectroscop-
ically confirmed or not, such high-z objects are often called "Lyman-break galaxies" or "U-band
drop-outs". The names reflect the fact that in these galaxies the UV flux is severly attenuated
1Based on observations made with the NASA/ESA Hubble Space Telescope obtained at the Space Telescope Science
Institute, which is operated by the Association of Universities for Research in Astronomy, Inc., under NASA contract
NAS 5-26555.
2Based on observations made at the Kitt Peak National Observatory, National Optical Astronomy Observatories,
which is operated by the Association of Universities for Research in Astronomy, Inc. (AURA) under cooperative
agreement with the National Science Foundation.
-- 2 --
by the combined effects of internal and intergalactic Lyman line and continuum opacity. Unlike
serendipitously discovered single objects, these new samples of high-z galaxies are large enough to
allow the study of global properties of the z > 2 galaxy population, such as their number density,
luminosity density, luminosity function, and rates of star formation and metal ejection.
Despite this recent progress, there is much that still remains unclear. Based on number densi-
ties, star formation rates, physical sizes, morphologies and masses inferred from spectral line widths,
Steidal et al. (1996a) and Giavalisco, Steidel, & Macchetto (1996) argue that Lyman break objects
are the spheroidal components of present-day luminous galaxies, seen in the act of formation. Using
diagnostics similar to those employed by Steidel et al. and Giavalisco et al., Lowental et al. (1997)
favor the interpretation that Lyman break objects are either progenitors of present-day low-mass
spheroidal galaxies or are starbursting building blocks of more massive galaxies of today (see also
Colley et al. 1997).
Estimates of star formation rates in Lyman break galaxies are based on their rest-frame UV
fluxes (e.g., Steidel et al. 1996a; Madau et al. 1996; Lowenthal et al. 1997). However, even moderate
amounts of dust may significantly suppress the UV flux and, hence, the inferred star formation rate.
Meurer et al. (1997) have studied the UV spectral indices of HDF Lyman break galaxies and have
deduced that, if ages of their stellar populations are ∼< 0.3 Gyr, reddening due to dust supresses
the UV flux fairly strongly (on average, a factor of 15 at 1600 A). Using the same spectral indices,
but assuming older ages (1 Gyr), Calzetti (1997) found the attenuation a factor of 3 lower than
that arrived at by Meurer et al. (1997). These two different results, obtained from the same data,
reflect the fact that both dust and aging have similar effects on the UV shape of the SED and that
one can mimick the effects of the other. This dust-age degeneracy can be lifted by the addition of
IR (i.e., rest-optical) data, which are sensitive to age but much less so to dust.
Ages of star-forming objects at z > 2 are not well known. Steidel et al. (1996a) feel that ages
younger than a few ×107yr are unlikely, arguing that such young ages would imply that a large
number of galaxies have to be bursting simultaneously. However, apart from the gravitationally
lensed z = 2.72 OB star absorption-line galaxy cB58 (Yee et al. 1996), for which fits to broadband
optical and IR colors give the age of the dominant stellar population to be ∼< 35 Myr (Ellingson
et al. 1996), no quantitative age measurements exist for Lyman break galaxies. Constraints on
ages will lift the dust-age degeneracy, in addition to providing clues to the evolutionary history of
Lyman break objects.
The Hubble Deep Field (HDF;Williams et al. 1996) is a small but intensly studied area of the
sky, for which both extremely deep multiwavelength imaging (e.g., Williams et al. 1996; Dickinson
et al. 1997; Hogg et al. 1997; Serjeant et al. 1997; Fomalont et al. 1997) and follow-up spectroscopy
(e.g., Steidel et al. 1996b; Cohen et al. 1996; Lowenthal et al. 1997; Zepf et al. 1997) have been
carried out. To date, 17 spectroscopically confirmed z > 2 Lyman break galaxies have been reported
in the HDF (Steidel et al. 1996a; Lowenthal et al. 1997).
In this paper we construct the broadband optical and IR (i.e., rest-frame UV and optical) spec-
-- 3 --
tral energy distributions (SEDs) of the 17 spectroscopically confirmed Hubble Deep Field Lyman
break galaxies (§ 2). These observed SEDs are then fitted with dust-corrected spectral synthesis
models described in § 3.1. These fits show that light from Lyman break galaxies is attenuated by
significant amounts of dust (typically E(B − V ) ≈ 0.3; § 3.2) and is dominated by very young
(∼< 0.2Gyr) stellar populations (§ 3.3). Furthermore, Lyman break objects have high rates of star
formation (typically ∼ 60h−2
100M⊙yr−1 for q0 = 0.5; § 3.4), and quite low stellar masses (typically,
∼ 109h−2
100M⊙; § 3.5). Since internal reddening in Lyman break galaxies appears to be significant,
in § 4.1 we calculate its effects on the recently-published estimates of z > 2 star formation and
metal ejection rates. In § 4.2 we speculate on the nature of Lyman break galaxies in view of the
high star formation rates and young ages of stellar populations which we found earlier. Our results
are summarized in § 5.
2. DATA
We study the broadband spectral energy distributions for the 17 spectroscopically confirmed
z > 2 HDF objects reported by Steidel et al. (1996b; 6 objects3, some with revised redshifts4)
and by Lowenthal et al. (1997; 11 objects). We measured broadband fluxes using both optical
(Williams et al. 1996) and infrared (Dickinson et al. 1997) data. Specifically, in the optical we used
the publicly-available Version 2, 4 × 4 binned, HST images of the HDF; in the IR we employed
the, also publicly available, Version 1 Kitt Peak 4m IRIM observations. The use of the IR data is
crucial, as at z ∼ 3 the optical data alone give coverage only bluewards of rest-2000 A, while the
IR extends this to rest-5500 A, thereby spanning the age-sensitive Balmer break. Altogether, the
available photometry spans 3000 -- 22000 A (or 750 -- 5500 A at z = 3) in 7 bandpasses (U300, B450,
V606, I814, J, H, Ks)5, although, as will be explained later, we do not include the U300 and B450
data in our fits.
We smoothed the HST images to match the poorer, ground-based point spread function (PSF)
of the IR images. We then performed photometry using the PPP faint galaxy photometry package
3Steidel et al. report five z > 2 redshifts. Since, morphologically, C4-06 appears to have two distinct components,
we have treated it as two objects at a common redshift.
4 The redshift of the Steidel et al. (1996b) object C5-02 has been revised from z = 2.845 to z = 2.008 (Steidel,
private communication). The reported redshift of C3-02 may also be erroneous: The observed broadband SED of
C3-02 (Figure 1) exhibits a break between the I814 and J filters and not, as would have been expected of the Balmer
break in a z = 2.775 object, redward of J. We note that the absorption lines of C IV 1550, Fe II 1608, and Al II 1670
could be misidentified as Ly α 1216, Si II 1260, and O I 1303. In this case the redshift of C3-02 would be z = 1.95,
which is consistent with the photometric redshift zphot = 2.1 of Sawicki et al. (1997) and with the placement of the
Balmer break between the I814 and J filters. While adopting the revised redshift for C5-02, we retain the published
value for C3-02. We tested the effects of this possible redshift misidentification on our analysis and found them to
be small.
5U300, B450, V606, and I814 are used to denote the HST filters F 300W , F 450W , F 606W , and F 814W .
-- 4 --
(Yee 1991). Colors were measured in "color apertures": for each object, a color aperture was chosen
so as to maximize the S/N. Photometric calibration was based on the zeropoints provided in the
STScI-HDF and KPNO-IRIM-HDF web pages (Ferguson 1996; Dickinson 1996), while conversion
to AB magnitudes (for the IR data) was done using the Fν (0 mag) fluxes of Wemstaker (1981).
The photometry results are presented in Table 1. The magnitudes listed are "total magnitudes"
obtained by correcting the object's magnitude within the color aperture by its growth curve in the
V606 image. A typical total magnitude aperture is 3.2′′.
3. COMPARISON WITH SPECTRAL SYNTHESIS MODELS
3.1. The Models
For the analysis of the stellar populations of Lyman break galaxies, we generated a set of
model broadband SEDs constructed in the following way: We started with the Bruzual & Charlot
(1996) multi-metallicity spectral synthesis models; specifically, we used the KL96 theoretical stellar
spectra6 with the Salpeter 0.1 < m
< 125 IMF. These Bruzual & Charlot models were attenuated
M⊙
with Calzetti's (1997; see also Calzetti et al. 1994) empirical reddening recipe for star-forming
galaxies7 , covering a range of E(B − V ) values. The reddened SEDs were appropriately redshifted
and then convolved with the instrumental transmission curves (in the case of the HST data) or filter
transmission curves (for the Kitt Peak IR observations) producing a suite of model colors. For each
object this suite of model colors contains two possible star formation histories (instantaneous burst
and constant star formation), 221 ages (0 -- 20 Gyr), three different metallicities (0.02Z⊙, 0.2Z⊙,
and 1.0Z⊙), and a range of reddening (E(B − V ) = 0.0 to 0.5 in steps of 0.02).
Why were these particular ingredients chosen for the models? Initial mass function: The initial
mass function, at least in low-redshift starbursts, appears to be independent of environment and is
consistent with the Salpeter IMF with a high (∼ 100M⊙) upper mass cutoff (Stasi´nska & Leitherer
1996; see also Massey et al. 1995). We have thus chosen to use the Salpeter 0.1 < m
< 125 IMF
M⊙
available in the Bruzual & Charlot (1996) library. Reddening curve: Since, by virtue of their color
selection criteria (Steidel et al. 1996b; Lowenthal et al. 1997), the 17 spectroscopically confirmed
HDF Lyman break objects are star-forming, the use of the Calzetti (1997) reddening curve for
6In general, substantial differences still exist between composite SEDs based on theoretical and empirical stellar
spectra. We have, however, tested our dust and age results with Bruzual & Charlot solar-metallicity empirical models
and found that the differences were small for our purposes.
7Calzetti (1997) gives:
Fobs(λ) = F0(λ)10−0.4E(B−V )k(λ),
where
k(λ) = (cid:26) [(1.86 − 0.48/λ)/λ − 0.1]/λ + 1.73
2.656(−2.156 + 1.509/λ − 0.198/λ2 + 0.011/λ3) + 4.88
for 0.63µm ≤ λ ≤ 1.0µm
for 0.12µm ≤ λ < 0.63µm
-- 5 --
star-forming regions is appropriate. Unlike the commonly-used LMC and SMC reddening curves
(Fitzpatrick 1986; Bouchet et al. 1985) which are are derived for stars alone, the Calzetti curve
holds for star-forming regions; it thus automatically includes such effects as back-scattering and
geometrical distribution of dust and is therefore much more appropriate for correcting the effects
of dust on the photometry of star-forming galaxies (Gordon, Calzetti, & Witt 1997). Using the
LMC reddening law would produce extinction values 2.5 -- 3 times lower than that which we find
(§ 4.1) with the more appropriate Calzetti extinction curve. Metallicity: Damped Ly-α systems at
z = 2− 3 have metallicities Z ∼< 0.1Z⊙ (Pettini et al. 1994; Lu et al. 1996; Pettini et al. 1997a); thus
0.2Z⊙ or 0.02Z⊙ are reasonable values to expect for the metallicity of Lyman break galaxies, while
1.0Z⊙ provides a useful check in case the metallicity of Lyman break objects were significantly
higher than that of damped Ly-α clouds. Ages: At z ≈ 3 the Universe is only 1 -- 2 Gyr old and no
object within it can be older than that; we have, however, allowed for model ages of up to 20 Gyr
as a consistency check of the fits. Star formation history: In reality, the most likely star-formation
scenario is that of a burst in which the star formation rate (SFR) declines with time. The two star
formation histories used (instantaneous burst and constant SFR) thus provide the bracketing cases
of the likely star formation histories of these objects. Dust contents and ages of stellar populations
obtained from these two extremes will then bracket the actual values of age and reddening.
For high-redshift objects, the neutral hydrogen contained in Ly-α clouds along the line of
sight provides an extra source of opacity in the UV (e.g., Madau 1995). The amount of extinction
provided by this mechanism is stochastic and depends on the numbers, redshift distribution, and
column densities of the Ly-α clouds along the line of sight to the target galaxy. The galaxy's
UV flux is affected blueward of rest-1216 A and is almost completely extinguished below rest-
912 A. Whereas this effect is of great usefulness in identifying high-z galaxies, its stochastic nature
introduces uncertainties below ∼ 1200 A in the SED models for individual galaxies. These stochastic
uncertainties, combined with the fact that the dust reddening curve is not well known below 1200 A,
led us to decide not to fit the two bluest filters (U300 and B450). Thus all the fits presented in this
paper have been done using the 5 reddest filters (V606, I814, J, H, and Ks).
Metallicity (Z = 0.02Z⊙, 0.2Z⊙, and 1.0Z⊙) and star formation history (instantaneous burst
and constant SFR) were fixed, giving six scenarios. In each of these six scenarios, age since the onset
of star formation and amount of dust were free parameters. The models and the data were compared
by means of χ2 minimization. As an illustration, Figure 1 shows the fits thus obtained with the
0.2Z⊙ continuous-SFR models. Instantaneous burst models, as well as those for Z = 0.02Z⊙ and
1.0Z⊙, produced fits of similar quality. Note that the two bluest filters, although not used to
produce the fit, match the best-fitting model SEDs quite well.
Because reddening increases towards shorter wavelengths, the part of the SED which contains
information about dust attenuation is the rest-frame UV. Dust attenuation can, however, be mim-
icked by aging of the stellar population: as the population ages, massive stars die out and the UV
flux decreases. For example, below rest-3000A a zero-age SED with E(B − V ) = 0.3 looks very
much like a 100 Myr, constant SFR, dust-free one. The availability of IR data allows the lifting
-- 6 --
of this degeneracy between age and reddening: at z > 2, IR data are probing the rest-optical part
of the SED and consequently measure the age by means of the size of the Balmer break and, to
a lesser degree, the slope of the continuum redward of rest-4000A. Metallicity effects could pose
a problem because of the age-metallicity degeneracy (e.g., Worthey 1994), but we have monitored
their effects by performing fits with models of different metallicity (0.02Z⊙, 0.2Z⊙, and 1.0Z⊙). As
will be seen later, no overwhelming metallicity effects have been found.
With increasing redshift, the rest wavelengths probed by each filter move more and more
blueward (by z ≈ 3, the H filter is centered at 4000A) and consequently less and less information
is available about the age-sensitive rest-optical part of the SED. Thus, past z = 3, only one filter
(Ks) is redward of the 4000A break. Because of this loss of age-sensitive information at higher
redshifts, we have chosen to split the sample at z = 3. We consider fits for the 11 objects at z < 3
to be much more reliable than those for the six at z > 3. The results and discussion presented
below are mainly based on the 11 z < 3 galaxies; we will, however, retain the six z > 3 objects for
completeness and as a consistency check.
3.2. Dust Content
For each object the models described above fit simultaneously the dust content and the age
of the dominant stellar population. The age and E(B − V ) parameters of the best-fitting 0.2Z⊙
models are plotted in Figure 2, with error bars corresponding to 90% confidence limits. The Figure
illustrates that the Lyman break objects are best fitted by models which require non-zero amounts
of dust. The median values for both the instantaneous burst and constant SFR 0.2Z⊙ scenarios
are E(B − V ) = 0.28, producing a factor of 16 attenuation at 1600A. This attenuation value is in
agreement with the result derived from the UV spectral indices by Meurer et al. (1997) (see also
Calzetti 1997). Note, however, that Meurer et al. had to assume (correctly, as we shall see in § 3.3)
that the z > 2 galaxies are no older than 0.3 Gyr.
Just as in the 0.2Z⊙ case, the 0.02Z⊙ and 1.0Z⊙ models reqiure significant amounts of dust.
The results are summarized in Figure 3 which shows the distribution of E(B − V ) values for the
six different scenarios. Dust is required for the vast majority of galaxies in all of the six scenarios.
For sub-solar metallicities (0.02Z⊙ and 0.2Z⊙) typical values are E(B − V ) ≈ 0.3 and somewhat
lower (E(B − V ) = 0.15 − 0.2) for the unlikely case of solar metallicity.
Can dust-free models be securely ruled out? The error bars in Figure 2 are 90% confidence
limits; hence, at 90% confidence, non-zero amounts of dust are unavoidable in all but one or two
of the 11 z < 3 objects. The presence of dust is further illustrated in Figure 4, where, for three
objects, the best possible dust-free models are compared with those in which dust is allowed. More
qualitatively, the median reduced χ2 are 1.34 and 1.68 for the instantaneous burst ad constant SFR
0.2Z⊙ fits with dust; the corresponding dust-free values are 5.04 and 10.1. Models with dust give
much better fits than those without it. Therefore we conclude that dust is present in Lyman break
-- 7 --
galaxies. A similar conclusion was reached by Ellingson et al. (1996) for the z = 2.72 galaxy cB58,
which has relatively high-precision IR photometry.
The high reddening values obtainded here from broadband photometry of Lyman break galaxies
are in contrast to the negligible extinctions inferred from chemical abundances in damped Lyα
systems (Lu et al. 1996; Pettini et al. 1997a). However, while quasar sightlines probe essentially
random regions of damped Lyα galaxies, broadband photometry targets regions of star formation.
Thus, if the star-forming regions have substantially different dust-to-gas ratios than "random"
parts of high-redshift galaxies, then the discrepancy between the damped Lyα spectroscopic and
our photometric results may simply reflect the differences in the environments that the two methods
probe.
Star formation rates have been derived from the fluxes of either Hα or Hβ for the small number
of high-redshift objects for which these fluxes have been measured (Bechtold et al. 1997; Pettini
et al. 1997b). In these objects, line-based SFR measurements are generally lower than the dust-
corrected star-formation rates derived from UV fluxes. This apparent discrepancy can, however,
be caused by a number of effects other than discrepant dust measurements; these effects include
leakeage of Lyman-limit photons from H II regions, absorption of Lyman-limit photons by dust,
and a non-standard IMF (see Bechtold et al. 1997 and references therein). These effects may affect
the reliability of SFR measurements based on emission lines. Ultimately, concurrent measurements
of Hα/Hβ ratios will provide a robust and independent estimate of the dust content in Lyman
break galaxies.
Based on our analysis, the presence of dust in Lyman break galaxies seems unavoidable with
E(B − V ) ≈ 0.3 being typical. As will be discussed in § 4.1, these amounts of dust are large enough
to significantly affect estimates of star formation and metal ejection rates at high redshift.
3.3. Ages of Stellar Populations
For each galaxy, we shall date the age of the galaxy's dominant stellar population. It must
be emphashized that the age of the dominant stellar population does not mean the age of the
galaxy itself, but rather corresponds to the time since the onset of the most recent major episode
of star formation. Note that the instantaneous burst and constant SFR models are the limiting-
case scenarios, while a galaxy's star formation history probably lies somewhere between these two
extremes. Thus, the instantaneous burst model provides the lower bound, while the constant SFR
model gives the upper bound on the likely age of the galaxy's current episode of star formation.
Figure 2 presents the parameters of the best-fitting 0.2Z⊙ models, while Figure 5 shows the
distribution of galaxy ages for all six combinations of metallicity and star formation history. As
these Figures illustrate, in the constant SFR scenario the vast majority of z < 3 objects, for
which spectral coverage is sufficient in the red, are best fit with models for which the ages of
dominant stellar populations are less than 0.2 Gyr. In the instantaneous burst scenario, the ages
-- 8 --
of virtually all the objects, including the six with z > 3, are less than 0.1 Gyr. The median ages
for the z < 3 objects are 36 Myr and 10 Myr in the 0.2Z⊙ constant SFR and instantaneous burst
scenarios, respectively. Although objects at z > 3 have insufficient IR coverage to constrain ages
reliably, their fits are also consistent with very young stellar populations; for the instantaneous
burst scenarios these ages are as young as for the z < 3 objects.
How confidently can we rule out older ages of stellar populations? The error bars in Figure 2
are 90% confidence limits. If we consider the constant SFR model (which is more favorable for long
ages), we see that of the z < 3 objects all but one must have ages < 1 Gyr at the 90% confidence
level. A further illustration of the necessity of young ages is given in Figure 6: the (forced) 1 Gyr-
old models produce poorer fits than the unforced younger models. More qualitatively, whereas the
median reduced χ2 are 1.34 and 1.68 for the 0.2Z⊙ instantaneous burst and constant SFR models
with free age, the corresponding values for the forced 1Gyr fits are 355.2 and 6.67, respectively.
We therefore conclude that the z > 2 Lyman break galaxies are dominated by very young stellar
populations.
< 125 IMF) may not reflect the true IMF in the Lyman break galaxies.
A possible concern is that the IMF employed to generate the models (i.e., the Salpeter
0.1 < m
In particu-
M⊙
lar, the determination of the age of the stellar population is primarily sensitive to the size of the
Balmer break, which grows as the population ages. Since low-mass stars are responsible for the
Balmer break, an IMF deficient in low-mass stars would produce a Balmer break that is smaller
than expected for a given age, thereby mimicking the signature of a younger stellar population.
Consequently, for an IMF deficient in low-mass star, our fits would underestimate the ages of stellar
populations. The multi-metallicity SEDs of Bruzual & Charlot (1996) are available only for the
0.1 < m
< 125 IMF. Hence, to estimate the impact of an IMF with a high lower-mass cutoff, we
M⊙
have compared the predicted colors of Bruzual & Charlot (1993) solar metallicity models with dif-
ferent IMF mass ranges. Specifically, we have compared the 0.1 < m
< 125
M⊙
Salpeter IMFs. The comparison reveals that, for the young ages in question and if the IMF is
deficient in low-mass stars, the age of the star-forming population can be underestimated by a
factor of ∼ 5 in the instantaneous burst scenarios; in the constant SFR models the underestimate
is only at the 10% level. These effects would then increase the median age in either scenario to no
more than ∼ 50 Myr. IMF effects are thus unlikely to affect our conclusion that the Lyman break
objects have very young stellar populations.
< 125 and 2.5 < m
M⊙
Figure 7 shows, as a function of redshift, ages of stellar populations obtained with the 0.2Z⊙
SEDs. The age of the q0 = 0.5, t0 = 12.5 Gyr Universe and the age of an object which formed at
zf = 4.5 are shown for comparison. Lyman break galaxies are, by and large, dominated by very
young stellar populations. The absence of older (age > 0.2 Gyr) objects, particularly at the lower
end of the redshift range covered, is remarkable. Possible reasons for this absence will be discussed
in § 4.2.
-- 9 --
3.4. Star Formation Rates
To estimate the star formation rates in Lyman break galaxies, we use the constant SFR fits.
For each of the observed objects, we obtain the SFR by comparing the normalization of the fit to
that of a 1M⊙yr−1 model of identical age and reddening. For the z < 3 objects, 0.2Z⊙ constant
SFR models, and q0 = 0.5, the median SFR is 59h−2
100M⊙yr−1 (167h−2
100M⊙yr−1 for q0 = 0.05). The
distribution of star formation rates in our sample is shown in Figure 8.
As a check on the star formation rates derived above from SED fitting, we can use the predicted
rest-frame 1500 A flux of a 9 Myr old, 106M⊙ galaxy with a 100 < m
< 1 Salpeter IMF (Leither et
M⊙
al. 1995). Comparing the dust-corrected, observed fluxes of the HDF Lyman break objects against
the fiducial flux of Leitherer et al. gives a median SFR of 63h−2
100M⊙yr−1 (q0 = 0.5), thereby
confirming the SFR value obtained earlier from SED fitting.
For illustrative purposes, we have also computed the SFRs under the assumption that the
Lyman break objects are dust-free. The resulting median star formation rate is 2h−2
100M⊙yr−1 for
q0 = 0.5 (6h−2
100M⊙yr−1 for q0 = 0.05), in agreement with the dust-free SFR estimates given by
Steidel et al. (1996a) and Lowenthal et al. (1996). However, since models which include dust produce
much better fits, we prefer the ∼ 60h−2
100M⊙yr−1 for q0 = 0.05) dust-corrected
rates of star formation.
100M⊙yr−1 (∼ 170h−2
3.5. Stellar Masses
To estimate the stellar masses produced in Lyman break galaxies, we can use the results of
either the instantaneous burst fits or the constant SFR fits. For the instantaneous burst fits, we
compute the stellar mass formed in each of the observed objects by comparing the normalization
of the fit to that of a 1M⊙ model of identical age and reddening. For the z < 3 objects, 0.2Z⊙
instantaneous burst models, and q0 = 0.5(0.05), the median mass in stars is 1(2) × 109h−2
100M⊙, or
about 1
15 ) the stellar mass of a present-day L∗ galaxy (3 × 1010h−1
30 ( 1
100M⊙; Cowie et al. 1995).
Under the constant SFR assumption, the typical duration of a starforing espisode is 2× the
median observed age. Using the ages and SFRs from the constant SFR fits, we get the median
stellar mass to be 2(4) × 109h−2
100M⊙ for q0 = 0.5(0.05) -- in agreement with the value derived from
instantaneous burst fits. As was the case for star formation rates, comparing dust-corrected UV
fluxes to Leitherer et al. models confirms the above masses. The distributions of stellar masses
for both the instantaneous burst and constant SFR models are shown in Figure 9(b -- c). In both
scenarios, the stellar masses of Lyman break objects are generally smaller than the stellar mass of
a present-day L∗ galaxy. Stellar masses appear to be relatively insensitive to the details of star
formation history.
As an illustration, the above analysis can also be applied to the fits obtained with dust-free
models. For q0 = 0.5(0.05), under the instantaneous burst assumption, the resulting median stellar
-- 10 --
mass is 1(4) × 109h−2
100M⊙. Using the ages and star formation rates of the constant-SFR fits gives
the median stellar mass of 6(16) × 109h−2
100M⊙. These values, obtained from dust-free model fits,
are surprisingly similar to the stellar masses derived from models which account for dust. In the
instantaneous burst case, this agreement arises because the dust-free fits give older ages, which
necessitates larger stellar masses to account for the observed fluxes. In the constant SFR case the
agreement occurs because the lower star formation rate of the dust-free models are offset by their
older ages.
The relative insensitivity of the total stellar mass formed to the assumed star formation history
and dust content of the galaxy allows us to obtain a robust result for the typical mass produced by
a star formation episode in a Lyman break galaxy. Averaging the median masses calculated from
the dust-corrected instantaneous burst and constant SFR fits, we have that the median stellar mass
produced in an episode of star formation in an HDF Lyman break galaxy is ∼ 1.5(3) × 109h−2
100M⊙
for q0 = 0.5(0.05). This median stellar mass is within the 1010M⊙ upper limits on the gravitational
masses of HDF Lyman break galaxies which were obtained by Lowenthal et al. (1997) on the basis
of Lyα emission line widths and is equivalent to ∼ 1
10 ) the stellar mass of a present-day L∗
galaxy.
20 ( 1
4.
IMPLICATIONS
4.1. Dust Corrections
Intervening dust has the effect of absorbing UV photons ,thereby suppressing the observed
UV flux. If, as the SED fits discussed in § 3.2 indicate, z > 2 objects have substantial amounts of
internal reddening, recent measurements of the rates of star formation (Steidel et al. 1996a; Madau
et al. 1996; Lowenthal et al. 1997) and metal ejection (Madau et al. 1996; Sawicki et al. 1997) need
to be revised.
Steidel et al. (1996a) have used rest-1500A fluxes to estimate the average star formation rate of
their Lyman break galaxy sample to be 2h−2
100M⊙yr−1 for q0 = 0.05).
If we assume that their objects suffer from the same amount of reddening as is typical for the Lyman
break objects in the Hubble Deep Field (E(B − V ) ≈ 0.28), then their SFR estimates need to be
adjusted upward to 38h−2
100M⊙yr−1 for q0 = 0.5 (and 6h−2
100M⊙yr−1 for q0 = 0.5 (and to 111h−2
100M⊙yr−1 for q0 = 0.05).
On the basis of 1500A fluxes, Lowenthal et al. (1997) made estimates of star formation rates
both in individual z ≈ 3 HDF objects and in the volume-averaged population. Reddening correction
would bring the Lowenthal et al. range of star formation rates to 13 − 35h−2
100M⊙yr−1 for q0 = 0.5
(31− 111h−2
100M⊙yr−1 for q0 = 0.05). The volume-averaged star formation rates given by Lowenthal
et al. need likewise be adjusted upwards by a factor of ∼ 18.
Madau et al. (1996) studied a sample of U- and B-band dropouts in the Hubble Deep Field.
They used the comoving UV (1620 A at hzi = 2.75 and 1630 A at hzi = 4.0) luminosity densities
-- 11 --
to estimate the volume-averaged rates of metal ejection and star formation. If we apply the typical
E(B − V ) = 0.28 extinction found to occur in z > 2 galaxies, the Madau et al. estimates of both
the star formation and metal ejection rates need to be revised upward by a factor of 16. This value
is consistent with the factor of 15 derived by Meurer et al. (1997) on the basis of optical data alone
and with the (as we have seen, correct) assumption of young ages. Note, however, that the Madau
et al. (1996) estimates do not include incompleteness corrections, so they remain lower limits even
after having been corrected for dust.
As part of their photometric redshift study of the Hubble Deep Field, Sawicki et al. (1997)
have calculated the expected metal density of the Universe as a function of redshift. To do so, they
have used rest-3000A luminosity densities. Applying E(B − V ) = 0.28 of extinction yields a factor
of 7 increase in flux at 3000A, requireing a factor of 7 increase in the 2 ∼< z ∼< 3.5 metallicity of the
Universe shown in Fig. 10. of Sawicki et al.
In summary, the values of star formation and metal ejection rates presented in the literature
are underestimates, since they do not account for intrinsic reddening. Though the exact correction
factors depend on the wavelength used for the original estimate, once correction for dust is made,
both star formation and metal ejection rates go up by about an order of magnitude. Note that
although the total mass of stars formed per object is similar for the dust-free and dust-included
models (§ 3.5), the difference in the volume-averaged star formation and metal ejection rates of the
Universe enters via the very different duty cycle of visibility which the Lyman break objects have
in these two models.
4.2. Ages and Star Formation Histories of Lyman Break Galaxies
For all but one of the z < 3 HDF Lyman break galaxies, the ages of the dominant stellar
populations are < 0.2 Gyr. Since our redshift range (2 < z < 3.5) spans a Gyr in time, the lack of
a significant number of galaxies with old (> 0.2 Gyr) stellar populations suggests that, in Lyman
break galaxies, the duration of star formation is short rather than extended and continuous.
In the rest of this section we will speculate on some of the possible interpretations of this
episodic star formation. Using simple calculations, we will show that under this single-burst scenario
the number densities of Lyman break objects, in addition to their stellar masses, are consistent with
Lyman break objects being progenitors of present-day L < L∗ galaxies. They are unlikely to be
the present-day L ≥ L∗ galaxies caught in the act of creation unless one invokes more complicated
mechanisms such as mergers or recurrent bursts of star formation.
In the simplest scenario, akin to the low-mass starburst model of Lowenthal et al. (1997), each
Lyman break galaxy undergoes a single episode of star formation. Since the stellar mass produced
in a typical Lyman break galaxy is ∼ 109h−2
100M⊙ (§ 3.5), such a burst would result in a low-mass
object whose stellar mass is ∼ 1
20 that of a present-day L∗ galaxy. This typical mass is thus too
low to account for the whole stellar mass of the bulge of a present-day L∗ galaxy.
-- 12 --
The young ages of their stellar populations, together with the fact that they will fade rapidly
after star formation ends (by ∼ 1 magnitude in a mere 30 Myr), mean that only a small fraction
of such short-burst galaxies are visible at any one time. The redshift range covered by the HDF
Lyman break sample (2 < z < 3.5), spans a time interval of ∼ 0.8 Gyr in a h100 = 0.7, q0 = 0.5
universe (∼ 2.3 Gyr for h100 = 0.7, q0 = 0.05). The median age from the 0.2Z⊙, constant SFR fits
is ∼ 35 Myr. Assuming that a typical star-forming burst lasts twice that time, and that it takes
∼ 30 Myr for the galaxy to fade ∼ 1 magnitude (and hence out of the spectroscopic sample), we
estimate that a typical Lyman break galaxy stays in the sample for ∼ 100 Myr. Therefore, for
every galaxy that is detected, there will be ∼ 8 (∼ 23 for q0 = 0.05) objects which have faded
beyond spectroscopic detectibility. Hence, in the volume and redshift range sampled, there should
be a total of ∼ 136 (∼ 391, for q0 = 0.05) visible plus faded objects which are undergoing, or have
undergone, brief but intense bursts of star formation.
The comoving number density of this combined (both visible and faded) population is 7 ×
100Mpc−3 for q0 = 0.05). In the Ks-band (which corresponds to ∼ rest-
100Mpc−3 (4 × 10−2h3
10−2h3
V ), the HDF Lyman break galaxies span a range of ∼ 2.5 magnitudes. Locally, over a corresponding
magnitude range, one expects to see ∼ 2 × 10−2h−3
20 L∗ (Loveday et
al. 1992; Lin et al. 1996). The number density of HDF Lyman break objects is thus similar to, or
perhaps somewhat higher than, that of local 1
20 L∗ galaxies. Thus, on the basis of their number
densities and stellar masses, Lyman break objects are consistent with being progenitors of present-
day sub-L∗ galaxies.
100Mpc−3 galaxies with L ≈ 1
One could postulate that present-day L > L∗ galaxies formed through bursts of star formation
which are analogous to those which seem to be producing the above-mentioned low-mass objects,
but which are more intense though less numerous than the bursts seen in the HDF. Based on the
present-day luminosity function of Lin et al. (1996), and assuming that the 17 objects seen in the
HDF are precursors of present-day sub-L∗ galaxies, one would expect the HDF to yield ∼ 5 Lyman
break objects with stellar masses ≥ 3×1010M⊙, the expected stellar mass of an L∗ galaxy. The fact
that such massive stellar populations are absent (Figure 9) implies that most present-day massive
galaxies did not form in single bursts of star formation at z > 2.
There are, however, alternative ways to assemble massive objects. One option is through
galaxy-galaxy merging: the present-day merger rate is rather low but increases as (1 + z)∼3−4, at
least to moderate redshift (e.g., Yee & Ellingson 1995; Patton et al. 1997). If this redshift trend
continues to z ∼ 3, then the merging of a dozen or so Lyman break objects can easily result in the
spheroid of a present-day L∗ galaxy.
Another possibility is akin to the "Christmas tree" model of Lowenthal et al. (1997; see also
Colley et al. 1996). In this model, star formation is episodic but recurrent within each galaxy: star-
forming episodes are separated by quiescent intervals during which the galaxy temporarily fades
out of the sample. The underlying stellar population, resulting from previous star-forming bursts,
would be too faded to be detectable in the presence of an ongoing burst (e.g., Ellingson et al. 1996).
-- 13 --
Each of the recurrent star-forming bursts would then add, typically, ∼ 109h−2
the galaxy, gradually building up its stellar mass. A spheroid containing ∼ 3 × 1010h−1
of stars could be thus accumulated in a dozen or two such recurrent starburst episodes.
100M⊙ of new stars to
100M⊙ worth
5. CONCLUSIONS
We have fitted the broadband spectral energy distributions of spectroscopically confirmed
z > 2 HDF Lyman break objects, using the multi-metallicity spectral synthesis models of Bruzual
& Charlot (1996). In the fits we have included correction for internal dust extinction typical of
local star-forming galaxies (Calzetti 1997). The fits also assume that the IMF at high redshift is
not unlike that seen locally, although some leeway is allowed, particularly in the lower mass cutoff.
We find that Lyman break galaxies are dominated by very young stellar populations (< 0.2
Gyr). The absence of objects with old stellar populations, particularly at lower redshifts (z ∼> 2),
implies that star formation in a typical Lyman break galaxy cannot go on continuously for a
prolonged time. Instead, star formation must occur in bursts of short duration (t < 0.2 Gyr).
A typical Lyman break galaxy is shrouded in enough dust to suppress its UV flux by a factor of
∼ 16 at 1600A. Consequently, recent UV-based estimates of the rates of star formation and metal
ejection at high redshift (Steidel et al. 1996a; Madau et al. 1996; Sawicki et al. 1997; Lowenthal et
al. 1997) need to be adjusted upwards by a similar factor.
Star formation rates in Lyman break galaxies are high (median of 59h−2M⊙yr−1 for q0 = 0.5).
These star formation rates typically produce 109h−2M⊙ of stellar mass during the lifetime of a star-
forming episode; this number is robust as it is relatively insensitive to the details of star formation
history and dust content. This median stellar mass is equivalent to 1
20 the stellar mass contained
in a present-day L∗ galaxy. Objects with larger stellar masses may be built up through recurrent
episodes of star formation, or can be assembled through mergers.
15 -- 1
A stellar population will fade rapidly after the end of a brief episode of star formation (∼
1 magnitude at rest-2000A in the first 30 Myr). Because star formation in Lyman break galaxies
is of brief duration, a substantial population of objects which have just recently faded out of the
spectroscopic sample may exist at high redshift. We estimate that if a typical Lyman break galaxy
stays in our sample for 0.1 Gyr, then there are 8 -- 23 such faded galaxies for every visible one. The
100 mag−1 Mpc−3) is
number density of the combined visible and faded population (a few ×10−2 h3
comparable to the number density of present-day sub-L∗ (MB ∼> −18) galaxies.
On the basis of the evidence presented above, we conclude that Lyman break objects are likely
direct progenitors of present-day sub-L∗ galaxies, or that they may form luminous galaxies through
merging or by repeated episodes of star formation. They do not appear to be steadily star-forming
direct progenitors of present-day massive galaxies.
-- 14 --
We thank Gabriela Mall´en-Ornelas for a very thorough reading of an earlier version of this
paper. We also thank Huan Lin, Bob Abraham, Gerhardt Meurer, and the anonymous referee for
discussions and comments. This work was financially supported by NSERC of Canada.
-- 15 --
REFERENCES
Bechtold, J., Yee, H.K.C., Elton, R., & Ellingson, E. 1997, ApJ, 477, L29
Bouchet, P., Lequeux, J., Maurice, E., Pr´evot, L., & Pr´evot-Burnichon, M. L. 1985, A&A, 149, 330
Bruzual A., G., & Charlot, S. 1993, ApJ, 405, 538
Bruzual A., G., & Charlot, S. 1996, in preparation
Calzetti D., 1997, to appear in the Proceedings of the Conference "The Ultraviolet Universe at Low
and High Redshift" preprint: astro-ph/9706121
Calzetti, D., Kinney, A., & Storchi-Bergmann, T. 1994, ApJ, 429, 582
Cohen, J. G., Cowie, L. L., Hogg, D. W., Songaila, A., Blandford, R., Hu, E. M., & Snopbell, P.
1996, ApJ, 471, L5
Colley, W.N., Gnedin, O.Y., Ostriker, J.P., & Rhoads, J.E 1997, ApJ, 488, 579
Connolly, A.J., Szalay, A.S., Dickinson, M., SubbaRao, M.U., & Brunner, R.J. 1997, ApJ, 486, L11
Cowie, L.C., Hu, E.M., & Songaila, A. 1995, Nature, 377, 603
Dickinson, M. 1996,
http://www.stsci.edu/ftp/science/hdf/clearinghouse/irim/irim zeropts.html
Dickinson, M., et al. 1997, in preparation.
Ellingson, E., Yee, H.K.C., Bechtold, J., & Elston, R. 1996, ApJ, 466, L71
Ferguson, H. 1996, http://www.stsci.edu/ftp/observer/hdf/logs/zeropoints.txt
Fitzpatrick, E.L., 1986, AJ, 92, 1068
Fomalont, E. B., Kellermann, K. I., Richards, E. A., Windhorst, R. A., & Partridge, B. P. 1997,
ApJ, 475, L5
Giavalisco, M., Steidel, C.C., & Macchetto, F.D. 1996, ApJ, 470, 189
Gordon, K. D., Calzetti, D., & Witt, A. N. 1997, ApJ, 487, 626
Hogg, D.W., Neugebauer, G., Armus, L., Matthews, K., Pahre, M.A., Soifer, B.T., & Weinberger,
A.J. 1997, AJ, 113, 474
Leitherer, C., Robert, C., & Heckman, T.M. 1995, ApJS, 99, 173
Lin, H., Kirshner, R.P., Shectman, S.A., Landy, S.D., Oemler, A., Tucker, D., & Schechter, P.L.
1996, ApJ, 464, 60
-- 16 --
Loveday, J., Peterson, B.A., Efstathiou, G., & Maddox, S.J. 1992, ApJ, 390, 338
Lowenthal, J.D., Koo, D.C., Guzm´an, R., Gallego, J., Phillips, A.C., Vogt, N.P., Illingworth, G.D.,
& Gronwall, C. 1997, ApJ, 481, 673
Lu, L., Sargent, W.L.W., Barlow, T.A., Churchill, C.W., & Vogt, S.S. 1996, ApJS, 107, 475
Madau, P. 1995, ApJ, 441, 18
Madau, P., Ferguson, H.C., Dickinson, M.E., Giavalisco, M., Steidel, C.C., & Fruchter, A. 1996,
MNRAS, 283, 1388
Massey, P., Lang, C.C., DeGioia-Eastwood, K., & Garmany, C.D. 1995, ApJ, 438, 188
Meurer, G.R., Heckman, T.M., Lehnert, M.D., Leitherer, C., & Lowenthal, J. 1997, AJ, 114, 54
Pascarelle, S. M., Windhorst, R. A., Keel, W. C., & Odewahn, S. C. 1996, Nature, 383, 45
Patton, D.R., Pritchet, C.J., Yee, H.K.C., Ellingson, E., & Carlberg, R.G. 1997, ApJ, 475, 29
Pettini, M., Smith, L.J., Hunstead, R.W., & King, D.L. 1994, ApJ, 426, 79
Pettini, M., King, D.L., Smith, L.J., & Hunstead, R.W. 1997, ApJ, 478, 536
Pettini, M., Steidel, C.C., Adelberger, K.L., Kellogg, M., Dickinson, M., & Giavalisco, M. 1997,
to appear in Origins, ed. J.M. Shull, C.E. Woodward, and H. Thronson (ASP Conference
Series); preprint: astro-ph/9708117
Sawicki, M.J., Lin, H., & Yee, H.K.C. 1997, AJ, 113, 1
Serjeant, S., Eaton, N., Oliver, S., Efstathiou, A., Goldschmidt, P., Mann, R.G., Mobasher, B.,
Rowan-Robinson, M., Sumner, T., Danese, L., Elbaz, D., Franceschini, A., Egami, E.,
Kontizas, M., Lawrence, A., McMahon, R., Norgaard-Nielsen, H.U., Perez-Fournon, I., &
Gonzalez-Serrano, I. 1997, MNRAS, 289, 457
Stasi´nska, G., & Leitherer, C. 1996, ApJS, 107, 472
Steidel, C. C., Giavalisco, M., Pettini, M., Dickinson, M., & Adelberger, K.L. 1996a, ApJ, 462, L17
Steidel, C. C., Giavalisco, M., Dickinson, M., & Adelberger, K.L. 1996b, AJ, 112, 352
Steidel, C. C., Adelberger, K. L., Dickinson, M., Giavalisco, M., Pettini, M., Kellogg, M. 1997,
ApJ, in press, preprint: astro-ph/9708125
Wemstaker, W. 1981, A&A, 97, 329
Williams, R. E., Blacker, B., Dickinson, M., Van Dyke Dixon, W., Ferguson, H. C., Fruchter, A. S.,
Giavalisco, M., Gilliland, R. L., Heyer, I., Katsanis, R., Levay, Z., Lucas, R. A., McElroy,
D. B., Petro, L., Postman, M., Adorf, H.-M., & Hook, R. N. 1996, AJ, 112, 1335
-- 17 --
Worthey, G. 1994, ApJS, 95, 107
Yee, H.K.C. 1991, PASP, 103, 396
Yee, H.K.C., & Ellingson, E. 1995 ApJ, 445, 37
Yee, H.K.C., Ellingson, E., Bechtold, J., Carlberg, R. G., & Cuillandre, J.-C. 1996 AJ, 111, 1783
Zepf, S.E., Moustakas, L.A., & Davis, M. 1997, ApJ, 474, L1
This preprint was prepared with the AAS LATEX macros v4.0.
-- 18 --
Table 1. Photometry
object
ref
z
U300,AB
B450,AB
V606,AB
I814,AB
JAB
HAB
Ks,AB
C2-05
hd2 0725 1818
hd2 2030 0287
hd2 0624 0266
C4-08
C3-02
C4-06a
C4-06b
hd4 0367 0266
hd4 2030 0851
hd2 0434 1377
hd2 1410 0259
hd2 1359 1816
C4-09
hd3 0408 0684
hd2 0705 1366
hd2 0698 1297
S
L
L
L
S
S
Sa
Sa
L
L
L
L
L
S
L
L
L
2.008
2.233
2.267
2.419
2.591
2.775
2.803
2.803
2.931
2.980
2.991
3.160
3.181
3.226
3.233
3.368
3.430
25.72(06)
25.83(05)
26.46(06)
28.09(29)
28.62(46)
25.84(05)
27.67(22)
27.33(12)
25.45(08)
> 29.37
29.74(26)
29.97(90)
> 29.56
26.81(19)
> 29.26
27.55(21)
> 29.28
23.69(03)
24.39(03)
24.47(03)
25.14(03)
24.57(03)
24.64(03)
24.59(03)
24.36(03)
24.77(04)
25.39(04)
25.54(04)
25.38(04)
25.82(04)
25.01(03)
25.88(04)
26.05(04)
26.55(05)
23.47(03)
24.28(03)
24.35(03)
25.06(03)
24.50(03)
24.54(03)
23.47(03)
24.03(03)
23.87(03)
24.74(04)
24.66(03)
24.72(03)
24.76(03)
24.12(03)
24.89(04)
25.04(03)
25.19(03)
23.17(03)
24.19(03)
24.35(04)
25.09(04)
24.31(03)
24.50(03)
23.08(03)
23.81(03)
23.54(03)
24.45(04)
24.37(03)
24.66(04)
24.35(03)
23.87(03)
24.40(04)
24.95(03)
24.81(03)
22.79(05)
24.60(24)
24.16(15)
24.85(21)
24.48(17)
24.09(13)
22.97(07)
23.46(08)
23.73(18)
23.97(15)
24.64(19)
26.02(97)
24.25(15)
23.60(09)
25.15(63)
26.08(74)
25.28(32)
22.65(07)
23.73(17)
24.70(46)
24.55(35)
24.50(32)
> 24.27
21.96(05)
23.13(09)
22.54(11)
23.97(27)
24.20(22)
24.58(56)
23.73(15)
24.01(24)
24.98(78)
24.24(21)
24.93(45)
22.23(05)
23.69(16)
24.22(30)
24.52(31)
24.17(22)
24.33(25)
21.82(04)
22.87(08)
22.39(09)
23.35(12)
24.18(20)
23.83(20)
24.47(26)
22.67(07)
23.47(17)
24.47(23)
23.62(12)
aSince object C4-06 has two very distinct components, we have treated it as two objects at the same redshift. C4-06a
is the lower object in the C4-06 panel of Fig. 2 of Steidel et al. (1996b), while C4-06b is the upper object.
Note. -- Upper limits (2σ) are indicated with a ">". Errors are indicated in parentheses: they are 1σ uncertainties
expressed as percentages of the flux.
References. -- (S) Steidel et al. (1996b); (L) Lowenthal et al. (1997)
-- 19 --
Fig. 1. -- The best-fitting 0.2Z⊙, constant SFR models. Dust content and age are free parameters.
Although their fluxes are plotted, the U300 and B450 data were not used in the fit (see text). The
model SEDs shown include the high-z Lyman suppression of the U V flux. The bottom two rows
contain objects at z > 3 -- objects for which fit quality suffers from the fact that only one filter is
present redward the Balmer break. Instantaneous burst and 0.02Z⊙ and 1.0Z⊙ models produced
fits of similar quality as the ones shown here.
-- 20 --
Fig. 2. -- Reddening and age of the best-fit
Z = 0.2Z⊙ models. The top panel shows the
constant SFR fits and the bottom one is for the
instantaneous burst model. Age is the time since
the onset of star formation. Galaxies at z < 3
are shown as solid symbols while those at z > 3,
for which fit quality is poorer, use broken ones.
Error bars come from χ2 fitting and correspond
to 90% confidence limits. Z = 0.02Z⊙ and 1.0Z⊙
models produce similar results.
Fig. 3. -- Intrinsic reddening for the 6 models
of different metallicity and star formation his-
tory. The shaded histogram corresponds to those
galaxies at z < 3, while the unshaded one is for
all 17 objects (i.e., including the lower-quality
z > 3 fits). Arrows indicate the median values
of E(B − V ) for the z < 3 galaxies. Substantial
amounts of dust are present under all scenarios.
-- 21 --
Fig. 4. -- Examples of fits with and without dust
for the Z = 0.2Z⊙ constant SFR models. The
broken line shows the best-fit models for which
age was allowed to vary freely, but E(B − V ) was
held at 0. The solid line shows the best-fit model
(as in Figure 1) for which both age and dust
were free parameters. Models with dust produce
better fits than the dust-free ones.
Fig.
5. -- Ages since the onset of the cur-
rent episodes of star formation. The shaded his-
togram corresponds to those galaxies at z < 3,
while the unshaded one is for all 17 objects. Ages
> 0.5 Gyr are grouped together in the right-most
bins (separated from the other bins by the dot-
ted lines). Most Lyman break galaxies are dom-
inated by very young stellar populations.
-- 22 --
Fig. 6. -- Examples of fits with and without
forced age for the Z = 0.2Z⊙ constant SFR mod-
els. The broken lines are the fits for which age
was fixed at 1 Gyr (while reddening was allowed
to vary freely). The solid lines are the fits where,
as in Figure 1, both age and reddening were
free parameters. Instantaneous burst models of
1 Gyr age would produce an even stronger dis-
agreement with the data, since in those models
the UV-producing massive stars would not have
continued to be replenished as in the constant
SFR models.
Fig. 7. -- The ages of dominant stellar popula-
tions of Lyman break galaxies determined by fit-
ting 0.2Z⊙ model SEDs. Error bars correspond
to 90% confidence limits. For comparison, the
dotted lines show the age of the universe and
the age of an object which formed at zf = 4.5;
both are for a Ω = 1 universe with a present-
day age fixed at t0 = 12.5 Gyr; note, however,
that the stellar population ages do not depend
on the value of H0. Objects at z > 3 suffer
from poor coverage above ∼ 4000 A and, con-
sequently, have age estimates of lower quality.
Models with Z = 0.02Z⊙ and 1.0Z⊙ produce
ages similar to these shown here. Stellar popu-
lations of the majority of z > 2 HDF galaxies
appear to have undergone recent (t < 0.2 Gyr)
episodes of star formation.
-- 23 --
Fig. 8. -- Distribution of star formation rates
obtained from the 0.2Z⊙ constant SFR fits. A
q0 = 0.5 universe was assumed. The shaded his-
togram is for the eleven z < 3 galaxies, and the
unshaded one includes all 17 objects. The me-
dian value for the z < 3 subset is indicated.
Fig. 9. -- Stellar masses from the 0.2Z⊙ fits.
Shaded histograms are for z < 3 galaxies, and
the unshaded ones include all 17 objects. Arrows
indicate median values for the z < 3 subset. A
q0 = 0.5 universe was assumed. Panel (a) shows
the distribution of stellar masses obtained from
the instantaneous burst fits. The stellar mass of
a present-day L∗ galaxy (3 × 1010h−1
100M⊙; Cowie
et al., 1995) is indicated for comparison. Panel
(b) shows the distribution of stellar masses de-
rived by multiplying the objects' star formation
rates and the ages of their stellar populations.
Under both the instantaneous burst and con-
stant SFR scenarios, stellar masses produced by
the observed episodes of star formation are gen-
erally substantially less than the stellar mass of
a present-day L∗ galaxy.
|
astro-ph/9910191 | 2 | 9910 | 1999-11-19T08:39:50 | Compton Dragged Gamma--Ray Bursts associated with Supernovae | [
"astro-ph"
] | It is proposed that the gamma-ray photons that characterize the prompt emission of Gamma-Ray Bursts are produced through the Compton drag process, caused by the interaction of a relativistic fireball with a very dense soft photon bath. If gamma-ray bursts are indeed associated with Supernovae, then the exploding star can provide enough soft photons for radiative drag to be effective. This model accounts for the basic properties of gamma-ray bursts, i.e. the overall energetics, the peak frequency of the spectrum and the fast variability, with an efficiency which can exceed 50%. In this scenario there is no need for particle acceleration in relativistic collisionless shocks. Furthermore, though Poynting flux may be important in accelerating the outflow, no magnetic field is required in the gamma-ray production. The drag also naturally limits the relativistic expansion of the fireball to Gamma < 10^4. | astro-ph | astro-ph | Submitted to ApJ Letters, Oct. 11, 1999
Preprint typeset using LATEX style emulateapj v. 04/03/99
9
9
9
1
v
o
N
9
1
2
v
1
9
1
0
1
9
9
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
COMPTON DRAGGED GAMMA -- RAY BURSTS ASSOCIATED WITH SUPERNOVAE
Davide Lazzati1, Gabriele Ghisellini
Osservatorio Astronomico di Brera, Via Bianchi 46, I -- 23807 Merate (Lc), Italy
Annalisa Celotti
Scuola Internazionale Superiore di Studi Avanzati, via Beirut 2/4, I -- 34014 Trieste, Italy
and
Martin J. Rees
Institute of Astronomy, University of Cambridge, Madingley Road, Cambridge CB3 0HA, U.K.
Submitted to ApJ Letters, Oct. 11, 1999
ABSTRACT
It is proposed that the gamma -- ray photons that characterize the prompt emission of Gamma -- Ray
Bursts are produced through the Compton drag process, caused by the interaction of a relativistic fireball
with a very dense soft photon bath. If gamma -- ray bursts are indeed associated with Supernovae, then the
exploding star can provide enough soft photons for radiative drag to be effective. This model accounts
for the basic properties of gamma -- ray bursts, i.e.
the overall energetics, the peak frequency of the
spectrum and the fast variability, with an efficiency which can exceed 50%. In this scenario there is no
need for particle acceleration in relativistic collisionless shocks. Furthermore, though Poynting flux may
be important in accelerating the outflow, no magnetic field is required in the gamma -- ray production.
The drag also naturally limits the relativistic expansion of the fireball to Γ ∼< 104.
Subject headings: gamma rays: bursts -- supernovae: general -- radiation mechanisms: non -- thermal
1.
INTRODUCTION
In the leading scenario for GRBs and afterglows, the
gamma -- ray event is produced by internal shocks in a
hyper -- relativistic inhomogeneous wind (Rees & M´esz´aros
1994) while the afterglow is produced as the fireball drives
a shock wave in the external interstellar medium (M´esz´aros
& Rees 1997). Even if there is a large consensus that
both gamma -- rays and afterglow photons are produced by
the synchrotron process, recently some doubts have been
cast on the synchrotron interpretation for the burst itself
(Liang 1997; Ghisellini & Celotti 1998; Ghisellini, Celotti
& Lazzati 1999).
The nature of the progenitor is still a matter of active
debate, as the sudden release of a huge amount of energy
in a compact region generating a fireball does not keep
trace of the way this energy has been produced. For this
reason, the study of the interactions of the fireball with
the surrounding medium seems to be the most powerful
mean to unveil the GRB progenitor. At least two models
are in competition: the merging of a binary system com-
posed of two compact objects (Eichler et al. 1989) and
the Hypernova -- Collapsar model (Woosley 1993, Paczy´nski
1998), i.e. the core collapse of a very massive star to form
a black hole.
After the discovery and the multiwavelength observa-
tions of many afterglows, circumstantial evidence has ac-
cumulated for GRB exploding in dense regions, associated
to supernova -- like phenomena.
In fact, (a) host galaxies
have been detected in many cases (Sahu et al. 1997; see
Wheeler 1999 for a review), and some of them show star-
burst activity (Djorgovski et al. 1998, Hogg & Fructher
1999); (b) large hydrogen column densities have sometimes
been detected in X -- ray afterglows (Owens et al. 1998); (c)
non -- detections of several X -- ray afterglows in the optical
band can be due to dust absorption (Paczy´nski 1998); (d)
a possible iron line feature has been detected in the X -- ray
afterglow of GRB 970508 (Piro et al. 1999, Lazzati et al.
1999) and (e) the rapid decay with time of several after-
glows can be explained by the presence of a pre -- explosion
wind from a very massive star (Chevalier & Li 1999). More
recently, the possible presence of supernova (SN) emission
in the late afterglows light curves of GRB 970228 (Galama
et al. 1999, Reichert 1999) and GRB 980326 (Bloom et al.
1999) has added support in favor of the association of some
GRBs with the final evolutionary stages of massive stars.
Although in these models the available energy is larger
than in the case of compact binary mergers, the very small
efficiency of internal shocks (see, e.g., Spada, Panaitescu &
M´esz´aros 1999) seems to be inconsistent with the fact that
more energy can be released during the burst proper than
the afterglow (Paczy´nski 1999, see also Kumar & Piran
1999).
In this letter we show that if GRBs are associated with
supernovae, Compton drag inside the relativistic wind can
produce both the expected energetics and the peak en-
ergy of the spectrum of a classical long duration GRB. In
this new scenario the efficiency is not limited by internal
shock interactions, and the successful modeling of after-
glows with external shocks is left unaffected.
The Compton drag effect has been already invoked for
GRBs by Zdziarski et al. (1991) and Shemi (1994). Cos-
mic background radiation (at high redshift), central re-
gions of globular clusters and AGNs were identified as
plausible sources of soft photons, but none of these scenarii
1Dipartimento di Fisica, Universit`a degli Studi di Milano, Via Celoria 16, I -- 20133 Milano, Italy
1
2
was able to account for all the main properties of GRBs.
However, the growing evidence of association of GRB ex-
plosions with star -- forming regions and supernovae opens
new perspectives for this scenario.
2. COMPTON DRAG IN A RELATIVISTIC WIND
We consider a relativistic (Γ ≫ 1) wind of plasma prop-
agating in a bath of photons with typical energy ǫseed. A
fraction ∼ min(1, τT) of the photons are scattered by the
inverse Compton (IC) effect to energies ǫ ∼ Γ2ǫseed, where
τT is the Thomson opacity of the wind. Due to relativistic
aberration, the scattered photons propagate in a narrow
cone forming an angle 1/Γ with the velocity vector of wind
propagation. By this process, a net amount of energy ECD
is converted from kinetic energy of the wind to a radiation
field propagating in the direction of the wind itself, where
ECD ∼ min(1, τT) V urad (Γ2 − 1). V is volume filled by
the soft photon field of energy density urad swept up by
the wind.
Let us assume that the GRB fireball, instead of being
made by a number of individual shells (see e.g. Lazzati et
al. 1999), is an unsteady (both in velocity and density)
relativistic wind, expanding from a central point. After
an initial acceleration phase, the density of the outflowing
wind decreases with radius as n(r) ∝ r−2, giving a scatter-
ing probability ∼ min[1, (r/r0)−2], where r0 is the radius
at which the scattering probability equals unity. After the
first scattering, the photons propagates in the same direc-
tion of the flow and the probability of a second scattering
is reduced by a factor ∼ Γ2.
If such a wind flows in a radiation field with energy den-
sity urad(r), the total energy transferred to the photons
when the fireball reaches a distance R is given by2:
ECD(R) = 4πΓ2"Z r0
urad(r)r2dr +Z R
r2
r0
0
0urad(r)r2dr# (1)
where for simplicity we assume that a constant Γ has been
reached (see also Section 3). The transparency radius r0
depends on the baryon loading of the fireball, which is pa-
rameterized by ηb ≡ E/(M c2), where E/M is the ratio
between the total energy and the rest mass of the fireball.
Then r0 is given by3:
r0 = 5.9 × 1013E1/2
52 η−1/2
{b,2} cm.
(2)
2.1. A simple scenario
We initially consider a simple scenario which can il-
lustrate the basic features of the Compton drag effect.
Let us assume that the GRB is triggered at a time ∆t
(of the order of a few hours) after the explosion of a
supernova (Woosley et al.
1999; Cheng & Dai 1999).
By this time, the supernova ejecta, moving with veloc-
ity βSNc, have reached a distance RSN = vSN∆t ∼ 5.4 ×
1013β{SN,−1}(∆t/5 hr) cm. Let us also imagine that the
supernova explosion is asymmetric, e.g. with no ejecta in
the polar directions. Despite this asymmetry, the ejecta
uniformly fill with radiation the entire volume within RSN.
If RSN < r0, the energy extracted by Compton drag is:
ECD =
ECD =
4πR3
3
4πr3
0
3
SN
Γ2 urad
RSN ≤ r0
(3)
Γ2 urad(cid:18)3
RSN
r0
− 2(cid:19) RSN > r0.
(4)
According to Woosley et al.
(1994), the average lumi-
nosity of a type II supernova4 during ∆t is of the order
of LSN ∼ 1044 erg s −1, with a black body emission at
a temperature TSN ∼ 105 K. It follows that in this case
{SN,5} erg cm−3 (consistent with
urad = aT 4
RSN assumed above). The efficiency ξ of Compton drag
in extracting the fireball energy is very large; from Eq. 3
we obtain:
SN ∼ 7.6 × 105T 4
ξ ≡
ECD
E
∼ 0.6 E−1
52 β3
5 h(cid:19)3
{SN,−1}(cid:18) ∆t
{SN,5} Γ2
T 4
2,
(5)
Note here that a high efficiency can be reached even for
Γ ∼ 100. Note that the drag itself can limit the maximum
speed of the expansion -- even in a wind with a very small
barion loading -- as discussed in Sect. 3. Each seed pho-
ton is boosted by ∼ 2Γ2 in frequency, yielding a spectrum
peaking at hν ∼ 2Γ2(3kTSN) ∼ 0.5Γ2
2T{SN,5} MeV.
2.2. A more realistic scenario
The previous scenario requires that the GRB explodes a
few hours after a supernova. There is however a plausible
alternative, independent of whether the massive (> 30M⊙)
star (assumed to be the progenitor of the GRB) ends up
with a supernova explosion or not, and can produce a
gamma -- ray burst even if the relativistic flow and the core
collapse of the progenitor star are simultaneous or sepa-
rated by a relatively small time interval (Woosley et al.
1999; MacFadyen, Woosley & Heger 1999).
In fact there is a somewhat general consensus (e.g. Mac-
Fadyen & Woosley 1999; Aloy et a. 1999, but see also
Khokhlov et al. 1999) that a relativistic wind can flow
in a relatively baryon free funnel created by a bow shock
following the collapse of the iron core of the star. Even
if details of this class of models are still controversial, the
formation of the funnel seems to be a general outcome. Let
us estimate its luminosity and more precisely the amount
of energy in radiation crossing the funnel walls at a time
tf after its creation. With respect to the total luminos-
ity of the star, assuming it radiates at its Eddington limit
∼ LEdd, there would be a reduction by the geometrical fac-
tor equal to the ratio of the funnel to star surfaces, which
is of the order of the funnel opening angle ϑ. However, im-
mediately after its creation, the funnel luminosity is much
larger than ϑLEdd, due to two effects which we discuss in
turn.
First the walls of the funnel contain an enhanced amount
of radiation with respect to the surface layers of the star:
the radiation once "trapped" in the interior of the star
can escape through the funnel walls, thus enhancing the
2All the calculations are made in spherical symmetry.
In case of beaming, all the quoted numbers should be considered as equivalent
isotropic values.
3Here and in the following we adopt the notation Q = 10xQx, using cgs units
4This luminosity decreases by a factor ∼ 100 for type Ibc supernovae, while the typical frequency increases by a factor of 10.
luminosity inside the funnel for a short time. Photons
produced at a distance s from the wall surface cross it at
a time tf ∼ τss/c = σns2/c, where σ is the relevant cross
section. This compares with the Kelvin time tK ∼ σnR2
⋆/c
needed for radiation to reach the star surface, yielding
s/R⋆ ∼ (tf /tK)1/2. After the time tf , the radiation pro-
duced in the layer of width ds crosses the funnel surface
carrying the energy dEf ∼ ϑτ⋆LEddds/c, the correspond-
ing luminosity being:
Lf ∼
ϑ
2
LEdd (cid:18) τ⋆R⋆
ctf (cid:19)1/2
.
(6)
For tf = 100 s and a 10M⊙ star with R⋆ ∼ 1013 cm
(τ⋆ ≈ 108), this effect can enhance the funnel luminosity
by ∼ 104.
Let us now consider a second plausible enhancing fac-
tor. If the funnel has been produced by the propagation
of a bow -- shock in the star, the matter in front of the ad-
vancing front is compressed, with a pressure increase of
M2, where M is the Mach number of the shock in the
star. This (optically thick) gas then flows along the sides
of the funnel and relaxes adiabatically to the pressure of
the external matter (its original pressure). The result is
that the funnel is surrounded by a sheath (cocoon) with
density lower than that of the unshocked stellar material
by a factor M3/2 (a polytropic index of 4/3 has been used
in the adiabatic cooling). The diffusion of photons through
this rarefied gas into the funnel is then even faster, result-
ing in a further increase of the luminosity by M3/4 ∼ 200,
where a shock speed βsf c = 0.1c (MacFadyen & Woosley
1999) and a sound speed βsc = 10−4c have been assumed.
By taking into account both effects, the funnel luminos-
ity corresponds to:
Lf ∼ LEdd
∼ 1045ϑ−1
M⋆
10M⊙
erg s−1,
ϑ
2 (cid:18) τ⋆R⋆
ctf (cid:19)1/2(cid:18) βsf
βs (cid:19)3/4
(7)
which leads to an energy loss for Compton drag LCD ∼
Γ2Lf ∼ 1049ϑ−1Γ2
2(M⋆/10M⊙), to be compared with the
observed luminosity hLGRBi ∼ 1049πϑ2
−1 erg s−1. Here
the average luminosity is considered over the entire burst
duration: for single pulses, we should take into account an
extra factor Γ2 in Eq. 7 due to the Doppler contraction of
the observed time.
The typical radiation temperature associated with this
luminosity, assuming a black body spectrum, is enhanced
with respect to the temperature of the star surface by
[Lf /(ϑLEdd)]1/4 ∼ (τ⋆R⋆)1/8(ctf )−1/8(βsf /βs)3/16. Adopt-
ing the numerical values used above, the enhancement is
of the order of 50, corresponding to a funnel temperature
Tf ∼ 2 × 105 K (for a surface temperature of the star of
∼5000 K). This value is similar to the one estimated in the
simple scenario of the previous subsection and thus leads
to similar Compton frequencies.
3. PROPERTIES OF THE OBSERVED BURSTS
If the wind is homogeneous the spectrum of the scat-
tered photons resembles that of the incident photons, i.e.
a broad black -- body continuum peaked at a temperature
Tdrag ∼ 2 Γ2T . While the observed characteristic photon
energy would be therefore ǫ ∼ 0.5Γ2
2 T5(1 + z)−1 MeV,
3
in good agreement with the observed distribution of peak
energies of BATSE GRBs (assuming again Γ = 100, see
below), the spectrum would not reproduce the observed
smoothly broken power -- law shape (Band et al. 1993).
The assumptions of a perfectly homogeneous wind and of
an isothermal radiation field are however very crude, and
one might reasonably expect that different regions of the
wind are characterized by different values of Γ and dif-
ferent soft field temperatures.
If we assume, e.g., that
the temperature of the soft photon field varies with ra-
dius according to a power -- law T (r) ∝ r−δ, the time in-
tegrated spectrum will have a high energy power -- law tail
F (ν) ∝ ν− 3−3δ
In addition, the bulk Lorentz factor of
the flow is likely to vary on a timescale much shorter than
the integration time required to obtain a spectrum with
the BATSE data (∼ 1 s), and hence the analysed spec-
tra are the superposition of drag spectra by many differ-
ent Lorentz factors. A third effect adding power to the
high energy tails of the spectrum is the reflection of up --
scattered photons in the pre -- supernova wind. This pho-
tons are scattered again by the fireball and can reach ener-
gies of ∼ 0.5Γ MeV ∼ 50Γ2 MeV. The computation of the
actual spectrum resulting from all these effects depends
from many assumptions and is beyond the scope of this
work.
.
δ
The effects described above, which can increase the fun-
nel luminosity over the Eddington limit, take place in non --
stationary conditions. At the wind onset, it is likely that
the temperature gradient in the walls of the funnel is large,
but this is soon erased due to the high luminosity of the
walls. This causes both the total flux and the character-
istic frequency of the soft photons to decrease, and hence
a hard -- to -- soft trend is expected. Moreover, it has been
shown by Liang & Kargatis (1996) that the peak frequency
of the spectrum in a single pulse at time t is strongly re-
lated to the flux of the pulse integrated from the beginning
of the pulse to the time t. In our scenario, this behaviour
can be easily accounted for if we consider a shell slowed
down by the drag itself: the Lorentz factor (and hence the
peak frequency of the spectrum) at a time t is related to
the energy lost by the shell, i.e. to the integral of the flux
from the beginning of the pulse to the time t.
The observed minimum variability time -- scale is related
to the typical size of the region containing the dense seed
photon field, which corresponds to either R⋆ or RSN de-
pending on which of the two scenarios described above
applies. The relevant light crossing time -- divided by the
time compression factor -- is thus
tvar ∼
R
cΓ2 ∼ 3 × 10−2 R13 Γ−2
2
s.
(8)
Longer time -- scales are instead expected if the relativistic
wind is smooth and continuous.
Another interesting feature of this scenario is the pos-
sibility that the bulk Lorentz factor of the wind is self --
consistently limited by the drag itself. The pressure of
the soft photons starts braking the fireball in competi-
tion with the pressure of internal photons. The limiting
Lorentz factor is hence reached when the internal pressure
p′
fb ∝ (T0/Γ)4 is balanced by the pressure of the exter-
nal photons as observed in the fireball comoving frame
p′ ∝ Γ2 T 4
SN (1 + τT)−1, where τT is the scattering optical
4
depth of the wind. This gives:
Γlim ∼ 2 × 104 T −1/2
{SN,5} E1/4
52 R−5/8
{0,7} η−1/8
{b,5},
(9)
where R0 is the radius at which the fireball is released.
Equation 9 reduces to Γlim ∼ 104(T0,11/TSN,5)2/3 if the
fireball becomes transparent before reaching the coasting
phase. With such high Γ the Compton drag would be max-
imally efficient, causing the fireball to immediately decel-
erate until its Γ reaches the value given by LCD = Lf Γ2,
implying:
Γ = (cid:18) Lkin
Lf (cid:19)1/2
∼ 300 (cid:18) L{kin,50}
L{f,45} (cid:19)1/2
.
(10)
These limits are in general smaller than the maximum Γ
set by the baryon load only, but still in agreement with
values recently inferred for GRB 990123 (Sari & Piran
1999). In addition, it is likely that the external parts of
the relativistic wind, which are in closer connection with
the funnel walls, are dragged more efficiently then the cen-
tral ones, since at the beginning the soft photons coming
from the walls can penetrate only a small fraction of the
funnel before being up -- scattered by relativistic electrons.
This may result in a polar structured wind, with higher
Lorentz factors along the symmetry axis, gradually de-
creasing as the polar angle increases.
4. DISCUSSION
A crucial requirement of our model is the association of
GRBs with the final evolutionary stages of very massive
stars, as these provide the large amount of seed photons
emitted at distances ∼ 1013 cm from the central trigger,
which are neeeded for the Compton drag to be efficient.
The efficiency of conversion of bulk kinetic energy of
the flow into gamma -- ray photons is large, solving the ob-
servational challenge of gamma -- ray emission being more
energetic than the afterglow one (Paczy´nski 1999). Fur-
thermore in this scenario there is no requirement for ei-
ther efficient acceleration in collisionless shocks or the
presence/generation of an intense (equipartition) magnetic
field, although Poynting flux may still be important in ac-
celerating the outflow (being more efficient than neutrino
reconversion into pairs).
We have investigated the main properties of a GRB
produced by Compton drag in a relativistic wind in a
very general case. A moderately beamed burst (ϑ ∼< 10◦,
Woosley et al. 1999) can be thus produced and, without
any fine tuning of the parameters, the basic features of
classic GRBs are accounted for.
In particular, the peak energy of the burst emission sim-
ply reflects the temperature of the supernova seed photons,
up -- scattered by the square of the bulk Lorentz factor.
The simplest hypothesis predicts a quasi -- thermal spec-
trum, however it is easy to imagine an effective multi --
temperature distribution which would depend on uncon-
strained quantities such as the variation of the spectrum
of the SN photons with radius and the degree of inhomo-
geneity of the wind.
Although in this scenario there is no requirement for
internal shocks to set up, they can of course occur, con-
tributing a small fraction of the observed gamma -- ray flux.
On the other hand, the wind is expected to escape from
the funnel of the star with still highly relativistic motion,
so that an external shock can be driven in the interstellar
medium and produce an afterglow, similar to the scenario
already studied by several authors. It is likely that this af-
terglow would develop in a non -- uniform density medium,
due to the presence of the massive star wind occurring
before the supernova explosion (Chevalier & Li 1999).
We thank Andy Fabian, Francesco Haardt, Piero Madau
and Giorgio Matt for many stimulating discussions. DL
thanks the Institute of Astronomy for the kind hospitality
during the preparation of this work. The Cariplo Founda-
tion (DL) and Italian MURST (AC) are acknowledged for
financial support.
REFERENCES
Aloy, M. A., Mueller, E., Ibanez, J. M., Marti, J. M. & MacFadyen,
A., 1999, ApJ submitted (astro-ph/9911098)
Band, D. et al., 1993, ApJ, 413, 281
Bloom, J. S., et al., 1999, Nature, 401, 453
Cheng, K. S. & Dai, Z. G., 1999, ApJ submitted (astro-ph/9908248)
Chevalier, R. A. & Li, Z., 1999, ApJ, 520, L29
Djorgovski, S. G., Kulkarni, S. R., Bloom, J. S., Goodrich, R., Frail,
D. A., Piro, L., Palazzi, E., 1998, ApJ, 508, L17
Eichler, D., Livio, M., Piran, T. & Schramm, D. M., 1989, Nature,
340, 126 & Taylor, G. B., 1997, Nature, 389, 261
Fryer, C. L., Woosley, S. E., Hartmann, D. H., 1999, ApJ in press
(astro-ph/9904122)
Galama, T. J., et al., 1999, ApJ submitted (astro-ph/9907264)
Ghisellini, G. & Celotti, A., 1998, ApJ, 511, L93
Ghisellini, G., Celotti, A. & Lazzati, D., 1999, MNRAS submitted
Hogg, D.W., & Fruchter, A.S., 1999, ApJ, 520, 54
Khokhlov, A. M., Hoflich, P., Oran, S. E., Wheeler, J. C. & Wang,
L., 1999, ApJ submitted (astro-ph/9904419)
Kumar, P. & Piran, T., 1999, ApJ submitted (astro-ph/9909014)
Lazzati, D., Campana, S. & Ghisellini, G., 1999a, MNRAS, 302, L31
Lazzati, D., Ghisellini, D., & Celotti, A., 1999b, MNRAS309 L13
Liang, E. P. & Kargatis, V., 1996, Nature, 381, 49
Liang, E. P., 1997, ApJ, 491, L15
Owens, A. et al., 1998, A&A, 339, L37
MacFadyen, A & Woosley, S. E., 1999, ApJ submitted (astro-
ph/9810274)
MacFadyen, A, Woosley, S. E. & Heger, A., 1999, ApJ submitted
(astro-ph/9910034)
M´esz´aros, P. & Rees, M. J., 1997, ApJ, 476, 232
Piro, L. et al., 1999, ApJ, 514, L73
Paczy´nski, B., 1986, ApJ, 308, L43
Paczy´nski, B., 1998, ApJ, 494, L45
Paczy´nski, B., 1999, to appear in: "The Largest Explosions Since
the Big Bang: Supernovae and Gamma Ray Bursts" (astro-
ph/9909048)
Rees, M. J. & M´esz´aros, P., 1994, ApJ, 430, L93
Reichart, D. E., 1999, ApJ, 512, L111
Sari, R. & Piran, T., 1999, ApJ, 517, L109
Sahu, K., Livio, M., Petro, L. & Macchetto, D. F., 1997, IAU Circ.
6606
Shemi, A., 1994, MNRAS, 269, 1112
Spada, M., Panaitescu, A. & M´esz´aros, P., 1999, ApJ submitted
(astro-ph/9908097)
Wheeler, J. C., 1999, to appear in: "The Largest Explosions Since
the Big Bang: Supernovae and Gamma Ray Bursts" (astro-
ph/9909096)
Woosley, S. E., 1993, ApJ, 405, 273
Woosley, S. E., Eastman, R. G., Weaver, T. A. & Pinto, P. A., 1994,
ApJ, 429, 300
Woosley, S. E., MacFadyen, A. I. & Heger, A., 1999, to appear
in: "The Largest Explosions Since the Big Bang: Supernovae and
Gamma Ray Bursts" (astro-ph/9909034)
Zdziarski, A. A., Svensson, R. & Paczy´nski, B., 1991, ApJ, 366, 343
|
astro-ph/9507027 | 2 | 9507 | 1995-07-10T10:57:50 | Neutrino pair synchrotron radiation from relativistic electrons in strong magnetic fields | [
"astro-ph"
] | The emissivity for the neutrino pair synchrotron radiation in strong magnetic fields has been calculated both analytically and numerically for high densities and moderate temperatures, as can be found in neutron stars. Under these conditions, the electrons are relativistic and degenerate. We give here our results in terms of an universal function of a single variable. For two different regimes of the electron gas we present a simplified calculation and compare our results to those of Kaminker et al. Agreement is found for the classical region, where many Landau levels contribute to the emissivity , but some differences arise in the quantum regime. One finds that the emissivity for neutrino pair synchrotron radiation is competitive, and can be dominant, with other neutrino processes for magnetic fields of the order $B \sim 10^{14} - 10^{15} G $.This indicates the relevance of this process for some astrophysical scenarios, such as neutron stars and supernovae. | astro-ph | astro-ph | Neutrino pair synchrotron radiation from relativistic electrons in
strong magnetic fields
A. Vidaurre1, A. P´erez2, H. Sivak 4, J. Bernab´eu2,3, and J. Ma. Ib´anez2
1 Departamento de F´ısica Aplicada, Universidad Polit´ecnica de Valencia, Spain
2 Departamento de F´ısica Te´orica, Universidad de Valencia
3 IFIC, Centro Mixto Univ. Valencia-CSIC, 46100 Burjassot (Valencia), Spain
4 D.A.R.C., Observatoire de Paris-Meudon, 92190 Meudon, France
Received
;
accepted
5
9
9
1
l
u
J
0
1
2
v
7
2
0
7
0
5
9
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
-- 2 --
ABSTRACT
The emissivity for the neutrino pair synchrotron radiation in strong magnetic
fields has been calculated both analytically and numerically for high densities
and moderate temperatures, as can be found in neutron stars. Under these
conditions, the electrons are relativistic and degenerate. We give here our
results in terms of an universal function of a single variable. For two different
regimes of the electron gas we present a simplified calculation and compare our
results to those of Kaminker et al. Agreement is found for the classical region,
where many Landau levels contribute to the emissivity , but some differences
arise in the quantum regime. One finds that the emissivity for neutrino pair
synchrotron radiation is competitive, and can be dominant, with other neutrino
processes for magnetic fields of the order B ∼ 1014 − 1015G. This indicates the
relevance of this process for some astrophysical scenarios, such as neutron stars
and supernovae.
Subject headings : Stars : Magnetic Fields -- Stars: Neutron.
-- 3 --
1.
INTRODUCTION
Estimates of the magnetic field strength at the surface of neutron stars are obtained
from several different scenarios: theoretical models of pulsar emission (Ruderman 1972), the
accretion flow in binary X-ray sources (Ghosh and Lamb 1978) and observation of features
in the spectra of pulsating X-ray sources which have been interpreted as cyclotron lines
(Trumper et al. 1978, Wheaton et al. 1979, Gruber et al. 1980, Mihara et al. 1990). For
a recent review of all these topics the interested reader is addressed to the book by Michel
(Michel 1991). In a sample of more than 300 pulsars the range of values of the surface
magnetic field strength runs into the interval: 10.36 ≤ logB (Gauss) ≤ 13.33 (Manchester
and Taylor 1981).
Very recently, several authors (Duncan and Thomson 1992, Thomson and Duncan
1993, Bisnovatyi-Kogan and Mosheenko 1992, Bisnovatyi-Kogan 1993) have proposed two
different physical mechanisms leading to an amplification of some initial magnetic field in a
collapsing star. Fields as strong as B ∼ 1014 − 1016G, or even more, might be generated in
new-born neutron stars.
According to (Bisnovatyi-Kogan and Mosheenko 1992, Bisnovatyi-Kogan 1993), a
mirror-asymmetric magnetic field distribution might arise in a rapidly and differentially
rotating proto-neutron star having, originally, both a toroidal and a poloidal component.
The field amplification due to differential rotation leads to the formation of an additional
toroidal field from the poloidal one by twisting of the field lines. After the first 20 seconds
of the life of a new-born neutron star (basically, its Kelvin-Helmholtz epoch) the induced
toroidal magnetic field could be as huge as B ∼ 1015 − 1017G.
In a second scenario (Duncan and Thomson 1992, Thomson and Duncan 1993), a
dynamo action in a differentially rotating and convective young neutron star is responsible
-- 4 --
for the strengthening of some initial dipole field up to values of B ∼ 1012 − 3 × 1013G if
the convective episodes arose during the main-sequence stage or to B ∼ 1014 − 1015G if the
dipole field is generated after collapse.
In presence of a strong magnetic field the so-called neutrino-pair synchrotron radiation
process becomes allowed :
[e−]
~B→ [e−] + ν + ¯ν
(1)
This reaction has been studied by a few groups (Landstreet 1967, Canuto et al.
1970, Yakovlev and Tschaepe 1981, Vidaurre 1990, Kaminker et al. 1991, Kaminker et
al. 1992, Kaminker and Yakovlev 1993) for different regimes of the electron plasma. The
calculation of the corresponding emissivity, both analytically or numerically, is far from
obvious. In fact, this calculation appears as a multiple integral and summation over the
variables and quantum numbers involving the wave functions of the initial and final electron
and the corresponding statistical weights. The wave function integrals lead to Laguerre or
Bessel functions with a complicated behaviour, so that approximations in order to simplify
the expressions are sometimes delicate. This in fact has lead to errors in the past literature.
In a recent paper, Kaminker et al. (Kaminker et al. 1991) have studied the neutrino
emissivity of the above process (1) for moderately high magnetic fields B ∼ 1012 − 1014G
and high densities, in the degenerate-relativistic regime for the electrons. They claim that,
within these conditions, the emissivity is independent of the electron density. So far, to our
knowledge, there are no numerical tests which confirm these results.
Given all these circumstances, together with the interest of the problem, both from
the theoretical and astrophysical point of view, we have considered useful to reexamine
the existing results for the above regime. Therefore, we have performed a numerical study
of this process at high densities ρYe ≥ 107g/cm3, where Ye is the electron fraction per
baryon and ρ the matter density in c.g.s. units, and moderate temperatures T < 109K, for
-- 5 --
values of the magnetic field strength B ≤ 1016G. Under these conditions, the electrons are
relativistic and degenerate. We have found a result which is in agreement with the one of
Kaminker et al. when the electron gas is in the classical regime. We also obtain agreement
for the corresponding analytical expressions, which we present in a simpler way than these
authors. For the quantum regime, however, we give analytical formulae which show the
correct dependence on B for large magnetic fields, in contrast to those given by Kaminker
et al. This fact would be particularly important if some of the magnetic field amplification
mechanisms described above were physically realizable in nature.
The influence of the neutrino-pair synchrotron radiation in presence of strong magnetic
fields merits to be examined in different astrophysical scenarios, such as the delayed
mechanism of type II Supernovae, neutrino cooling of proto-neutron stars during the Kelvin
epoch or the secular cooling of the neutron star. Furthermore, the neutrinos originated
from the combined effect of one of the proposed mechanisms to enhance the magnetic field
and the neutrino-pair synchrotron radiation process could be envisaged as a signature of
the mechanism itself.
This paper is organized as follows. In section 2 we discuss the calculation of the
emissivity for the synchrotron process and present our main results. These results are
illustrated with more detail in section 3 and 4 for two different regimes of the electron gas.
We end in section 5 with some conclusions and remarks.
2. CALCULATION OF THE EMISSIVITY
The emissivity for the process (1) can be written as
-- 6 --
εν = G2eB
3(2π)6 P∞n=1Pn−1
n′=0 R +∞
−∞
dpz R~q2≤ω2 d3q ω
× A f (E) [1 − f (E′)]
(2)
where G is the Fermi coupling constant. The initial (final) electron, of mass m, has
an energy E = qm2 + p2
⊥
+ p2
z (E′ = qm2 + p′2
⊥
+ p′2
z ) characterized by the Landau
quantum number n (n′) and momentum pz (p′z) along the B-direction. We have introduced
= 2eBn and p′2
⊥
p2
= 2eBn′, which correspond to the classical transverse momenta of the
⊥
electron. In the above equation, q is the four-momentum transfer and ω = E − E′ the
energy which is carried away by the neutrino pair. f (E) is the Fermi-Dirac distribution
function : f (E) = hexp( E−µ
T ) + 1i−1
integration region over ~q is restricted by :
, where µ is the electron chemical potential. The
q2 = ω2 − ~q2 ≥ 0.
The expression for A in the equation of the emissivity can be found in (Vidaurre
1990, Kaminker et al. 1992). In the relativistic limit, the relevant expression for A is :
A =
(C 2
V + C 2
A)
EE′
n[−2(EE′ − pzp′z)2
+ (EE′ − pzp′z)(p2
+ p2
t [(EE′ − pzp′z) −
t − 2q2
⊥
1
2
t )# Ψ(u)
q2
⊥
) −
2
(p2
t − q2
⊥
(q2
⊥ − p2
)(cid:21) Φ(u)(cid:27)
where
Ψ =
n′!
n!
use−u[
u
n′
(Ls+1
n′−1)2 +
n
u
(Ls−1
n′ )2]
(3)
(4)
-- 7 --
Φ =
n′!
n!
use−u[
n
n′
(Ls
n′−1)2 + (Ls
n′)2]
(5)
n1 are Laguerre functions with argument u = q2
Ln2
2eB (q⊥ is the component of ~q orthogonal
to the magnetic field) and s = n − n′. CV (CA) is the effective vector (axial) coupling of the
neutrino pair to the electron current, coming from both the Fierz reordered charged current
⊥
(for electron neutrinos) and neutral current (for all neutrino species) weak interactions.
In Eq. (4) we have dropped a term proportional to (C 2
V − C 2
A), which disappears in the
extreme relativistic limit, and the interference term proportional to CV CA, which does not
contribute to the integrated emissivity in Eq. (2). The argument goes as follows. The
corresponding integrand, for CV CA, is odd under the simultaneous change of sign of the
longitudinal momenta of the initial and final electrons, so that a symmetric integration of
both pz and qz in Eq.(2) cancels this asymmetric term. We have defined p2
t = p2
⊥
+ p′2
⊥
. If
one takes for the electroweak mixing angle sin2θW = 0.23, then C 2
V + C 2
A = 1.6748.
One has the relationship
ω2 − q2
z = 2m2 + p2
t − 2(EE′ − pzp′z) ≈ p2
t − 2(EE′ − pzp′z)
(6)
where the latter approximation corresponds to considering relativistic electrons.
By substituting in Eq. (4) we obtain
A =
(C 2
V + C 2
A)
2EE′
(ω2 − q2
z − q2
⊥
)[p2
t (Ψ − Φ) − (ω2 − q2
z − q2
⊥
)Ψ]
(7)
As mentioned above, for the range of temperatures and densities we are interested in,
the electrons are degenerate. The product of distribution functions appearing in (2) will
then restrict the energies to E, E′ ∼ µ. Moreover, one can write the following identity :
-- 8 --
f (E) [1 − f (E′)] = B(ω) [f (E′) − f (E)]
(8)
where B(ω) = [exp(ω/T ) − 1]−1 is a Bose-Einstein distribution function with zero chemical
potential. From the last equation, it is apparent that the energy difference ω will be
restricted to a few times T and will be much lower than the relevant values of E and
E′, if the electrons are degenerate. Similarly, one can easily estimate that qz and q⊥ will
contribute as ∼ T , whereas pt, pz and p′z contribute as ∼ µ. This allows us to neglect the
last term in Eq. (7). In this case one gets
A =
(C 2
V + C 2
A)
2EE′
p2
t (ω2 − q2
z − q2
⊥
)Θ
(9)
with Θ = Ψ − Φ. We have tested numerically that the complete expression Eq.(4) or
Eq.(7) gives approximately the same result as Eq.(9) for the physical conditions we are
considering here.
We can also perform the following approximations, in agreement with the above
discussion :
n + n′ =
ω =
=
p2
t
2eB
2seB + p2
E + E′
(E + E′)2 + ω2 − 2(p2
z − p′2
4eB
2seB + 2pzqz
z
z + p′2
z )
≈
E + E′
(E + E′)2 − 4p2
z
4eB
≈
(10)
Due to the Pauli principle, the quantity f (E)[1 − f (E′)] will be nonzero only when E
and E′ are within a narrow interval around µ of width ω ∼ T . With this in mind, we have
replaced all the slowly varying functions in (2) by their value around E, E′ ∼ µ.
The latter equation, together with the phase space restriction (3), requires that qz and
q⊥ must lie inside the elliptical domain (Kaminker et al. 1991)
-- 9 --
+
q2
⊥
p2
µ2 (qz − χz)2 ≤ χ2
⊥
⊥
(11)
and χ⊥ = seB
p⊥
.
where χz = seBpz
p2
⊥
We have numerically calculated the emissivity of the synchrotron process (1) for values
of the electron density ρYe ≥ 107g/cm3, moderate temperatures T < 109K, and magnetic
fields B ≤ 1016G. The results, for this range of values, can be expressed in a compact way
as :
εν = 1.47 1014 B2
13 T 5
9 f (x) erg/cm3/s
(12)
where B13 is the magnetic field in units of 1013G, T9 is the temperature in units of 109K,
and x = µT
eB is a dimensionless variable. This result is useful in order to perform analytical
calculations. We have plotted in Fig. 1 the function f (x). As can be seen from this figure,
f (x) first increases as x grows, and is almost constant for large values of x. This behaviour
corresponds to different physical regimes of the electron plasma, and will be explained in
the next sections.
3. Large x Regime
Let us discuss the behaviour of the emissivity corresponding to large values of the
parameter x defined above. This situation corresponds, for example, to sufficiently high
electron densities for a fixed magnetic field and temperature. The number of Landau levels
2eB . On the other hand, the maximum of s = n− n′
which are involved in Eq. (2) is nmax = µ2
can be estimated from Eq. (10) as smax ∼ µT
eB = x. We then have nmax >> smax >> 1
-- 10 --
for µ
T >> 1. According to this idea, we have used the following approximations for the
Laguerre polynomials in Eq. (5) :
s−1 + J 2
s+1
Ψ → J 2
Φ → 2J 2
s
(13)
with Js(a) a Bessel function and a = q2(n + n′)u. One can prove that a ≤ s always. The
approximation shown by Eq. (13) can be used if n >> 1 and s << n. Because it is time
saving and more suitable numerically than Eq. (5), we made use of it in our numerical
computation of the emissivity, whenever large values of n (and s << n) were encountered.
By changing the sum over n and n′ to a sum over n′ and s, and integrating the angle of ~q
around the ~B direction, one arrives to the following expression :
εν = G2eB
3(2π)5 (C 2
× (1 − p2
V + C 2
µ2 )(ω2 − q2
A) P∞n′=0P∞s=1 R +∞
−∞
dpz Rq2≤ω2 dqzdq⊥q⊥ ω
) B(ω) [f (E′) − f (E)] Θ(a)
z − q2
⊥
z
(14)
The argument of Θ can be written as a = p⊥q⊥/(eB), with p⊥ = qµ2 − p2
z, and ω = seB+pzqz
.
µ
Further approximations can be made in the above equation for degenerate electrons, if one
considers the distribution functions f (E) and f (E′) as step functions in the energy. Within
this assumption, the sum over n′ can be done explicitly. One gets
∞
Xn′=0
[f (E′) − f (E)] =
ω
2eB
(E + E′) ≃
µω
eB
(15)
In deriving Eq.(15), we have used the fact that s is lower than nmax. By inserting the
latter expression into Eq. (14) we obtain
-- 11 --
εν = G2µ
3(2π)5 (C 2
V + C 2
× (1 − p2
A) P∞s=1 R +∞
dpz Rq2≤ω2 dqzdq⊥q⊥ ω2
−∞
z − q2
µ2 )(ω2 − q2
⊥
) B(ω)Θ(a)
z
(16)
Next we make use of recurrence relationships for the Bessel functions and obtain the
formula :
Θ(a) = 2(J′s)2 + 2(s2/a2 − 1)J 2
s
(17)
For the large values of s involved here, one can use the following approximation to the
Bessel functions (Gradshteyn and Ryzhik 1980):
Js(a) ≈
1
πs 2(s − a)
3a
K1/3(z)
(18)
where z = [2(s−a)]3/2
3√x
approximated as :
and K1/3 is the modified Bessel function, which can be further
K1/3(z) ≈ r π
2z
exp (−z)
In this way one obtains, after some algebra,
Θ(a) ≈
2
π
s−4/3(2z)1/3 exp (−2z)
(19)
(20)
As can be seen from Fig. 2, the latter equation provides a reasonable approximation
to Eq.(17). In this figure, we have plotted Θ(a) as obtained from Eq.(20) (dotted line) and
from Eq.(17) (solid line) for s = 200. Another important feature is that only values of the
-- 12 --
argument a close to s will contribute. This can be understood from Eq.(20), due to the
exponential behaviour, which effectively limits z. In fact, if we define σ as the 'width' of
the exponential, one has (1 − a/s) < ( σ
important values of qz and q⊥ are restricted to
s )2/3. By substituting into Eq.(11) one obtains that
σ
s
q⊥
χ⊥
)2/3
) ≤ (
(1 −
qz − χz ≤ ∆(q⊥) ≤
√2
µχz
pz
(
σ
s
)1/3
(21)
We have defined ∆(q⊥) = µχ⊥
. Thus qz is restricted to a narrow interval of
p⊥ r1 − q2
χ2
⊥
⊥
width ∆(q⊥) around χz. This allows us to perform the integral over qz approximately. To
the first order in ∆(q⊥) one can write
Z χz+∆(q⊥)
χz−∆(q⊥)
dqz −→ 2∆(q⊥)
(22)
with the replacement qz → χz in the integrand of Eq. (16). This means that ω2 will be
. The integral over q⊥ is then immediate. Since the total number of level
differences s is large, one can substitute the sum over s by an integral over the continuous
replaced by χ2
z + χ2
⊥
variable t = s/x. Thus by changing Ps −→ xR ∞0 dt and performing the remaining integrals
one finally obtains
εν =
G2
π6 (C 2
V + C 2
A)ζ(5)(eB)2T 5 = 1.16 1015 B13 T 5
9 erg/cm3/s
(23)
This equation implies that the emissivity does not depend on the electron density in
this regime, in agreement with the result previously found in Ref. (Kaminker et al. 1991).
In fact, our Eq. (23) is close to the one obtained in this reference. It is also in good
-- 13 --
agreement with the values of f (x) obtained numerically (and plotted in Fig. 1), as can be
seen by comparing Eq.(23) with Eq.(12).
A numerical fit which reproduces the behaviour of f (x) for x > 2 to better than 4% is
given by
f (x) = −5.0224 − 8.1289x + 9.2892x2
1.0293 + 2.0605x + 1.0727x2
(24)
4. Low x Regime
We now address the question of whether an increase of the magnetic field will always
give a larger neutrino emission. In Fig. 3 we present the corresponding emissivity (in cgs
units and logarithmic scale) for a fixed temperature T9 = 1 and four values of the electron
density ρYe = 109, 1011, 1013, 1014g/cm3, as a function of B13 (B13 ranging from unity up
to 103), as obtained numerically. For these high magnetic fields, the number of populated
Landau levels (nmax, defined above) can be of order unity, and we enter into the quantum
regime. We have used, in these cases, the expression of A as given by Eqs. (4) and (5)
directly, instead of making the approximations shown in section 3.
As can be seen from Fig. 3, for a given density and temperature, the emissivity first
increases and, after reaching a maximum value, will fall to zero for large values of the
magnetic field. This can be understood since the number of possible n → n′ transitions
decreases as eB
µT ≫ 1. In fact, a rough analytic expression in this region can be obtained by
putting s = 1 in Eq. (14) and ω = eB
µ . Therefore one has :
Ψ ≃ 1
Φ ≃ 0
(25)
-- 14 --
The integrals in Eq. (14) can be done analytically and one obtains the following
expression :
εν = 6.6 1012 B2
13 T 5
9 x−5e−1/x erg/cm3/s
(26)
We have verified that the latter expression gives values which are in agreement with
our numerical results around the maximum of the emissivity. This is shown in Fig. 4,
where we have compared our results for ρYe = 1011g/cm3 (solid line) with the prediction
of Eq. (26) (dotted line). We have also plotted (dashed line) the corresponding analytical
approximation given by (Kaminker et al. 1991) for this case. As can be seen from this
figure, Eq. (26) provides a reasonable approximation to the emissivity, which works better
than the formula of Kaminker et al.
5. Comparison with other processes
In order to investigate the relevance of the process studied here for neutron star cooling,
we have made the comparison with the emissivities corresponding to other neutrino processes
which are competitive with the synchrotron emission. We have considered pair production
e+e− −→ ν ¯ν ,plasmon decay Γ −→ ν ¯ν, bremmstrahlung e−(Z, A) −→ e−(Z, A)ν ¯ν and
photoproduction γe− −→ e−ν ¯ν. For these processes, we have made the assumption that
they do not vary significantly with the magnetic field. The numerical fit to these emissivities
have been taken from Munakata et al. 1985. Although these rather simple formulae are
not the most up-to-date available fits to the above processes, they serve to our purpose as
a first approximation (see, for example, Itoh et al. 1989, for a more elaborated fit). The
bremmstrahlung process is taken from Maxwell 1979. A more complete calculation, as in
-- 15 --
Itoh et al. (Itoh et al. 1989), gives the same order of magnitude in the region where this
process dominates.
In Fig. 5 we present the result of comparing all these processes for a temperature
T = 108K as the product ρYe varies from zero up to 1012g/cm3. The bremmstrahlung
energy emission was calculated assuming that the dominant nucleus is 56F e. The
synchrotron emissivity is plotted (solid lines) for two values of the magnetic field : B13 = 10
and B13 = 100. Other relevant processes are : bremmstrahlung (dashed-dotted line),
plasma (long dash), and photoneutrino emission (short dash). As can be seen from this
figure, the synchrotron emission is competitive with the above processes within the range
ρYe ∼ 109 − 1012. Moreover, as pointed out by Pethick and Thorsson (Pethick and Thorsson
1993), band-structure effects can suppress bremmstrahlung by a factor of 10 or more for
temperatures less than about 109K. In this case, the synchrotron emission would be the
dominant process in the above electron density range, if the magnetic field reaches values
of the order ∼ 1014G.
For higher temperatures, the synchrotron emission corresponding to a given value of
B 'switches on' at lower electron densities, as can be inferred from Eq. (26). However,
other processes have a faster increase with temperature and, therefore, the dominance
of synchrotron emission reduces to a narrow interval of densities, although it effectively
competes for high densities. This is shown in Fig. 6, where we have made the above
comparison for a temperature T = 109K (pair emission is represented by the dotted line).
6. CONCLUSIONS
We have performed numerical calculations of the synchrotron emissivity from
relativistic degenerate electrons. This calculations allow us to present the results in terms
-- 16 --
of an universal function f (x) which can be used in astrophysical codes. For two different
regimes of the electron gas we have derived analytical formulae, in a simpler way than
previous references. These formulae have been tested, and we have found a reasonable
agreement with our numerical calculations. We also have compared our results to the
analytical formulae derived recently by Kaminker et al. Agreement is found for the classical
region, where many Landau levels contribute to the emissivity , but some differences arise
in the quantum regime, for the values of the magnetic field recently suggested in new-born
neutron stars.
We have shown that neutrino-pair synchrotron radiation for moderately high magnetic
fields B ≥ 1014G is an efficient cooling mechanism for temperatures not larger than 109K,
and can compete effectively (or even dominate) with other processes.
We claim that the influence of the neutrino-pair synchrotron radiation in presence of
strong magnetic fields (∼ 1015G) merits to be examined in different astrophysical scenarios,
such as the delayed mechanism of type II Supernovae, neutrino cooling of proto-neutron
stars during the Kelvin epoch or the secular cooling of the neutron star. Furthermore, these
neutrinos originated from the combined effect of the dynamo action with the neutrino-pair
synchrotron radiation process could be envisaged as a signature of this mechanism. This
will be the subject of future investigations.
Acknowledgments This work has been partially supported by the Spanish DGICYT
(grant PB91-0648) and CICYT (grant AEN 93-0234). Calculations were carried out in a
VAX 6000/410 at the Instituto de F´ısica Corpuscular and in a IBM 30-9021 VF at the
Centre de Inform`atica de la Universitat de Val`encia. We are grateful to J.A. Miralles for
useful comments.
-- 17 --
REFERENCES
Bisnovatyi-Kogan, G.S., Mosheenko, I., 1992, Sov. Astron., 36, 285.
Bisnovatyi-Kogan, G.S., 1993, Astron. Astrophys. Transactions, 3, 287.
Canuto, V., Chiu, H.Y., Chou, C.K., 1970, Phys. Rev., D2, 281.
Duncan, R.C., Thompson, C., 1992, Ap. J., 392, L9.
Ghosh, P., and Lamb, F.K., 1978, Ap. J. Letters, 223, L83.
Gradshteyn, I.S., Ryzhik, I.M., "Table of Integrals, Series and Products". Academic Press,
1980.
Gruber, D.E., Matteson, J.L., Nolan, P.L., Knight, F.K., Baity, W.A., Rotschild, R.E.,
Peterson, L.E., Hoffman, J.A., Scheepmaker, A., Wheaton, W.A., Primini, F.A.,
Levine, A.M., Lewin, W.H.G., 1980, Ap. J. Letters, 240, L127.
Itoh, N., Adachi, T., Nakagawa, M., Kohyama, Y., Munakata, H., 1989, Ap. J., 339, 354.
see also Haft, M., Raffelt, G., 1993, Max-Planck Institut fur Astrophysik preprint.
Schinder, J., Schramm, D.N., Wiita, P.J., Margolis, S.H., Tubbs, D.L. 1987, Ap. J.,
313, 531.
Kaminker, A.D., Levenfish, K.P., Yakovlev, D.G., 1991, Sov. Astron. Lett., 17, 450.
Kaminker, A.D., Levenfish, K.P., Yakovlev, D.G., Amsterdamski, P., Haensel, P., 1992,
Phys. Rev. D, 46, 3256.
Kaminker, A.D., Yakovlev, D.G., 1993, JETP, 76, 229.
Landstreet, J.D., 1967, Phys. Rev., 153, 1372.
-- 18 --
Manchester, R.N., Taylor, J.H., 1981, Astron. J., 86, 1953.
Maxwell, O.V., 1979, Ap. J., 231, 201.
Michel, F.C., Theory of Neutron Star Magnetospheres, University of Chicago Press (1991).
Mihara, T., Makishima, K., Ohashi, T., Sakao, T., Tashiro, M., Nagase, F., Tanaka, Y.,
Kitamoto, S., Miyamoto, S., Deeter, J.E., Boynton, P.E., 1990, Nature, 346, 250.
Munakata, H., Kohyama, Y., Itoh, N., 1985, Ap. J., 296, 197
Pethick, C.J., Thorsson, V., Nordita preprint 93/72 A/S/N.
Ruderman, M.A., 1972, Ann. Rev. Astron. Astrophys., 10, 427.
Thompson, C., Duncan, R.C., 1993, Ap. J., 408, 194.
Trumper, J., Pietsch, W., Reppin, C., and Voges, W., 1978, Ap. J. Letters, 219, L105.
Vidaurre, A., Ph.D. Thesis, Universidad de Valencia (1990).
Wheaton, W.A., Doty, J.P., Primini, F.A., Cooke, B.A., Dobson, C.A., Goldman, A., Hecht,
M., Hoffman, J.A., Home, S.K., Scheepmaker, A., Tsiang, E.Y., Lewin, W.H.G.,
Matteson, J.L., Gruber, D.E., Baity, W.A., Rotschild, R., Knight, F.K., Nolan, P.,
Peterson, L.E., 1979, Nature, 272, 240.
D.G. Yakovlev and R. Tschaepe, 1981, Astron. Nachr. 302, 167.
This manuscript was prepared with the AAS LATEX macros v3.0.
-- 19 --
Figure Captions
Figure 1.- The function f (x) appearing in Eq. (12). See the text for the definition of
the dimensionless variable x and comments about its behaviour.
Figure 2.- The function Θ(a) as obtained from the approximation Eq.(20) (dotted
line) compared to Eq.(17) (solid line) for s = 200.
Figure 3.- Neutrino synchrotron emissivity as a function of the magnetic field B13 for
different values of the electron density. The temperature is the same (T = 109K) in all
cases.
Figure 4.- Comparison of our numerical results for ρYe = 1011g/cm3 (solid line) with
the analytical approximation Eq. (26) (dotted line). We have also plotted (dashed line) the
approximation given by (Kaminker et al. 1991) for this case.
Figure 5.- Competition of synchrotron neutrino emission (solid lines) with other
processes, as a function of the electron density, for a temperature T = 108K. Two
different values of the magnetic field (B13 = 10 and B13 = 100 have been considered. The
bremmstrahlung emissivity (dash-dotted line) has been calculated for 56F e. Plasma process
is represented by long-dashed line, and photoneutrino by short-dashed line.
Figure 6.- Same as figure 5 for T = 109K. Dotted line corresponds to pair emission.
|
astro-ph/9704110 | 1 | 9704 | 1997-04-12T03:18:13 | A New Model for Soft Gamma-Ray Repeaters | [
"astro-ph"
] | We consider a model in which the soft gamma-ray repeaters (SGRs) result from young, magnetized strange stars with superconducting cores. As such a strange star spins down, the quantized vortex lines move outward and drag the magnetic field tubes because of the strong coupling between them. Since the terminations of the tubes interact with the stellar crust, the dragged tubes can produce sufficient tension to crack the crust. Part of the broken platelet will be dragged into the quark core, which is only $10^4$ cm from the surface, leading to the deconfinement of crustal matter into strange quark matter and thus the release of energy. We will show that the burst energy, duration, time interval and spectrum for our model are in agreement with the observational results. The persistent X-ray emission from the SGRs can be well explained by our model. | astro-ph | astro-ph |
A New Model for Soft Gamma-Ray Repeaters
K.S. Cheng1
and Z.G. Dai2
1Department of Physics, University of Hong Kong, Hong Kong
2Department of Astronomy, Nanjing University, Nanjing 210093, China
We consider a model in which the soft gamma-ray repeaters (SGRs) result from young, magnetized
strange stars with superconducting cores. As such a strange star spins down, the quantized vortex
lines move outward and drag the magnetic field tubes because of the strong coupling between them.
Since the terminations of the tubes interact with the stellar crust, the dragged tubes can produce
sufficient tension to crack the crust. Part of the broken platelet will be dragged into the quark core,
which is only 104 cm from the surface, leading to the deconfinement of crustal matter into strange
quark matter and thus the release of energy. We will show that the burst energy, duration, time
interval and spectrum for our model are in agreement with the observational results. The persistent
X-ray emission from the SGRs can be well explained by our model.
PACS numbers: 98.70.Rz., 12.38.Mh, 26.60.+c, 97.60Jd
The soft gamma-ray repeaters (SGRs) are a small,
enigmatic class of gamma-ray transient sources. There
are three known SGRs which are charaterized by short
rise times (as short as 5 ms) and duration (∼ 50-150 ms,
FWHM, some less than 16 ms), spectra with character-
istic energies of ∼ 30-50 keV and little or no evolution,
and stochastic burst repetition within a timescale of ∼ 1
month [1]. SGR 0525−66, the source of the 1979 March
5 event, appears to be associated with the N49 supernova
remnant (SNR) in the Large Magellanic Cloud and hence
is apparently the most distant known SGR source at ∼ 55
kpc from Earth [2]. The second burster, SGR 1806−20,
which produced ∼ 110 observed bursts during a 7-year
span [3] and recently became active again [4], appears
to be coincident with the SNR G10.0−0.3 [5], confirming
an earlier suggestion [6]. Thus, this source is at a dis-
tance of ∼ 15 kpc. The third burster, SGR 1900+14, is
associated with SNR G42.8+0.6 [7], and its age is ∼ 104
yrs and its distance from Earth is ∼ 7 kpc. Accept-
ing these SGR-SNR associations, the burst peak lumi-
nosities can be estimated to be a few orders higher than
the standard Eddington value for a star with the mass
of ∼ 1M⊙. For example, SGR 1806−20 has produced
events that are ∼ 104 times the Eddington luminosity [8].
In addition to short bursts of both hard X-rays and soft
γ-rays, the persistent X-ray emission was also detected
from SGRs [5,7,9]. The luminosities of the persistent
X-ray sources are ∼ 7 × 1035 ergs s−1 for SGR 0525−66,
∼ 3×1035 ergs s−1 for SGR 1806−20, and ∼ 1035 ergs s−1
for SGR 1900+14. These observations show that the re-
peaters may be young, magnetized neutron stars which
power the surrounding luminous plerionic nebulae.
There may be three classes of models for explaining the
energy source of SGRs. In the first class of models, SGRs
were thought to result from accretion of neutron stars (for
a brief review see [10]). Since the highly super-Eddington
flux requires the accretion inflow and radiation outflow
to be channeled in different directions so that it makes
any accretion model very difficult. Second, it was sug-
gested [11] that glitches of normal pulsars are an energy
source of SGRs. However, the current models for pulsar
glitches [12,13] seem to give glitching intervals and dura-
tions much larger than those of SGRs. Moreover, no SGR
bursts have so far been detected from the Crab pulsar.
These two facts may disfavor the glitch model for SGRs.
Third, it was argued [14] that SGRs are magnetars, a
kind of neutron stars with superstrong magnetic fields of
≥ 5 × 1014 G. Although the motivations for this model
(e.g., rapid spin-down to 8 s period in 104 years) sound
attractive, there may be several unsettled issues [10], e.g.,
(i) a power output from such a strong magnetic field may
be inconsistent with the plerion energy range; (ii) in such
a strong field the radiation output is highly anisotropic
but the observed shape seems to be angle independent.
In this letter we suggest that SGRs result from young,
magnetized strange stars with superconducting cores.
The structure of strange stars has been studied [15].
Strange stars near 1.4M⊙ have thin crusts with thickness
of ∼ 104 cm and mass of ∼ 10−5M⊙. Some arguments
may be unfavorable to the existence of strange stars.
First, most important, the relaxation behavior of glitches
of pulsars which seem to be isolated neutron stars with
masses of ∼ 1.4M⊙ is well described by the neutron su-
perfluid vortex creep theory [12], but the current strange
1
star models scarely explain the observed pulsar glitches
[16]. This may also mean that at least pulsars in which
glitches occur must be neutron stars, not strange stars.
Second, the conversion of a neutron star to a strange star
requires the formation of a strange matter seed at a den-
sity (6-9 times the nuclear matter density) much larger
than the central denisty of the 1.4M⊙ star with a rather
stiff equation of state [17]. This shows that strange stars
are not easy to be produced in the universe.
However, it was argued [18] that when neutron stars in
low-mass X-ray binaries accrete sufficient mass, they may
convert to strange stars. This mechanism was further
suggested as a possible origin of cosmological gamma-ray
bursts. In this Letter we suggest that strange stars may
also be formed during the core collapse of massive stars or
during the accretion phase of newly born neutron stars.
The birth rate of strange stars due to these processes
must be low. This is beacuse (i) if the rate were high,
the number of resulting strange stars would be too high
to explain the observed glitch phenomena; (ii) although
the current type II supervova models believe that neutron
stars can be produced during the core collapse of massive
stars in some controversial mass range and the evolution
of more massive stars can result in the formation of black
holes, these models have neither given the upper limit of
the masses of massive stars which evolve to neutron stars
nor the lower limit of the masses of massive stars which
evolve to black holes. We conjecture that massive stars
in a narrow mass range may finally evolve to strange
stars. There are two cases for this evolution: (i) dur-
ing the core collapse the nucleon matter directly convert
into strange matter [19], in which case the shock wave
for the supernova can obtain more energy; (ii) the cen-
tral density of a newly born neutron star may reach the
deconfinement density due to hypercritical accretion in a
supernova circumstance [20] and then the whole neutron
star may undergo a phase transition to a strange star.
After the birth, a strange star must start to cool due
to neutrino emission. As a neutron star does, the strange
star core may become superconducting when its interior
temperature is below the critical temperature. Using a
relativistic treatment of BCS theory, Bailin & Love [21]
suggested that strange matter forms superconducting.
They showed that the pairing of quarks is most likely
to occur in both ud and ss channels. The pairing state of
the former is likely in s-wave and that of the latter is in
p-wave. The superconducting transition temperature is
about 400 keV. Therefore, a strange star with age older
than 103 years after its supernova birth should have a
core temperature lower than the normal-superconducting
temperature [22]. The quark superconductor is likely to
be marginally type II with zero temperature critical field
Bc ∼ 1016-1017 G [21,23] which sensitively depends on
the interactions between quarks.
On the other hand, the existence of quantized vortex
lines in the rotating core of a strange star is unclear.
Since different superconducting species inside a rotating
strange star try to set up different values of London fields
in order to compensate for the effect of rotation. Using
the Ginzburg-Landau formulism, Chau [23] showed that
instead of setting a global London field vortex bundles
carrying localized magnetic fields can be formed. The
typical field inside the vortex core is about 1016-1017 G
(the accurate value depends on strong interaction param-
eters). Using the similar idea proposed for the interaction
between the proton fluxoids and magnetic neutron vor-
tics in the core of a neutron star [24], he argued that the
vortex bundle and the flux tubes can interpin to each
other by interaction of their core magnetic fields. He
estimated that the pinning energy per intersection is
Ep ∼ 690N 1/2
flux MeV ,
(1)
where Nflux is the number of flux quantum in a flux tube.
Such strong binding between vortex lines and flux tubes
implies that when the vortex lines moving outward due
to spinning-down of the star will induce the decay of the
magnetic field [23]. One of the important consequences
of this coupling effect will be discussed next text.
We now propose a plate tectonic model for strange
stars which is, in principle, similar to that proposed by
Ruderman [24] for neutron stars. As described in last
subsection, there might exist two different types of quan-
tized flux tubes in the core of a strange star. The first
type of flux tubes are formed when the stellar magnetic
field penetrates through the superconducting core. The
second type of flux tubes (vortex lines) result from the
requirement of minimizing the rotating energy of the
core superfluid. When the star spin down due to mag-
netic dipole radiation, the vortex lines move outward
and pull the flux tubes with them.
Inductive currents
do not strongly oppose this flux tube motion because
of current screening by the almost perfectly diamagnetic
superconducting quarks. However, the terminations of
flux tubes are anchored in the base of highly conduct-
2
ing crystalline stellar crust. When the stellar spin-down
timescale τs = Ω/2 Ω is shorter than the typical ohmic
diffuse timescale,
τD ∼
σA
4πc2
∼ 3 × 104σ21R2
6 yrs ,
(2)
where σ is the conductivity and R6 is the radius in units
of 106 cm. The motion of flux tubes is limited by their
terminations in the crust unless the resulting pull on
the crust by these flux tubes exceeds the crustal yield
strength, namely,
BBc
8π
sin θ > µθs
l
R
,
(3)
where B is the stellar magnetic field, θ is the angle bew-
teen the stellar magnetic moment and the flux tubes,
µ is the shear modulus, θs is the shear angle, and l is
the crustal thickness. Substituting the typical values of
strange star parameters into equation (3), we obtain
sin θ ∼ θ > θc ≡ 3 × 10−6B−1
c,17B−1
12 θs,−3µ27l4R−1
6
rad ,
(4)
where Bc,17 is in 1017 G, B in 1012 G, θs,−3 in 10−3, µ27 in
1027 dyn cm−2, and l4 in 104 cm. When θ > θc, the stel-
lar crust will crack and θ will be reduced by an amount
δθ ∼ min(θ, ∆l/R) (∆l is the displacement of the crustal
plate). In the case of neutron stars, Ruderman [25] esti-
mated that ∆l ∼ 2×102 cm for the Crab and Vela pulsars.
For a strange star with a much thinner crust than that
of a neutron star, we expect that l > ∆l > 2 × 102 cm,
which implies δθ ∼ θ. Since the flux tubes move outward
with the same speed as the vortex lines which is given by
v ∼
R
τs
= 3 × 10−6R6τ −1
s,4 cm s−1 ,
(5)
where τs,4 is in 104 yrs, the time interval between two
successive cracking events is estimated to be
τint ∼
Rδθ
v
∼ 106 B−1
c,17B−1
12 θs,−3µ27l4R−1
6 τs,4 s .
(6)
This value is consistent with the typical interval timescale
of SGRs.
When the crust cracks, a small platelet could be
dragged from the crust into the strange matter core which
is only 104 cm from the surface. In the following we make
an estimate of the timescale for the platelet motion. The
force pulling the craking platelet horizontally by the flux
tubes is
Fp =
BBc
8π
θAp ,
(7)
where Ap is the area of the platelet. Thus, the timescale
opening a hole with area ∼ Ap is approximated by
τdrag =(cid:18)2l
ApMcr
4πR2
1
Fp(cid:19)1/2
∼ 80(cid:18) Mcr,−5
θs,−3µ27R6(cid:19)1/2
ms ,
(8)
where Mcr,−5 is the total mass of the crust in units of
10−5M⊙. The durations of the SGRs are expected to be
of the same order as this timescale. As normal matter
falls into the core continuously, the baryons will deconfine
into quarks. Because each baryon can release ∼ (20 − 30)
MeV (the accurate value is dependent upon the QCD
parameters), which are a sum of gravitational energy and
deconfinement energy, the total amount of energy release
is estimated as
∆E ∼ 3 × 1042Mcr,−5(Ap/l2)l2
4R−2
6
ergs .
(9)
where Mcr,−5 is the total mass of the crust in units
of 10−5M⊙. At least half of this amount will be car-
ried away by thermal photons with the typical energy
kT ∼ 30 MeV. These thermal photons will be released
continuously in a timescale of ∼ τdrag. In the presence of
a strong magnetic field, the thermal photons will convert
into electron-positron pairs when
Eγ
2mc2
B
Bq
sin Φ ∼
1
15
,
(10)
where Eγ is the photon energy, Bq = m2c3/¯he = 4.4 ×
1013 G and Φ is the angle between the photon propaga-
tion direction and the direction of the magnetic field [27].
The energies of the resulting pairs will be lost via syn-
chrotron radiation. The characteristic synchrotron en-
ergy is given by
E(1)
syn ∼
3
2
γ2
e ¯h
eB
mc
sin Φ ∼ 1.5 MeV ,
(11)
where γe is the Lorentz factor of the electron (∼ 30).
The first generation of synchrotron photons will be con-
verted into the secondary pairs because the optical depth
of photon-photon pair production is much large than 1.
The characteristic synchrotron energy of the secondary
pairs is given by
3
2mc2!2
2 E(1)
syn
¯h(cid:18) eB
mc(cid:19) ∼ 37B12 keV .
(12)
E(2)
syn ∼
3
Foundation of China.
-- -- -- -- -- -- -- -- -- -- -- --
[1] J.P. Norris, P. Hertz, K.S. Wood, and C. Kouve-
liotou, Astrophys. J. 366, 240 (1991).
[2] W.D. Evans et al., Astrophys. J. 237, L7 (1980);
T.L. Cline et al., ibid. 255, L45 (1982).
[3] J.G. Laros et al., Astrophys. J. 320, L111 (1987).
[4] C. Kouveliotou et al., Nature (London) 368, 125
(1994).
[5] T. Murakami et al., Nature (London) 368, 127
(1994).
[6] S.R. Kulkarni, and D.A. Frail, Nature (London) 365,
33 (1993).
[7] G. Vasisht, S.R. Kulkarni, D.A. Frail, and J. Greiner,
Astrophys. J. 431, L35 (1994).
[8] E.E. Fenimore, J.G. Laros, and A. Ulmer, Astrophys.
J. 432, 742 (1994).
[9] R.E. Rothschild, S.R. Kulkarni, and R.E. Ligenfelter,
Nature (London) 368, 432 (1994).
[10] E.P. Liang, Astrophys. Space Sci. 231, 69 (1995).
[11] F. Melia, and M. Fatuzzo, Astrophys. J. 438, 904
(1995); M. Fatuzzo, and F. Melia, ibid. 464, 316
(1996).
[12] D. Pines, and M.A. Alpar, Nature (London) 316,
27 (1985).
[13] B. Link, and R.I. Epstein, Astrophys. J. 457, 844
[14] R.C. Duncan, and C. Thompson, Astrophys. J.
392, L9 (1992); C. Thompson, and R.C. Duncan,
Mon. Not. R. Astron. Soc. 275, 255 (1995).
[15] C. Alcock, E. Farhi, and A. Olinto, Astrophys. J.
310, 261 (1986); P. Haensel, J.L. Zdunik, and R.
Schaeffer, Astron. Astrophys. 160, 121 (1986).
[16] M.A. Alpar, Phys. Rev. Lett. 58, 2152 (1987).
[17] G. Baym, in Neutron Stars: Theory and Observa-
tions eds. J. Ventura and D. Pines (Kluwer Aca-
demic Publishers), 21 (1991).
Lx ∼
ξ∆E
τint
∼ 3 × 1036 ξMcr,−5(Ap/l2)l4R−1
6
×Bc,17B12θ−1
s,−3µ−1
27 τ −1
s,4 ergs s−1 ,
(1996).
(13)
Since the optical depth of photon-electron scattering near
the star is also much larger than 1, a radiation-pair fire-
ball in thermal equilibrium will have an initial temper-
ature of the same order as E(2)
syn, and will expand adi-
abatically as a fluid. During the expansion the radia-
tion energy is converted into a bulk kinetic energy of
the outflow. The fireball will cool with T = T0(R0/R),
and the relativistic Lorentz factor Γ of the bulk motion
is Γ = T0/T = R/R0 [28]. Therefore, when the optical
depth of the fireball is one, an observer at infinity will see
a blueshifted spectrum with the typical energy of ∼ E(2)
syn
due to the relativistic bulk motion of the fireball.
Finally, we want to discuss an astrophysical implica-
tion of our model. The persistent X-ray emission from
the SGRs was detected. If the sources are normal neu-
tron stars with typical magnetic fields of ∼ 1012 G, it is
obvious that the persistent X-ray luminosities from the
SGRs may not be explained by the surface blackbody
radiation. This is because calculations for the cooling of
neutron stars [29] predict that after (0.5 −1) ×104 yrs the
bolometric luminosities will be at least two orders smaller
than the persistent X-ray ones from the SGRs. Recently
Usov [30] suggested that if the sources of the SGRs are
magnetars the persistent X-ray emission may be the ther-
mal radiation of these stars which is enhanced by a factor
of 10 or more due to the effect of ultrastrong magnetic
fields. We can also explain the observed persistent X-
ray emission by using our model. After each cracking
event, at least half of the resulting thermal energy from
the deconfinement of normal matter into strange quark
matter will be absorbed by the stellar core and thus the
surface radiation luminosity at thermal equilibrium may
be estimated to be
where ξ is a parameter which accounts for both the ratio
of the absorbed thermal energy to the released total en-
ergy during a cracking event and the ratio of the surface
blackbody radiation energy to the absorbed thermal en-
ergy. We expect that this parameter is of the order of 0.5.
Taking B−1
12 θs,−3µ27 ∼ 3 to account for τint ∼ 3×106
s, we have Lx ∼ 5 × 1035 erg s−1. This estimated lumi-
nosity seems to agree with the observed ones from the
SGRs.
c,17B−1
This work was supported in part by a RGC grant of
Hong Kong and in part by the National Natural Science
4
[18] K.S. Cheng, and Z.G. Dai, Phys. Rev. Lett. 77,
1210 (1996).
[19] N.A. Gentile et al., Astrophys. J. 414, 701 (1993);
Z.G. Dai, Q.H. Peng, and T. Lu, ibid. 440, 815
(1995).
[20] R.A. Chevalier, Astrophys. J. 346, 847 (1989); G.E.
Brown, ibid. 440, 270 (1995).
[21] B. Bailin, and A. Love, Phys. Rep. 107, 325 (1984).
[22] O.G. Benvenuto, H. Vucetich, and J.E. Horvath,
Nucl. Phys. B (Proc. Suppl.) B24, 160 (1991).
[23] H.F. Chau, Astrophys. J., in press (1997).
[24] H.F. Chau, K.S. Cheng, and K.Y. Ding, Astrophys.
J. 399, 213 (1992).
[25] M.A. Ruderman, Astrophys. J. 366, 261 (1991).
[26] M.A. Ruderman, preprint (1996).
[27] M.A. Ruderman, and P.G. Sutherland, Astrophys.
J. 196, 51 (1975).
[28] J. Goodman, Astrophys. J. 308, L47 (1986); A.
Shemi, and T. Piran, ibid. 365, L55 (1990).
[29] K. Nomoto, and S. Tsuruta, Astrophys. J. 312, 711
(1987).
[30] V.V. Usov, Astron. Astrophys. 317, L87 (1997).
5
|
0803.1893 | 1 | 0803 | 2008-03-13T03:58:16 | Physical Properties of Tidal Features in Interacting Disk Galaxies | [
"astro-ph"
] | We explore tidal interactions of a galactic disk with Toomre parameter Q ~ 2 embedded in rigid halo/bulge with a point mass companion moving in a prescribed parabolic orbit. Tidal interactions produce well-defined spiral arms and extended tidal features such as bridge and tail that are all transient, but distinct in nature. In the extended disks, strong tidal force is able to lock the perturbed epicycle phases of the near-side particles to the perturber, shaping them into a tidal bridge that corotates with the perturber. A tidal tail develops at the opposite side as strongly-perturbed, near-side particles overtake mildly-perturbed, far-side particles. The tail is essentially a narrow material arm with a roughly logarithmic shape, dissolving with time because of large velocity dispersions. Inside the disks where tidal force is relatively weak, on the other hand, a two-armed logarithmic spiral pattern emerges due to the kinematic alignment of perturbed particle orbits. While self-gravity makes the spiral arms a bit stronger, the arms never become fully self-gravitating, wind up progressively with time, and decay after the peak almost exponentially in a time scale of ~ 1 Gyr. The arm pattern speed varying with both radius and time converges to Omega-kappa/2 at late time, suggesting that the pattern speed of tidally-driven arms may depend on radius in real galaxies. We present the parametric dependences of various properties of tidal features on the tidal strength, and discuss our findings in application to tidal spiral arms in grand-design spiral galaxies. (Abridged) | astro-ph | astro-ph |
Accepted for publication in ApJ
Physical Properties of Tidal Features in Interacting Disk Galaxies
Sang Hoon Oh, Woong-Tae Kim, Hyung Mok Lee
Department of Physics and Astronomy, FPRD, Seoul National University, Seoul 151-742,
Republic of Korea
[email protected], [email protected], [email protected]
and
Jongsoo Kim
Korea Astronomy and Space Science Institute, Daejeon 305-348, Republic of Korea
[email protected]
ABSTRACT
We investigate the physical properties of tidal structures in a disk galaxy
created by gravitational interactions with a companion using numerical N-body
simulations. We consider a simple galaxy model consisting of a rigid halo/bulge
and an infinitesimally-thin stellar disk with Toomre parameter Q ≈ 2. A per-
turbing companion is treated as a point mass moving on a prograde parabolic
orbit, with varying mass and pericenter distance. Tidal interactions produce
well-defined spiral arms and extended tidal features such as bridge and tail that
are all transient, but distinct in nature. In the extended disks, strong tidal force
is able to lock the perturbed epicycle phases of the near-side particles to the
perturber, shaping them into a tidal bridge that corotates with the perturber.
A tidal tail develops at the opposite side as strongly-perturbed, near-side parti-
cles overtake mildly-perturbed, far-side particles. The tail is essentially a narrow
material arm with a roughly logarithmic shape, dissolving with time because of
large velocity dispersions. Inside the disks where tidal force is relatively weak,
on the other hand, a two-armed logarithmic spiral pattern emerges due to the
kinematic alignment of perturbed particle orbits. While self-gravity makes the
spiral arms a bit stronger, the arms never become fully self-gravitating, wind up
progressively with time, and decay after the peak almost exponentially in a time
– 2 –
scale of ∼ 1 Gyr. The arm pattern speed varying with both radius and time con-
verges to Ω− κ/2 at late time, suggesting that the pattern speed of tidally-driven
arms may depend on radius in real galaxies. Here, Ω and κ denote the angular
and epicycle frequencies, respectively. We present the parametric dependences of
various properties of tidal features on the tidal strength, and discuss our findings
in application to tidal spiral arms in grand-design spiral galaxies.
Subject headings: galaxies: spiral — galaxies: structure — galaxies: interactions
— galaxies: evolution — methods: numerical
1.
Introduction
Spiral arms are the most outstanding morphological features in disk galaxies. They not
only provide information on the dynamical states of the background stellar disks but also af-
fect galactic evolution by triggering large-scale star formation in the gaseous component (see
Elmegreen 1995; Bertin & Lin 1996; see also McKee & Ostriker 2007 and references therein).
Regarding the nature of the spiral structure, two pictures have been proposed. In one pic-
ture, the arms are viewed as quasi-stationary density waves that live long, rotating almost
rigidly around the galactic centers (Lin & Shu 1964, 1966). Nonaxisymmetric instability of
the stellar disks may grow to form a sort of self-sustained standing density waves in stellar
disks (Bertin et al. 1989a,b; Bertin & Lin 1996). In the other picture, the arms are transient
features driven, for example, by gravitational interaction with a companion galaxy (Toomre
1969; Toomre & Toomre 1972) or by swing amplification of leading waves (Julian & Toomre
1966; Goldreich & Lynden-Bell 1965; Toomre 1981). In this case, spiral features are short
lived, lasting only for several rotation periods (∼ 1 Gyr) and perhaps requiring intermittent
external forcing (e.g., Sellwood & Carlberg 1984).
Observations indicate that the probability to have grand-design arms is much higher for
galaxies in binaries or groups than in the field (Kormendy & Norman 1979; Elmegreen & Elmegreen
1982, 1987). This suggests that regardless of their nature, some of grand-design spiral arms
are clearly excited by nearby galaxies through tidal interactions. Prototypical examples in-
clude M51 and M81 that possess, respectively, companion galaxies NGC5195 and M82 within
50 kpc in distance.
In a pioneering work, Toomre & Toomre (1972) used non-interacting
test-particle simulations to demonstrate that gravitational interaction of a disk galaxy with
its companion generates features such as tidal bridge and tail which are commonly seen in
extended disks of interacting galaxies. Inclusion of self-gravity tends to enhance spiral struc-
ture in the disks (e.g., Hernquist 1990). Grand-design spiral arms can be produced even by
a low-mass perturber if the interaction involves a very close passage, indicating that tidal
– 3 –
arms may be more frequent than previously thought (Byrd & Howard 1992).
Since Toomre & Toomre (1972), there have been many numerical studies of tidal in-
including the detailed modelings of the arm morphologies in the
teractions of galaxies,
M51/NGC5195 system (e.g., Hernquist 1990; Howard & Byrd 1990; Barnes 1998; Salo & Laurikainen
2000a,b; Durrell et al. 2003) and in the NGC 7753/7752 system (Salo & Laurikainen 1993),
formation of tidal tails and tidal dwarf galaxies therein (e.g., Barnes & Hernquist 1992, 1996;
Elmegreen et al. 1993; Wetzstein et al. 2007), and formation of bars at the central parts of
galaxies (e.g., Noguchi 1987; Gerin et al. 1990). In particular, by using numerical simula-
tions with self-gravitating stars and gas, Salo & Laurikainen (2000a) argued that a bound
multiple-passage orbit of NGC5195 better reproduces the observed kinematics of an extended
H I tail of M51 (Rots et al. 1990), whereas the radial velocity data of the planetary nebu-
lae associated with the tidal structures favor an unbound single-passage orbit (Durrell et al.
2003). Although these authors considered primarily the outer, extended tidal features for
comparison, the spiral arms in the main disk may more tightly constrain the orbital pa-
rameters of the M51 system. This is because the structure and kinematics of tidal tails
depend rather sensitively on the observationally uncertain parameters such as a halo mass
distribution (Dubinski et al. 1996).
While the aforementioned work has improved our understanding of the tidally-induced
morphological changes of galaxies chiefly in extended disks, these studies did not focus on
spiral structures in the main disks that are more relevant to large-scale star formation.
Stellar spiral arms are certainly one of the main agents that greatly influence dynamical
evolution of the interstellar gas in disk galaxies. Since the turbulent and thermal sound
speeds of the gas are small, the gas responds very strongly to the gravitational potential per-
turbations imposed by the stellar spiral arms, readily forming galactic spiral shocks near the
potential minima (Roberts 1969; Shu, Milione, & Roberts 1973; Woodward 1975). In optical
images, these shocks appear as narrow dust lanes that represent regions where giant molec-
ular clouds and new stars form (e.g., Elmegreen & Elmegreen 1983; Vogel et al 1988; Rand
1993; Elmegreen 1994; Shetty et al. 2007). Nonaxisymmetric gravitational instability occur-
ring inside the dust lanes (e.g., Balbus 1988; Kim & Ostriker 2002, 2006; Shetty & Ostriker
2006) is most likely responsible for observed arm substructures including gaseous spurs (or
feathers) that jut perpendicularly from the arms (e.g., Scoville et al. 2001; Willner et al.
2004; Calzetti et al. 2005; La Vigne et al. 2006).
The strength of spiral shocks and their susceptibility to gravitational instability are
strongly affected by the physical properties of stellar arms such as amplitude, pitch angle,
pattern speed, etc, yet it is quite difficult to characterize them observationally. While the
arm pitch angle can be determined relatively straightforwardly if the inclination of a galaxy
– 4 –
is known, it is challenging to measure the pattern speed unambiguously. For instance, the
Tremaine & Weinberg (1984) method that has been applied to the CO data of several grand-
design spiral galaxies (e.g., Zimmer et al. 2004) assumes, among others, that the pattern
speed is independent of radius and that the molecular gas satisfies the mass conservation
equation. In the case of M51, however, a recent study by Shetty et al. (2007) shows that
observed density and velocity profiles across the disk do not obey the continuity equation in
any frame rotating at a fixed angular speed. Also, the strength of spiral arms determined
from K-band observations is prone to contamination by red supergiants in the arm regions
(e.g., Rix & Rieke 1993; Patsis et al. 2001). Given these observational uncertainties, it is
desirable to run numerical simulations to pin down the arm parameters and thus access the
connection between stellar arms and large-scale star formation in gaseous arms.
Motivated by these considerations, we in this paper use numerical N-body simulations
to explore in detail the properties of stellar spiral arms resulting from tidal interactions.
Since the parameter space is large, we consider an idealized galaxy model in which an
infinitesimally-thin, two-dimensional, exponential stellar disk is immersed in a combined
potential due to rigid halo and bulge. Self-gravitating particles comprising the disk respond
to a point-mass perturber that passes on a prograde parabolic orbit in the same plane as
the disk rotation; we vary the mass and pericenter distance of the perturber to study the
situations with various tidal strength. The particles are not allowed to move out of the disk
plane, and the effect of gas is ignored. A fully self-consistent treatment of the problem,
using active halo and bulge as well as three-dimensional disks consisting of both stars and
gas, will be studied in subsequent papers. Similar simulations have been carried out by
Elmegreen et al. (1991) who showed that a cold stellar disk (with zero velocity dispersion)
turns into a transient ocular shape if tidal perturbations are strong. In this work, we instead
consider a disk galaxy with realistic velocity dispersions. Our main objectives are to study
the quantitative changes in the properties of spiral arms as the tidal strength varies, and
also to clarify the development and physical nature of tidal features known as bridge and
tail in extended disks.
This paper is organized as follows. Section 2 describes the galaxy model and the orbital
parameters of tidal interactions as well as the numerical method we use. In §3, we focus on
the transient extended-disk structures produced by strong tidal perturbations and show that
the tidal bridge and tail form by distinct mechanisms. In §4, we measure the properties (pitch
angle, strength, and pattern speed) of the spiral arms and present their temporal and radial
variations. Finally, we summarize our results and discuss their astronomical implications in
§5.
– 5 –
2. Model and Numerical Method
2.1. The Model Galaxy
In this paper we investigate the generation of tidal features and their properties in a
disk galaxy via gravitational interactions with a point-mass perturber. The disk galaxy
consists of three components: a spherical “dark” halo, a spherical bulge, and an exponential
stellar disk; we do not consider a gaseous disk in the current work. The halo and bulge
accounting for the inner linearly-rising part and the outer nearly-flat part of the rotation
curve are represented by fixed gravitational potentials for simplicity. This will ignore the
potential consequences on the disk through the tidal deformation of the halo and bulge.1 On
the other hand, this inert halo and bulge enables a large number of particles for the stellar
disk. In order to maximize the particle number for stars near the disk midplane, we impose
a constraint that the disk remains infinitesimally thin during its whole evolution.
Appendix A describes the specific model we employ for each component of the galaxy:
a truncated logarithmic potential for the dark matter halo,2 a Plummer potential for the
spherical bulge, and an exponential density profile for the disk. The total galaxy mass of
Mg = 3.24 × 1011 M⊙ inside R = 25 kpc is dominated by the halo; the disk takes 16% of
the total. We realize the infinitesimally-thin disk by distributing N = 514, 000 equal-mass
particles on the disk plane and by assigning to them random velocities corresponding to
the Toomre parameter of Q ≈ 2. This value of Q fairly well represents the stellar disk
in the solar neighborhood and is large enough to prevent spontaneous generation of spiral
arms via swing amplification in the absence of tidal forcing (e..g., Sellwood & Carlberg 1984;
Bertin et al. 1989b). Before applying tidal perturbations, we evolve the galaxy in isolation
for two Gyrs to relax the phase space distribution into a global equilibrium. Appendix B
presents the temporal evolution of an isolated disk and radial profiles of various quantities
when an equilibrium is reached. We take the particle distribution at 1 Gyr and use it as an
initial condition for tidal encounter experiments. This guarantees that morphological and
structural changes of the disk occurring during interactions with the perturber are entirely
due to tidal perturbations.
1Since the velocity dispersions of dark matter particles are usually much larger than those of disk
stars, the impact of the perturber to the disk through the live halo and bulge is small, as confirmed by
Salo & Laurikainen (2000b).
2We have also run models without halo truncation and checked that the properties of tidal features inside
25 kpc are almost indistinguishable from those under the truncation.
– 6 –
2.2. Perturber and Model Parameters
As a perturbing companion, we consider a point particle with mass Mp that moves on
a parabolic orbit relative to the center of the galaxy in a prograde fashion. To study the
excitation of spiral arms as cleanly as possible (i.e. without disk warping and bending waves)
and to be consistent with the thin-disk approximation, its trajectory is confined to the same
plane as the galactic disk. Assuming that the galaxy whose center lies at R = 0 is spherical,
the relative orbit (Rp, φp) of the perturber in the polar coordinates is given parametrically
by
Rp = Rperi(1 + x2),
t =(cid:20)
1/2
2R3
peri
G(Mg + Mp)(cid:21)
(x + x3/3),
(1)
(2)
where Mg is the total galaxy mass within 25 kpc, x ≡ tan(φp/2), and Rperi is the pericenter
distance (e.g., Press & Teukolsky 1977). Note that t = 0 (or x = 0) corresponds to the
pericenter passage of the perturber.
To explore tidal encounters with various strength, we consider nine self-gravitating mod-
els that differ only in the mass and the pericenter distance of the perturber. We also run
one non-self-gravitating model to study the effect of self-gravity on the arm properties. Ta-
ble 1 lists the parameters of each model and some simulation outcomes. Column (1) labels
each run. Columns (2) and (3) give the perturber masss relative to the total galaxy mass
and the pericenter distance, respectively. Column (4) lists the dimensionless tidal strength
parameter defined by
S =(cid:18)Mp
Mg(cid:19)(cid:18) Rg
Rperi(cid:19)3(cid:18) ∆T
T (cid:19) ,
(3)
which measures the momentum imparted by the perturber to a disk particle at Rg = 25 kpc
relative to its original angular momentum (Elmegreen et al. 1991). Here, ∆T is the time
elapsed for the perturber to move over one radian near the pericenter relative to the galaxy
g/GMg)1/2 is the time taken by stars at R = Rg to rotate one radian
center, and T ≡ (R3
about the galaxy center. Columns (5) and (6) give the fractions of the disk particles that are
captured by the companion and those escaping from the whole system, respectively. Column
(7) gives the time ttail when the tidal tail becomes strongest, while columns (8) and (9) list
the pitch angle itail and surface density Σtail of the tail at t = ttail. Finally, column (10) gives
the peak strength of the spiral arms. Model A2* is identical to model A2 except that the
self-gravity of density perturbations in the disk is artificially taken to zero in the former. Note
that the self-gravitational potential of the unperturbed axisymmetric disk, as represented by
equation (A4), is still included in model A2* to make the rotation curve intact. Models A1
and C3 correspond to the strongest and weakest encounters, respectively.
– 7 –
In our presentation, the units of length and velocity are 1 kpc and 1 km s−1, respectively,
which give the characteristic time unit of t0 = 0.98 Gyr. All the simulations run from
t/t0 = −1.0, corresponding to (Rp, φp) = (178.3 kpc,−45.6◦) for our fiducial model A2, to
t/t0 = 3.0. Seen from the above, the perturber passes through the pericenter (Rperi, 0) at
t = 0 in the counterclockwise direction which is the same sense as the disk rotation.
2.3. Numerical Method
To evolve disk particles in response to tidal perturbations, we use the GADGET code
In GAD-
that is parallelized on a distributed-memory platform (Springel et al. 2001).
GET, the evaluation of gravitational force uses the Barnes-Hut hierarchical tree algorithm
(Barnes & Hut 1986) and assumes a spline-softened mass distribution of a point mass. Ex-
cept at the beginning of the simulations, GADGET employs a new cell-opening criterion
Ml4 > αaoldr6, which produces, at a lower computational expense, force accuracy compa-
rable to that obtained from the standard criterion r > l/θ. Here, α and θ specify prescribed
force error tolerances, M and l are the total mass and size of a cell, r is the distance of a
particle to the center-of-mass of the cell, and aold is the gravitational acceleration on the
particle computed at the previous timestep. For all the models presented in this paper, we
adopt α = 0.02 and θ = 0.8.
For the gravitational softening, we take a softening length of h = 0.4 kpc; the equivalent
Plummer softening length is ǫ = h/2.8 = 0.14 kpc (Springel et al. 2001). The relaxation time
associated with the force softening amounts to tR ≈ σ3ǫ/(πG2Σ0m), where σ = √σRσφ and
m is the particle mass (Rybicki 1971). Since this time is longer than 10 Gyrs for R > 1
kpc and N = 5 × 105, tidal features that form in the stellar disk are not contaminated by
particle noises and relaxation (e.g., White 1988). Particles are advanced by a second-order
leapfrog scheme with fully adaptive and individual timesteps. All the simulations have been
performed on an IBM p690 cluster using 16 processors, taking typically ∼ 25 hours for a
single run.
2.4. Limitations of This Work
In this work we employ highly idealized models of galaxies, perturbers, and their tidal
encounters, and consider a limited range of tidal strength. This obviously introduces a few
important caveats that should be noted from the outset:
1. An infinitesimally-thin stellar disk imposed in the simulations neglects non-planar mo-
– 8 –
tions of stars such as in vertical oscillations and warps. It also overestimates self-gravity
at the disk midplane.
2. While a perturbing companion more likely has an extended density profile in real
situations, we represent it as a point mass, which may overestimate the tidal force at
the closest approach, possibly affecting the shape and structure of tidal bridge and tail
that form in extended disks.
3. Since we treat the galactic halo and bulge as being dynamically inactive, it is convenient
to evolve the entire system in the coordinates centered on the center of the galaxy. This
naturally ignores centrifugal and Coriolis forces that arise from the orbital motion of
the galaxy relative to the center of mass of the whole system. The neglect of the
indirect forces may spuriously suppress the growth of m = 1 spiral modes in the stellar
disk, where m is the azimuthal wavenumber (e.g., Adams et al. 1989; Ostriker et al.
1992), although it is unlikely to much affect m = 2 and higher order modes.
4. By employing the prescribed parabolic orbit for a perturber, we neither consider the
back reaction of the stellar disk to the perturber nor allow multiple encounters that
would occur if the perturber is in a bound orbit. Furthermore, the prescribed orbit and
the rigid halo and bulge do not allow us to capture the potential effects of dynamical
friction and ensuing orbital decay of the interacting galaxies, which may make the tidal
tails longer and stronger (see e.g., Barnes 1988).
5. Limited to the cases with S <∼ 0.25, tidal tails created in our models are relatively weak
and survive only for ∼ 0.3 Gyrs (see §3.2). The current weak- or moderate-encounter
models preclude the possibility of prominent tails found in many interacting systems
that live long (∼ 1 Gyr or longer) and sometimes fragment into tidal dwarf galaxies
(e.g., Barnes 1988, 1992; Barnes & Hernquist 1992; Wetzstein et al. 2007), which may
occur when tidal interactions are very strong.
Given these constraints and limitations, we by no means attempt to reproduce tidal
deformation of real galaxies. We instead focus on the formation mechanisms and physi-
cal nature of tidally-driven disk structures, and compare the simulation results with the
predictions of analytic theories, for which the simplifications made above are appropriate.
3. Extended Tidal Features
Using a restricted three-body technique, Toomre & Toomre (1972) demonstrated that
tidal perturbations distort the extended portions of a disk to produce elongated and narrow
– 9 –
features, phenomenologically termed “bridge” and “tail”. The bridge is built at the near side
of the disk toward the perturber, while the tidal tail or “counterstream” forms at the far side
(e.g., Pfleiderer 1963). Self-consistent numerical simulations including the disk self-gravity
show that tidal perturbations excite not only extended tidal streams but also spiral arms
in the main disks (e.g., Hernquist 1990; Byrd & Howard 1992; Salo & Laurikainen 1993).
In this section, we focus on extended tidal features and distinguish between the physical
mechanisms that form bridge and tail, some of which have previously been overlooked.
3.1. Tidal Bridge
To illustrate the dynamical responses of extended disks to a tidal perturber, we begin
by presenting in detail the results from our fiducial model A2 with Mp/Mg = 0.4 and
Rperi = 35 kpc. Evolution of the other models are qualitatively similar. Figure 1 shows the
morphological evolution of the stellar disk in model A2. The arrow and the associated number
in each panel indicate the direction and distance (in kpc) to the perturber, respectively. Only
20% of the particles are plotted to delineate tidal features from the disk. Figure 2 displays
the perturbed surface densities of model A2 in the φ − log R plane. At early time when
the perturber is far away from the galaxy (t <∼ − 0.1), the tidal deformation of the disk is
vanishingly small. As the perturber approaches the pericenter, the disk begins to undergo
significant morphological changes, first forming a bridge (t ∼ 0.0 − 0.1) at the outskirts of
the disk close to the perturber and then a tail at the opposite side (t ∼ 0.2).
Tidal force imposed by the perturber excites the epicycle orbits of individual particles.
In Appendix C, we use an impulse approximation to estimate the amplitudes δR of the
perturbed epicycle orbits in an averaged sense. Figure 3 plots as thin lines the resulting δR
with differing Mp based on the impulse approximation. The thick line is for the case with no
tidal perturbations in which the radial oscillations of particles are purely due to the initial
velocity dispersions. Also plotted as various symbols are the dispersions h(R − R0)2i1/2 of
the particle positions R at t = 0 with respect to the initial locations R0 for models A2, B2,
and C2. Here, the angular brackets h i denote an average over the particles in a given radial
bin. Note that the numerical results are in good agreement with the corresponding analytic
In regions of disks with R <∼ 15 kpc, the deviation
estimates over a wide range of radii.
from the original epicycle orbits is quite small. It nevertheless enables well-defined spiral
structure there, as we will discuss in §4. In the extended disks, on the other hand, strong
tidal perturbations severely affect the orbits of particles, causing them to traverse over large
radial distances.
Since the tidal force is asymmetric, particles at the near side to the perturber are
– 10 –
more easily pulled radially outward and will subsequently find themselves subject to even
greater tidal force at larger R.3 Particles whose velocities exceed the escape velocity become
unbound, and either are captured by the perturber or escape from the combined galaxy-
perturber system (Toomre & Toomre 1972). The fraction of the captured particles and
the non-captured, freely escaping particles are given in columns (5) and (6) of Table 1,
respectively; these are fitted roughly with Mcap/Md = 0.95S2.08 and Mesc/Md = 0.67S2.92
for 0.04 <∼ S <∼ 0.3. In model C3 with S = 0.029, the tidal force is too weak to accelerate
particles to the escape velocity. Although less than 7% of the total even in our strongest
encounter model A1, the amount of mass stripped off by the tidal force depends fairly steeply
on the tidal strength, and can be substantial for encounters with large S.
While the bridge is a pathway through which mass transfer occurs, it also contains a
significant amount of bound particles. Due to strong tidal force, the orbits of these bound
particles are eventually arranged in such a manner that the maximum radial velocities always
occur in the direction to the perturber while the perturber remains close to the pericenter.
This is well illustrated in Figure 4 which plots the azimuthal distributions of the particle
velocities at R = 20 kpc in model A2 for t ≤ 0.3. In each panel, the vertical dotted lines
indicate the direction, φp, to the perturber. Although the morphological change of the disk
is almost absent when t = −0.1 (see Fig. 1), the signature of the tidal interaction is already
apparent in the azimuthal variations of the particle velocities. Note that in the bridge both
vR and ∂vφ/∂φ are maximized at φ = φp(t) at the epochs shown in Figure 4. That is,
the phases of particle orbits are locked to the perturber during this time interval. Since the
epicycle motions occur in the opposite sense to the disk rotation, this phase locking implies
that vφ steadily decreases as the particles continue galactic rotation past the perturber. It
attains minimum values near the leading edge of the bridge. It is at this leading edge where
the particles fall rapidly radially inward, rendering the leading boundary of the bridge rather
sharp.
Figure 5 displays distortions of rings at several different initial radii R0 during the early
phase of the tidal encounter. Near-side particles in a ring with larger R0 are pulled out earlier
and by greater amount toward the perturber, shaping the ring into an egg-shaped oval. The
tips of outer ovals become lagging behind the perturber.4 At the same time, new particles
from inner rings that rotate fast are pulled out to lead the perturber. This constructs a
3In the case of model A2, the ratio of the tidal forces at the near and far sides is 2.0 and 7.5 at R = 8
and 20 kpc, respectively.
4The perturber in model A2 has an angular velocity of Ω = 9.54 km s−1 kpc−1 at the pericenter. Since
the corresponding corotation radius is R = 25 kpc in the disk, all the near-side particles shown in Figure 5
would lead the perturber were it not for strong tidal perturbations and the resulting phase locking.
– 11 –
transient pattern that persists while the perturber is close to the pericenter (∼ a few tenths
of Gyrs), with the pattern speed roughly equal to the instantaneous angular velocity of the
perturber. As Figure 1 shows, the bridge in model A2 (and also in other models) lasts
until t ∼ 0.3 after which the perturber is too far away to tightly enforce the alignment of
the epicycle orbits. Therefore, the bridge is a transient structure that not only allows mass
transfer to the companion but also consists of bound particles that execute coherent forced
oscillations in response to the applied tidal perturbations.
3.2. Tidal Tail
Tidal torque applied at the far side of the disk causes the leading (lagging) particles
with respect to the line connecting the disk center and the perturber to lose (gain) angular
momentum and thus to rotate slower (faster). This gives rise to a negative gradient of the
circular velocity along the azimuthal direction. One may naively expect that the compressive
velocity fields in the azimuthal direction should be a cause of a tidal tail at the far side, but
this is not the case. The third panel of Figure 4 shows that the velocity gradient amounts
to ∂vφ/∂φ ∼ −50 km s−1 rad−1 for the far-side particles at R = 20 kpc when the perturber
is at the pericenter. Assuming that this value remains constant over a time interval ∆t, the
resulting fractional change δΣ/Σ of the surface density would be ∼ ∆t∂vφ/(R∂φ) ∼ 0.5 for
∆t = 0.2 Gyr, which is too small to build a tidal tail in its own right. Therefore, the tail
formation must involve additional processes.
Figure 5 demonstrates the tail-making process in our models. Let us pay attention to the
two groups of particles, denoted by dots in black or cyan, in the ring with R0 = 22− 24 kpc.
The dots in cyan representing a group of far-side particles at t = −0.05 are slowly rotating
about the disk center, with a period of ∼ 0.6 Gyr, by following moderately perturbed epicycle
orbits. With relatively weak tidal force, the locking of the epicycle phases is not significant
at the far side. On the other hand, the near-side particles in black that were ahead of the
perturber at t = −0.05 have highly perturbed orbits, plunging toward the disk center as
deep as R ∼ 9 kpc at t = 0.05. The constraint of angular momentum conservation requires
the particles to rotate faster at small R, providing them with a shortcut route to reach the
far side (t ∼ 0.10 − 0.15). A tidal tail develops as these strongly-perturbed, fast-rotating,
near-side particles catch up with those mildly-perturbed, far-side particles (t ∼ 0.15 − 0.2)
(e.g., Pfleiderer 1963; Toomre & Toomre 1972).
Note that the outer-disk particles located in between the black and cyan dots, i.e., the
particles with φ ∼ π/4 − π at t = 0 in the red ring in Figure 5, are all gathered into the tail
extending to ∼ 40 kpc from the disk center. Since the tail at a given radius is comprised of
– 12 –
particles from a wide range of radii in the unperturbed disk, it has large velocity dispersions
both in the radial and azimuthal directions (t = 0.2 frame of Fig. 4). Accordingly, the tail
in model A2 becomes weak and dispersed as the particles continue galactic rotation. This
implies that the tails in our models are transient material arms.
Figure 2 shows that the tail in model A2 forms at t ≈ 0.2, is more or less logarithmic
in shape with a pitch angle of tan i ∼ 0.5, and becomes more pronounced than the bridge.
Time of tail formation ttail, its pitch angle itail, and its strength as defined by the surface
density Σtail at R = 20 kpc and t = ttail of course depend on the the strength of tidal
perturbations and the presence of self-gravity. Columns (7)-(9) in Table 1 list ttail, tan itail,
and Σtail for models with S > 0.07; models B3, C1, C2, and C3 with very weak tidal
perturbations do not produce readily identifiable tail structure. These values are plotted in
Figure 6 as solid circles against S, which are well fitted with power laws: ttail = 0.07S −0.54,
tan itail = 0.75S0.22, and Σtail/Σ20 = 79.0S0.72, where Σ20 indicates the surface density of the
initial disk at R = 20 kpc. Definitely, a tail develops earlier and stronger for stronger tidal
perturbations, although the pitch angle depends weakly on S. Note that the tidal tail in the
non-self-gravitating model A2* is weaker and more loosely wound than that in model A2.
As mentioned above, a tidal bridge in the near side consists of particles in coherent
forced oscillations, while a tail in the opposite side forms by temporary particle overlapping.
Both are transient features whose amplitudes decay after t ∼ 0.2 − 0.3. As the perturber on
a parabolic orbit moves away from the galaxy in our models, the diminished tidal force no
longer aligns the phases of the particle orbits in the bridge. In addition, the large velocity
dispersions of the tail are unable to keep it as narrow as when it first forms. Consequently,
the particles making up the bridge and tail gradually spread out and follow galactic orbits
with large eccentricities. They interact with each other and also with spiral arms, producing
complicated structures seen at the extended parts of the disk in Figures 1 and 2. The further
diffusion and interactions of particles eventually make the outer disk almost featureless in
our simulations.
4. Disk Structure
We have seen in §3.1 that the enhancement of epicycle amplitudes due to tidal per-
turbations is rather small in regions of disks with R <∼ 15 kpc. Nevertheless, the phases of
perturbed epicycle orbits at different radii drift at different rates and are kinematically orga-
nized to develop a trailing two-armed spiral pattern there (e.g., Toomre 1969; Donner et al.
1991). Figure 7 displays close-up views of density snapshots of model A2 in the x–y plane.
A well-defined, two-armed spiral pattern is apparent for t ∼ 0.2 − 1.0, becoming most con-
– 13 –
spicuous at t ∼ 0.3 − 0.4. The spiral arms that appear as straight lines in the φ− ln R plane
(see Fig. 2) are approximately logarithmic, with a pitch angle varying with time.
An inspection of Figure 2 reveals that the arms extend inward up to R ≈ 4 kpc,
corresponding to the inner Lindblad resonance (ILR), and smoothly join the the extended-
disk features at t = 0.2. Since the pattern speed of the arms is smaller than the angular
speed of the disk rotation, however, they soon decouple from the tidal tail (t = 0.3), and
from the bridge at later time (t = 0.4 − 0.5) when the phase locking becomes inefficient.
The spiral arms in our models are not stationary in the sense that their pattern speed is
not constant over radius and that their pitch angle and amplitude vary with time. In this
section, we explore the quantitative properties of the spiral arms.
4.1. Pitch Angle
Since the spiral arms that form in our models are logarithmic, it is useful to define the
Fourier coefficients in φ and ln R as
A(m, p) =
1
N
N
Xj=1
exp(i[mφj + p ln Rj]),
(4)
where N is the total number of particles, (Rj, φj) are the coordinates of the j-th par-
ticle, and p is related to the pitch angle of an m−armed spiral through tan i = m/p
(Sellwood & Carlberg 1984; Sellwood & Athanassoula 1986). A positive (negative) value
of p corresponds to trailing (leading) spirals.
Figure 8 plots the temporal evolution of the Fourier amplitudes A(2, p) of the m = 2
logarithmic spiral mode in model A2 before the arms reach the maximum strength (t < 0.3).
We consider particles only at R = 5 − 10 kpc where the pattern in model A2 achieves
large amplitudes and contamination from the bridge and tail is almost absent. At early
time (t <∼ 0.04), the modal growth occurs as the dominant p shifts from negative to positive
values. This is suggestive of mild swing amplification in which seed perturbations grow as
they change from leading to trailing. Since the corresponding amplification factor is less than
10 when Q ∼ 2 (Goldreich & Lynden-Bell 1965; Julian & Toomre 1966; Toomre 1981) and
since swing amplification becomes no longer efficient at p >∼ 5 (e.g., Sellwood & Carlberg
1984), however, the further growth of the spiral modes cannot be attributable to swing
amplification.
It is rather due to the kinematic effects, enhanced by self-gravity, of the
perturbed epicycle orbits in a manner described in Toomre (1969). As the phases of the
epicycle orbits drift and are coherently arranged, the density associated with the pattern
grows quite rapidly and saturates at t ∼ 0.3 in model A2.
– 14 –
It is well known that kinematic density waves without self-gravity tend to wind up due
to the background differential rotation, with the pitch angle varying as
−1
tan i = t−1(cid:12)(cid:12)(cid:12)(cid:12)
d(Ω − κ/2)
d ln R (cid:12)(cid:12)(cid:12)(cid:12)
,
(5)
(e.g., Binney & Tremaine 1987). In the theory of quasi-stationary density waves hypoth-
esized by Lin & Shu (1964, 1966), self-gravity of the spirals compensates for the winding
tendency of the arms, keeping their pattern speed at a constant value over a wide range of
radii. In order to check if this is the case in our simulations, we calculate the pitch angle of
the arms determined from p that maximizes A(2, p) at a given time. Figure 9 shows the
temporal changes in tan i of the spiral arms located at R = 5−10 kpc, 8−13 kpc, and 11−16
kpc for the self-gravitating models A2, B2, and C2, respectively. For comparison, Figure 9
also plots the results of the non-self-gravitating model A2* for the arm segments in an an-
nulus with R = 8.0 − 8.5 kpc, over which d(Ω − κ/2)/d ln R is almost constant. Although
the arm pitch angle in model A2* exhibits small fluctuations, the late-time portion can be
well described by tan i ∝ t−1, consistent with the theoretical prediction (eq. [5]).5 For the
self-gravitating models, the arms have moderate pitch angles amounting to tan i ∼ 0.3 − 0.4
when they grow and stand out initially. After attaining substantial strength, they begin
to wind as tan i ∝ t−0.5∼−0.6, with a smaller power index corresponding to stronger arms.
This suggests that although self-gravity reduces the winding rate considerably, it cannot
completely suppress the winding tendency of the spiral arms in our models.6
Once finding the arm pitch angle and pattern speed (see below), we are able to compare
the WKB theory of linear density waves with the simulation results. The local theory for
tightly-wound linear waves in a stellar disk states that the perturbed radial velocity δvR and
azimuthal velocity δvφ are related to the perturbed surface density δΣ through
δvR = −
δvφ = −
i
2
κ2
ΩkR
νκ
Σ (cid:19) ,
kR (cid:18)δΣ
F (2)
Fν(x) (cid:18)δΣ
Σ (cid:19) ,
ν (x)
(6)
(7)
5 Since the Fourier method picks up, in a given annulus, the most dominant spiral modes that propagate
radially inward, the time dependence of tan i can also be affected by the radial variation of d(Ω− κ/2)/d ln R
if the annulus is wide enough. For instance, the average pitch angle of the arms in the R = 8− 13 kpc region
in model A2*, over which d(Ω−κ/2)/d ln R varies by 13% relative to the mean value, decays as tan i ∝ t−0.94.
6 In addition to the background shear, short trailing waves in the presence of self-gravity would increase
their radial wavenumber kR as they propagate inward from the corotation radius, capable of decreasing the
pitch angle further (Toomre 1969).
– 15 –
(Lin, Yuan, & Shu 1969). Here, kR ≡ m/(R tan i) is the local radial wavenumber of the
waves, x ≡ (kRσR/κ)2, ν ≡ (Ωp − mΩ)/κ is the dimensionless angular frequency with Ωp
denoting the pattern speed, and Fν(x) and F (2)
ν are the reduction factors defined by equations
(B9) and (B17) of Lin, Yuan, & Shu (1969), respectively. In equation (7), the imaginary unit
i represents the phase shift between δvφ and δΣ.
Figure 10 gives exemplary comparisons between the numerical results and the linear-
theory predictions for the azimuthal variations of the perturbed variables. Two sets of
numerical data near R = 10 kpc at t = 0.4 in model A2 and at t = 0.5 in model C2 are
arbitrarily taken. In the top panels, black curves with some fluctuations draw δΣ/Σ from
numerical simulations, while red lines plot the corresponding m = 2 Fourier modes δΣm=2.
In the middle and bottom panels, red curves draw equations (6) and (7) corresponding to
δΣm=2. Blue curves represent the azimuthally-binned averages of vR and δvφ = vφ − ¯vφ that
are plotted as dots from the simulations. Apparently, the perturbed density in model C2 is in
the linear regime and dominated by the m = 2 mode. Note that in spite of large dispersions
in vR and δvφ, there is fairly good agreement between the numerical and analytic results
for model C2. On the other hand, the spiral arms in model A2 are asymmetric and clearly
in the nonlinear regime. In this case, the perturbed velocities have significant contributions
from high-m modes (e.g., Vandervoort 1971), rising more steeply than a simple sinusoidal
curve as particles leave the spiral arms.
Among the models listed in Table 1, we found that models B3, C2, and C3 with relatively
weak tidal perturbations (S < 0.06) produce linear spiral arms with sinusoidal density
distributions. All the other models we considered show significantly nonlinear features in
the density and velocity profiles. This implies that tidally-excited stellar spiral arms in
grand-design spiral galaxies probably have non-linear amplitudes.
4.2. Arm Strength
One of the key parameters that directly influence gas flows in spiral galaxies is the
strength of stellar spiral arms. Stronger spiral arms imply larger enhancement of gas density
at the galactic shocks and hence more active star formation. To quantify the arm strength,
we define
2πGδ Σm=2
RΩ2
F ≡
,
(8)
where δ Σm=2 denotes the amplitude of δΣm=2. Since the corresponding gravitational poten-
R + m2R−2)1/2 for a tightly-wound spiral
tial perturbation is given by δΦm = −2πGδΣm/(k2
in an infinitesimally-thin disk, F measures the gravitational force due to the spiral arms in
– 16 –
the direction perpendicular to the arms relative to the the axisymmetric radial force RΩ2 in
the unperturbed state (e.g., Roberts 1969; Shu, Milione, & Roberts 1973; Kim & Ostriker
2002).
Figure 11 plots the radial variations of F for the arms averaged over the time interval
∆t = 0.4 centered at the epoch when the arm amplitudes are maximized. In a given model,
F broadly peaks at a certain range of radii; Rmax ∼ 5 − 10 kpc for models A1, A2, B1, and
C1, Rmax ∼ 8 − 13 kpc for models A2*, A3, and B2, and Rmax ∼ 11 − 16 kpc for models
B3, C2, and C3. This demonstrates that more distant encounters excite spiral features in
regions with larger R. The arms become progressively weaker toward the disk center since
the ratio of the tidal perturbations to the background gravity is proportional roughly to
R2 for small radii. They eventually attain vanishingly small amplitudes inside R ≈ 4 kpc
corresponding to the ILR through which stellar spiral waves cannot propagate.7 Although
the tidal perturbations are strong in extended disks, on the other hand, F still decreases
with increasing R (> Rmax). This is because the amount of mass available to construct
spiral arms in the background stellar disk declines very rapidly with radius. Figure 11 also
shows that self-gravitating spiral arms in model A2 are stronger by about a factor of 1.5 and
located relatively closer to the center than non-self-gravitating arms in model A2*.
To study how rapidly spiral arms grow and how long they survive, we plot in Figure 12
the temporal variations of F averaged over a range of radii where the arms are strongest
in each model. Obviously, the arms grow earlier and more rapidly for models with stronger
tidal perturbations. For instance, it takes only ∼ 0.1 − 0.3 Gyrs for the strong-encounter
models A1 and A2 to achieve the peak strength, while more than 1 Gyrs are required for
the weak encounter models. Figure 13 plots the peak value Fmax of the arm strength as a
function of the tidal strength S, showing roughly Fmax = 0.79S0.83.
Since the formation of tidal spiral arms in a disk involves the gathering of particles from
different radii, the velocity dispersions increase as the arms grow. In addition, gravitational
scatterings of stellar particles off the arms become efficient to heat the disk once the arms
acquire considerable amplitudes, counterbalancing the arm-amplifying effect of self-gravity
(e.g., Sellwood & Carlberg 1984; Binney & Tremaine 1987; Binney 2001). In all the models
we have considered, the arms never become fully self-gravitating. They stop growing and
decay as the enhanced velocity dispersions make the once well-organized epicycle orbits
kinematically less coherent. Figure 12 shows that for the self-gravitating models, F decreases
7It is unclear whether the absence of spiral arms at R < 4 kpc in our models is mainly due to the ILR
barrier or just because the tidal perturbations are too weak to excite density waves there. We have run a
model simulation (not listed in Table 1) corresponding to model A2 but without the bulge (hence no ILR)
only to find that the inner disk is contaminated by the formation of a central bar (e.g., Noguchi 1987).
– 17 –
after the peak almost exponentially in a time scale of ∼ 1 Gyrs, whereas spiral arms in the
non-self-gravitating model A2* decay much more slowly since they do not experience secular
disk heating. Strong encounter models possess spiral arms with F ≥ 5% for ∼ 1 Gyrs,
corresponding to four disk rotations at R = 10 kpc, while spiral arms in the weak-encounter
model C3 have F <∼ 3% throughout the entire evolution. Small-amplitude fluctuations of F
at t >∼ 1 are caused by the interactions with the particles once pertaining to the bridge and
tail.
4.3. Pattern Speed
Finally, we discuss the pattern speed of the tidal arms formed in our simulations. To
measure the pattern speed at a given radius, we define the normalized cross-correlation of
the perturbed surface densities at two fixed times separated by ∆t as
C(R, θ, t) =
1
Σ0(R)2 Z 2π
0
δΣ(R, φ, t)δΣ(R, φ + θ, t + ∆t)dφ.
(9)
For a sufficiently small value of ∆t, the instantaneous arm pattern speed at a given radius
is then determined by Ωp(R, t) = θmax/∆t, where θmax denotes the phase angle at which
C(R, θ, t) is maximized. We take ∆t = 0.1 in calculating Ωp from the numerical data.
Figure 14 plots as contours the amplitudes of C(R, θ, t) on the radius (R)−frequency
(θ/∆t) domain for some selected time epochs of models A2 and A2*. The solid and dashed
lines draw the radial variations of Ω and Ω± κ/2, respectively, from the initial disk rotation.
At t = 0.1, the spiral arms in both models are relatively weak and the cross-correlation is
dominated by the extended-disk features, especially by the tidal bridge. The bridge rotates
almost rigidly at a fixed pattern speed (∼ 9.5 km s−1 kpc−1), corresponding to the angular
frequency of the perturber at the pericenter. This evidences the phase locking of particle
orbits in the bridge explained in §3.1. The tail at the opposite side of the perturber becomes
strong at about t = 0.2, significantly contributing to C(R, θ, t) at R >∼ 17 kpc. Interestingly,
the instantaneous pattern speed of the tail is similar to that of the bridge at this time. As
time evolves further, the extended tidal structures become weaker since the perturber moves
farther away, while the spiral arms become more pronounced in the distribution of C(R, θ, t).
When the arms are quite strong (t ∼ 0.2−0.6) in model A2, their patten speed decreases
with radius, indicating that they are not a “pattern” in a strict sense. This is the reason
why the pitch angle of the arms decreases with time. Since the axisymmetric background
state of the stellar disk as well as the shape and pitch angle of the arms are already known,
– 18 –
one can calculate the theoretical pattern speed predicted from the WKB dispersion relation
ν2 = 1 −
2πGΣkR
κ2
Fν(x),
(10)
for tightly-wound density waves (Lin, Yuan, & Shu 1969). The dotted line in each of the
left panels of Figure 14 shows Ωp obtained from equation (10), which traces the loci of
maximum C(R, θ, t) fairly well. Note that equation (10) would simply yield ν ≈ −1 or
Ωp = Ω − κ/2 without self-gravity, in excellent agreement with the pattern speed of spiral
arms in model A2* for t >∼ 0.3. Although the presence of self-gravity tends to enhance
the arm pattern speed, our numerical results suggest that its effect is quite small; for all
the models considered, Ωp is below ∼ 20 km s−1 kpc−1 even when the arms reach the peak
strength, and it comes very close to the Ω−κ/2 curve at t >∼ 0.6. This implies that the spiral
arms at least at late time are kinematic spiral waves in which the large velocity dispersions
of particles as well as the kinematic winding of the arms make self-gravity unimportant.
5. Summary & Discussion
5.1. Summary
Galactic spiral shocks and their substructure-forming instabilities in disk galaxies are
strongly affected by stellar spiral arms that are often triggered by tidal interactions with
a companion galaxy. To gain an insight on the large-scale star formation occurring in
the gaseous component and related evolution of disk galaxies, it is crucial to understand
the physical properties of tidally-induced stellar arms. While the literature abounds with
studies of tidal interactions of galaxies, most of them concentrate mainly on morphological
transformation, especially in the extended parts, of disk galaxies.
In this paper, we have initiated numerical N-body experiments for tidal encounters to
quantify the properties of spiral arms that form in the disks and study how their properties
vary with tidal strength. We also study the nature of the tidal bridge and tail that develop
in the outer regions. We consider a simple galaxy model consisting of a rigid halo/bulge and
a razor-thin stellar disk with Toomre stability parameter of Q ≈ 2. A perturbing companion
galaxy is treated as a point-mass potential moving on a prescribed, prograde, parabolic orbit
in the same plane as the galactic disk. By varying the mass and pericenter distance of the
perturber, we explore tidal interactions with strength in the range of 0.03 <∼ S <∼ 0.3, where
S is the dimensionless momentum applied by the perturber to stars at outer disks (see eq.
[3]).
Our main results are summarized as follows.
– 19 –
1. The tidal bridge forms at the near side to the perturber as particles in outer disks
are pulled out by strong tidal perturbations. Some particles with velocities exceeding the
escape velocity become unbound, and either are captured by the perturber or escape from
the system, but these are less than 7% of the total for S <∼ 0.3. On the other hand, bound
particles with low velocities in the bridge execute coherent forced oscillations in such a way
that the maximum radial velocities vR and the maximum gradient of the azimuthal velocities
∂vφ/∂φ are always attained in the direction toward the perturber. This phase locking of
the perturbed particle orbits allows the bridge to construct a transient pattern that corotates
with the perturber as long as the perturber remains close to the pericenter (t <∼ 0.3). The
phase locking is also a cause of the sharp leading edge of the bridge, where particles begin
to fall radially inward during their forced oscillations.
2. Only strong tidal encounters with S > 0.07 produce a recognizable tail (or counter-
stream) at the far side of the disk. The tail develops as strongly-perturbed, near-side particles
overtake mildly-perturbed, far-side particles. When the tail achieves a peak strength, it is
very narrow and in a roughly logarithmic shape. For 0.07 <∼ S <∼ 0.3 we have considered, the
formation epoch ttail, pitch angle itail, and the surface density Σtail of the tail depend on the
tidal strength parameter S as ttail = 0.07S −0.54, tan itail = 0.75S0.22, and Σtail/Σd = 79S0.72
at R = 20 kpc. Comprising of particles collected from a wide range of radii in the unper-
turbed disk, the tail is a material arm and has large velocity dispersions, so that it widens
and weakens with time.
3. Even though the boost of epicycle amplitudes due to tidal perturbations is quite
small in regions with R <∼ 15 kpc, the perturbed particle orbits are kinematically organized
to generate two-armed global spiral arms there. With Q ≈ 2 in the unperturbed disk, the self-
gravity of stars does not play a dominant role in growing the spiral modes, although it appears
to enhance the amplitudes considerably when the arms are nonlinear. The spiral arms are
approximately logarithmic in shape and subject to kinematic winding. For the parameters
we have explored, the pitch angle of the spiral arms is in the range of tan i ∼ 0.3 − 0.4 when
the arms attain peak amplitudes and then decreases as tan i ∝ t−0.5∼−0.6, with a smaller
decay rate corresponding to stronger arms.
4. Stronger encounter models tend to develop stronger spiral arms earlier and more
toward the galaxy center, resulting in the arms at R ∼ 5 − 10 kpc, ∼ 0.1 − 0.3 Gyr after
the pericenter passage for models with S > 0.13. Arms are absent inward of R = 4 kpc
corresponding to the inner Lindblad resonance. In terms of the parameter F (eq. [8]) that
measures the perturbed radial force due to the spiral arms relative to the mean axisymmetric
gravity, the maximum strength of the spiral arms behaves as Fmax = 0.79S0.83. Because of
large velocity dispersions associated with the particle gathering and secular heating, the
– 20 –
arms never become fully self-gravitating and decay after the peak almost exponentially in a
time scale of ∼ 1 Gyr.
5. Analyses using the normalized cross-correlation of the perturbed densities reveal that
the arm pattern speed Ωp is not constant in both radius and time, indicating that spiral arms
that form in our models are not exactly a pattern. In fact, Ωp decreases with radius, causing
the pitch angle to decrease with time. Self-gravity tends to increase Ωp, but only below
∼ 20 km s−1 kpc−1 even when the arms are strongest. Self-gravity becomes unimportant as
the arms decay, resulting in Ωp ≈ Ω − κ/2 at late time.
5.2. Discussion
We have seen in this paper that spiral arms produced by tidal encounters are approxi-
mately logarithmic in shape, similarly to observed spiral arms in many grand-design spiral
galaxies (e.g., Kennicutt 1981; Elmegreen et al. 1989; Shetty et al. 2007). The occurrence
of the logarithmic arms in our models can be understood as follows. As mentioned above,
the arms are kinematic density waves modified by self-gravity. Ignoring the effect of self-
gravity and assuming that the phases of the waves are aligned along φ = φp = 0 at t = 0,
corresponding to the impulsive tidal perturbations applied at the pericenter, the pitch angle
of kinematic density waves with m = 2 is given by equation (5). If the right-hand side of
equation (5) is independent of R, the arms have a perfect logarithmic shape. It turns out
that the galaxy rotation curve we adopt (Fig. 15) has an approximately constant value of
d(Ω − κ/2)/d ln R ∼ 3.5 ± 0.5 km s−1 kpc−1 over the distance from the ILR radius out to
the edge of the disk. This results in ∆ tan i/ tan i ∼ 0.15 over a range of radii where spiral
arms are strong, indicating that the variation of the pitch angle along the arms is in fact
very small. The presence of self-gravity as well as epicycle motions of particles are likely to
further smooth out the local variation of tan i.
Our numerical results show that self-gravity is unable to keep the arm pitch angles
fixed over time. A larger rate of shear in the rotation curve implies a smaller arm pitch
angle for kinematic arms. Indeed, Seigar et al. (2005, 2006) reported a well-defined negative
correlation between the arm pitch and the shear rate for a sample of (not necessarily tidally-
driven) spiral galaxies, suggesting that spiral arms in real galaxies are unlikely to be fully
self-gravitating.
While we adopt highly simplified models for both the disk galaxy and the orbital param-
eters of tidal interactions, it is still interesting to compare the arm properties found in our
simulations with those of observed spiral arms. In the case of the M51/NGC5195 system,
– 21 –
the mass ratio of the target galaxy to the companion is estimated to be ∼ 0.3 − 0.55 (e.g.,
Smith et al. 1990; Salo & Laurikainen 2000a). The encounter models that well reproduce
the kinematics and morphologies of the M51 system favor an inclined orbit with the pericen-
ter distance of 20 − 30 kpc (Salo & Laurikainen 2000a). Since the thin disk approximation
and non-inclined orbits taken in our models tend to produce stronger tidal arms than in the
thick-disk, inclined-orbit counterparts, models A1 and A2 can perhaps be best compared
with the M51/NGC5195 system. K-band observations indicate that the radially-averaged
spiral arm strength F is around 20% for M51 (e.g., Scoville et al. 2001; Salo & Laurikainen
2000b; see also Rix & Rieke 1993; Rix & Zaritsky 1995), which is not much different from
∼ 17 − 22% found for models A1 and A2 at t ∼ 0.1 − 0.3 (Figs. 11 and 12). The arms in
M51 are logarithmic spirals with a pitch angle of tan i ∼ 0.39 (Shetty et al. 2007), which is
again close to the arm pitch angle in model A2 at t ∼ 0.2 − 0.3.
Among the properties of spiral arms, the most intriguing is the pattern speed that is not
well constrained by observations. Elmegreen et al. (1989) identified 4 : 1 resonance features
in the arms of M51 to find Ωp ∼ 40 km s−1 kpc−1, while Zimmer et al. (2004) determined
Ωp = 38 ± 7 km s−1 kpc−1 using the Tremaine-Weinberg method. By running collisional
models for cloud dynamics under a given spiral potential, Garc´ıa-Burillo et al. (1993) found
Ωp ∼ 27 km s−1 kpc−1 for the best fit to the observed morphologies of the CO arms in M51.
All of these works were based on the premise that the arm pattern speed is a constant with
radius. However, our numerical results show that the pattern speed of tidal arms depends
on the radius. In the case of model A2, Ωp is a decreasing function of radius, varying when
the arms are strongest from ∼ 20 km s−1 kpc−1 at the ILR to ∼ 10 km s−1 kpc−1 at the
outer parts, and at later time converging to the Ω− κ/2 curve. A similar trend was obtained
by Salo & Laurikainen (2000b) who ran more realistic encounter models (with a star-only
disk) for the M51 system and found that Ωp is close to the Ω− κ/2 curve for a range of radii
where the spiral arms are strong. Although much remains uncertain regarding the effects
of the cold gaseous component and rotation curve, these results suggest that tidally-driven
arms may have a pattern speed that varies with radius in real spiral galaxies.
An age distribution of star clusters in M51 shows a narrow peak at 4 − 10 Myrs and
a broad peak at 100 − 400 Myrs (Lee et al. 2005), indicating active star formation at these
epoches. This enhanced star formation is most likely due to strong spiral arms induced by the
tidal interactions with the companion NGC 5195. Since it takes about ∼ 100− 200 Myrs for
the spiral arms in our models A1 and A2 to attain a substantial amplitude, say F = 10%,
after the perturber passes the pericenter, this implies that the closest passages of NGC
5195 might have occurred ∼ 100 − 200 Myr and ∼ 200 − 600 Myrs ago. Salo & Laurikainen
(2000a) proposed two encounter models for the M51 system: a near-parabolic, single-passage
orbit occurred 400–500 Myrs ago and a bound double-passage orbit having taken place 400–
– 22 –
500 Myrs and 50–100 Myrs ago. Considering the delay between the pericenter passage and
the development of strong arms, the cluster age distribution appears to be more consistent
with the double-passage scenario, although it is uncertain what effects the second passage
will make on the pre-existing arms generated at the first passage.
It is well known from the seminal paper of Toomre & Toomre (1972) that tidal inter-
actions distort the outer parts of a galactic disk and create a tidal bridge extending toward
the perturber as well as a narrow tail at the opposite side. They noted a fraction of the disk
material is stripped and transferred through the bridge to the perturber. In this work, we
further show that the bridge is in fact a transient pattern constructed by bound particles
whose orbits are strongly locked to the perturber. As these particles follow galactic rotation,
they are pulled out toward the perturber and then move radially inward at the leading edge,
making the bridge rather sharp. By mapping the final to initial particle positions under
an impulse approximation, Donner et al. (1991) showed that the sharp boundary of a tidal
bridge corresponds to the loci (caustics) of zero Jacobian of the mapping where the orbits
of neighboring particles come very close together. Indeed, Figure 4 shows that the leading
edge has a large velocity dispersion, consistent with the Liouville theorem that dictates the
conservation of the particle density in the phase space.
Unlike a bridge, a tail at the opposite side is a material arm resulting from the overlap-
ping of near-side particles with far-side particles in the extended parts of the disk. Conse-
quently, the tail forms later than the bridge by about a half orbital time, consistent with the
results of Donner et al. (1991) and Byrd & Howard (1992). Our experiments show that the
formation time and pitch angle of a tail are well correlated with the tidal strength parameter
S. While we employed simple models for tidal interactions and limited our simulations to
the cases with S < 0.3, our results appear to be applicable to models with quite strong tidal
perturbations as well. In simulations of merger encounters, for example, Barnes (1992) ran
self-consistent models consisting of a live halo/bulge and a disk with both stars and gas. One
of his models considered interactions between equal-mass disk galaxies, in which one disk
passes directly through the other with the pericenter distance Rperi/Rg = 0.5, corresponding
to S = 1.48. Figure 3 of Barnes (1992) shows that the tail in this model becomes strongest
at t ≈ 1.25, corresponding in our units to ttail ≈ 0.053 after the pericenter passage, and has
a logarithmic shape with tan itail ≈ 0.83, which are remarkably similar to the extrapolation
of our results in §3.2 that yield ttail ≈ 0.057 and tan itail ≈ 0.81. Through a comprehensive
survey of the parameter space, Toomre & Toomre (1972) found that tail shape is insensitive
to the orbital eccentricity e for 0.6 ≤ e ≤ 1 as long as the inclination of the orbit is not so
large (see also Barnes 1998), which is also consistent with our result that tan itail is weakly
dependent on S.
– 23 –
Numerical studies on tidal encounters often report the formation of double arm structure
at the opposite side to the perturber (e.g., Sundin 1989; Elmegreen et al. 1991; Donner et al.
1991). Our simulations also exhibit such double features (see, e.g., t = 0.3 frame of Fig. 1)
which come out as the tidal tail decouples from the spiral arms that, because of the smooth
alignment with the former, are not readily discernible at t = 0.2. Elmegreen et al. (1991)
found that the lagging arm forms by gathering particles streaming away from the near side
and soon merges with the leading arm. This might be a consequence of the zero velocity
dispersion in their unperturbed disk since the ratio of the velocity impulse due to tidal torque
to the initial random velocity is too large to set up well-defined spiral arms in the disks of their
models. Elmegreen et al. (1991) also found that a prograde, in-plane encounter produces a
“ocular” galaxy with oval-shaped, sharp boundaries, provided S > 0.019. A similar structure
can be seen in the t = 0.2 panel of Figure 1, although the boundaries in our models are less
sharp since, as they noted, the formation of ocular shape requires the injected energy from
the perturber to be much larger than the kinetic energy in random particle motions.
We are grateful to an anonymous referee for stimulating suggestions, and to L. Hernquist,
N. Hwang, M. G. Lee, and E. C. Ostriker for helpful discussion. This work was supported in
part by KASI (Korea Astronomy and Space Science Institute) through a grant 2004-1-120-
01-5401. J. K. was supported in part by KOSEF through the Astrophysical Research Center
for the Structure and Evolution of Cosmos and the grant of the basic research program R01-
2007-000-20196-0. The authors would like to acknowledge the computational support from
KISTI Supercomputing Center under KSC-2007-S00-1007.
A. Galaxy Model
In this Appendix we describe the model galaxy we use for tidal encounter experiments.
The galaxy consists a rigid halo/bulge and a live stellar disk. For a fixed spherical halo, we
adopt a truncated logarithmic potential
Φh(r) =(cid:26) 1
0 log (r2
2v2
−GMh(rtr)/r
c + r2) + constant
for r ≤ rtr
for r > rtr
(A1)
where r is the three-dimensional distance from the halo center, rc is the halo core radius,
rtr is the truncation radius, and v0 is the constant rotation velocity the disk would have
at large r if the halo were not truncated (e.g., Lee et al.. 1999). The corresponding halo
mass distribution is Mh(r) = v2
c + r2)] for r < rtr and Mh(r) = Mh(rtr) for r > rtr.
0r2
The constant in equation (A1) should equal −v2
tr) to make
the potential continuous at r = rtr. For the simulations presented in this paper, we take
tr) − 1
0r3/[G(r2
0 log(r2
c + r2
tr/(r2
c + r2
2v2
– 24 –
rc = 7.5 kpc, rtr = 25 kpc, v0 = 220 km s−1, corresponding to Mh(rtr) = 2.58 × 1011 M⊙. A
spherical bulge is modeled by a Plummer potential
Φb(r) = −
GMb
√r2 + a2
,
(A2)
with the scale radius a = 0.23 kpc and the total bulge mass Mb = 1.0 × 1010 M⊙.
Although stars in real galactic disks are distributed with a finite vertical thickness,
for example, amounting to ∼ 330 pc in the solar neighborhood (e.g., Chen et al. 2001;
Karaali et al. 2004), we impose an infinitesimally-thin stellar disk by setting the vertical co-
ordinates and velocities equal to zero throughout the simulations. For the radial distribution
of stellar surface density, we adopt an exponential form
Σd(R) = Σ0 exp(−R/Rd),
(A3)
where R is the galactocentric radius in the disk, Rd is the disk scale length, and Σ0 is the
surface density at the galaxy center. The total disk mass is Md = 2πΣ0R2
d. The gravitational
potential of the disk is given by
Φd(R) = −(GMd/Rd) RhI0( R)K1( R) − I1( R)K0( R)i ,
(A4)
where In and Kn represent modified Bessel functions of the first and second kinds, respec-
tively, and R ≡ R/2Rd (see e.g., Binney & Tremaine 1987). We take Rd = 3.4 kpc and
Σ0 = 711 M⊙ pc−2, corresponding to Md = 5.2 × 1010 M⊙.
To obtain the equilibrium velocity distribution of disk particles under the total gravita-
tional potential Φtot = Φh + Φb + Φd, we follow a method suggested by Hernquist (1993) and
Quinn et al. (1993). We first assume that the radial and azimuthal components, vR and vφ,
of particle velocities obey initially the Schwarzschild distribution function
f (vR, vφ, R) =
Σd
2πσRσφ
exp"−
v2
R
R −
2σ2
(vφ − ¯vφ)2
2σ2
φ
#,
(A5)
where σR and σφ are the radial and azimuthal velocity dispersions, respectively (e.g., Toomre
1964). The mean azimuthal streaming velocity ¯vφ differs from the circular velocity vc deter-
mined solely from the total gravitational potential as v2
c (R) = −dΦtot/d ln R. In the local
approximation in which Σd, σR, and σφ are assumed to vary slowly with R, one can show
that σR and σφ are related to each other through
σ2
φ/σ2
R = κ2/4Ω2,
(A6)
– 25 –
where Ω ≡ vc/R is the local rotational angular velocity and κ2 ≡ 4Ω2 + dΩ2/d ln R is the
square of the local epicycle frequency (e.g., Binney & Tremaine 1987). Then, the usual Jeans
equation in the radial direction for an equilibrium disk leads to
φ − v2
¯v2
c = σ2
R(cid:18)1 −
κ2
4Ω2 − 2
R
Rd(cid:19)
(A7)
(Barnes 1992; Hernquist 1993).
Finally, we express the radial velocity dispersion σR in terms of the Toomre stability
parameter
Q =
κσR
3.36GΣd
,
(A8)
which determines local gravitational stability of a razor-thin disk to axisymmetric perturba-
tions. We adopt a fixed value of Q = 2 everywhere initially. This value of Q corresponds
roughly to solar neighborhood conditions with κ = 36 km s−1 kpc−1, σR = 30 km s−1
(Binney & Tremaine 1987), and Σd = 35 M⊙ pc−2 (Kuijken & Gilmore 1989), and is large
enough to make swing amplification of non-axisymmetric disturbances inefficient. This pre-
cludes the possibility of spiral arms driven spontaneously by the stellar self-gravity (e.g.,
Sellwood & Carlberg 1984; Bertin et al. 1989b).
Figure 15 plots the circular velocity vc(R) and the mean rotational velocity ¯vφ(R) of
our model galaxy as solid and dotted lines, respectively. Also shown as dashed lines are
the separate contributions to vc from halo, bulge, and disk, which have a mass ratio of
Mh : Mb : Md = 0.81 : 0.03 : 0.16 inside R = 25 kpc.
It is apparent that ¯vφ is usually
smaller than vc, indicating that stars, on average, lag behind a circular orbit at the same
galactocentric radius, a phenomenon known as asymmetric drift.
B.
Initial Disk Setup
We initialize the exponential stellar disk (eq. [A3]) by distributing N=514,000 equal-
mass particles and place it under the combined halo and bulge potential (eqs. [A1] and
[A2]). Strictly speaking, the model disk constructed in this way is not in perfect equilibrium
because equations (A5) and (A7) hold true only in a local sense, that is, only when the
gravitational potential and the stellar velocity dispersions do not vary much with radius
(e.g., Sellwood 1985). In addition, when the disk is allowed to evolve, any non-axisymmetric
modes that grow may interact with particles, feeding them with random kinetic energy.
Two-body interactions of particles tending to heat the disk are not completely negligible,
either. All of these may cause the disk structure to deviate considerably from the desired
one even before undergoing tidal encounters.
– 26 –
To obtain a disk configuration representing a dynamically well-relaxed, global equilib-
rium, we evolve our model galaxy in isolation for two Gyrs. Figure 16 displays snapshots of
particle distributions from the isolated disk evolution. The disk is rotating in the counter-
clockwise direction and time is expressed in units of Gyr. Other than weak non-axisymmetric,
trailing structures seen at its outskirts, the disk does not suffer from dramatic morphological
changes. This implies that the disk is globally stable, a consequence of the fact that, when
Q ∼ 2, the growth of perturbations by swing amplification and other instabilities is quite
mild (Toomre 1981; Sellwood 1989). No additional perturbation from the rigid halo and
bulge also helps to keep the disk featureless (Hernquist 1993).
potential) varies rapidly with radius, rendering the local approximation invalid there (e.g.,
Figure 17 shows the radial distributions of various physical quantities averaged over the
azimuthal direction at t = 0, 1, and 2 Gyrs. While ¯vφ and σR change promptly (within less
than 0.1 Gyr) from the initial profiles, Σd remains almost unchanged. The changes in σR
and Q are largest at R <∼ 5 kpc where the circular velocity (hence the total gravitational
Sellwood 1985). The small increases of σR at R >∼ 10 kpc from the initial values are likely
caused by mild swing amplification. Except the slight variations of σR near the center, the
changes of the disk properties between 1 and 2 Gyrs are practically negligible, indicating
that at late time the disk is in a sufficiently well-relaxed, new equilibrium.
C.
Impulse Approximation
In the absence of tidal perturbations, the motions of individual disk particles are in
general a superposition of the radial oscillations with epicycle frequency κ around their
guiding centers and the circular oscillations of the guiding centers about the disk center. The
dispersion δR in the epicycle amplitudes is related to the radial velocity dispersion through
δR = σR/κ. Tidal perturbations are able to enhance the epicycle amplitudes for particles
whose orbital periods are not so small compared with the duration of a tidal encounter.
Using an impulse approximation, one can estimate δR of disk particles subject to tidal
perturbations. Let us assume that the tidal forcing is applied impulsively near the pericenter
during the time interval of Rperi/vp. Then, the increment ∆vR in the radial velocities of
particles at radius R0 is given approximately by
∆vR =
2GMpR0
vpR2
peri
,
(C1)
where vp = [2G(Mg + Mp)/Rperi]1/2 is the orbital velocity of the perturber at the pericenter
(e.g., Binney & Tremaine 1987). Assuming that the kinetic energy associated with ∆vR is
– 27 –
absorbed into the epicycle motions, one obtains
δR = (σ2
R + ∆v2
R)1/2/κ
(C2)
as a measure of the mean radial excursion of disk particles under the influence of tidal
perturbations. Figure 3 plots as thin curves δR from equations (C1) and (C2) with differing
Mp corresponding to models A2, B2, and C2, while the thick curve draws σR/κ.
Adams, F. C., Ruden, S. P., & Shu, F. H. 1989, ApJ, 347, 959
REFERENCES
Balbus, S. A. 1988, ApJ, 324, 60 (B88)
Barnes, J. 1988, ApJ, 331, 699
Barnes, J. 1992, ApJ, 393, 484
Barnes, J. 1998, in Galaxies: Interactions and Induced Star Formation eds. D. Friedli, L.
Martinet, & D. Pfenniger (Springer: Heidelberg), 275
Barnes, J., & Hernquist, L. 1992, Nature, 360, 715
Barnes, J., & Hernquist, L. 1992, ApJ, 471, 115
Barnes, J., & Hut, P. 1986, Nature, 324, 446
Bertin, G., & Lin, C. C. 1996, Spiral Structure in Galaxies: A Density Wave Theory (Cam-
bridge: MIT Press)
Bertin, G., Lin, C. C., Lowe, S. A., & Thurstans, R. P. 1989a, ApJ, 338, 78
Bertin, G., Lin, C. C., Lowe, S. A., & Thurstans, R. P. 1989b, ApJ, 338, 104
Binney, J. 2001, in ASP Conf. Ser. 230, Galaxy Disks and Disk Galaxies, ed. J. G. Funes &
E. M. Corsini (San Francisco: ASP) 63
Binney, J., & Tremaine, S. 1987, Galactic dynamics (Princeton: Princeton Univ. Press)
Byrd, G. G., & Howard, S. 1992, AJ, 103, 1089
Chen, B., Stoughton, C., Smith, A. et al. 2001, ApJ, 553, 184
Donner, K. J., Engstrom, S., & Sundelius, B. 1991, A&A, 252, 571
– 28 –
Dubinski, J., Mihos, J. C., & Hernquist, L. 1996, ApJ, 462, 576
Durrell, P. R., Mihos, J. C., Feldmeier, J. J., Jacoby, G. H., & Ciardullo, R. 2003, ApJ, 582,
170
Elmegreen, D. M., & Elmegreen, B. G. 1982, MNRAS, 201, 1021
Elmegreen, D. M., & Elmegreen, B. G. 1987, ApJ, 314, 3
Elmegreen, D. M., Sundin, M., Elmegreen, B. G., & Sundelius, B. 1991, A&A, 244, 52
Elmegreen, B. G. 1994, ApJ, 433, 39
Elmegreen, B. G. 1995, in The 7th Guo Shoujing Summer School on Astrophysics: Molecular
Clouds and Star Formation, eds. C. Yuan & Hunhan You (Singapore:World Scientific),
149
Elmegreen, B. G., & Elmegreen, D. M. 1983, MNRAS, 203, 31
Elmegreen, B. G., Elmegreen, D. M., & Seiden, P. E. 1989, ApJ, 602
Elmegreen, B. G., Kaufman, M., & Thomasson, M. 1993, ApJ, 412, 90
Garc´ıa-Burillo, S., Combes, F., & Gerin, M. 1993, A&A, 274, 148
Gerin, M., Combes, F., & Athanassoula, E. 1990, A&A, 230, 37
Goldreich, P., & Lynden-Bell, D. 1965, MNRAS, 130, 125
Hernquist, L. 1990, Dynamics and Interactions of Galaxies, ed. R. Wielen (Berlin: Springer),
108
Hernquist, L. 1993, ApJS, 86, 389
Howard, S., & Byrd, G. G. 1990, AJ, 99, 1798
Julian, W. H., & Toomre, A. 1966, ApJ, 146, 810
Karaali, S., Bilir, S., & Hamzaoglu, E. 2004, MNRAS, 355, 307
Kennicutt, R. 1981, AJ, 86, 1847
Calzetti, D., Kennicutt, R. C. et al. 2005, ApJ, 633, 871
Kim, W.-T., & Ostriker, E. C. 2002, ApJ, 570, 132
– 29 –
Kim, W.-T., & Ostriker, E. C. 2006, ApJ, 646, 213
Kormendy, J., & Norman, C. A. 1979, ApJ, 233, 539
Kuijken, K., & Gilmore, G. 1989, MNRAS, 239, 605
La Vigne, M. A., Vogel, S. N., & Ostriker, E. C. 2006, ApJ, 650, 818
Lee, C. W., Lee, H. M., Ann, H. B., & Kwon, K. H. 1999, ApJ, 513, 242
Lee, M. G., Chandar, R., & Whitmore, B. C. 2005, AJ, 130, 2128
Lin, C. C., & Shu, F. H. 1964, ApJ, 140, 646
Lin, C. C., & Shu, F. H. 1966, Proceedings of the National Academy of Science, 55, 229
Lin, C. C., Yuan, C., & Shu, F. H. 1969, ApJ, 155, 721
McKee, C. F., & Ostriker, E. C. 2007, ARA&A, 45, 565
Noguchi, M. 1987, MNRAS, 228, 635
Ostriker, E. C., Shu, F. H., & Adams, F. C. 1992, ApJ, 399, 192
Patsis, P. A., H´eraudeau, Ph., & Grosbøl, P. 2001, A&A, 370, 875
Pfleiderer, J. 1963, ZAp, 58, 12
Press, W. H., & Teukolsky, S. A. 1977, ApJ, 213, 183
Rand, R. J. 1993, ApJ, 410, 68
Rix, H.-W., & Rieke, M. J. 1993, ApJ, 418, 123
Rix, H.-W., & Zaritsky, D. 1995, ApJ, 447, 82
Roberts, W. W. 1969, ApJ, 158, 123
Rots, A. H., Bosman, A., van der Hulst, J. M., Athanassoula, E., & Crane, P. C. 1990, AJ,
100, 387
Rybicki, G. B. 1971, Ap&SS, 14, 15
Quinn, P. J., Hernquist, L., & Fullagar, D. P. 1993, ApJ, 403, 74
Salo, H., & Laurikainen, E. 1993, ApJ, 410, 586
– 30 –
Salo, H., & Laurikainen, E. 2000a, MNRAS, 319, 377
Salo, H., & Laurikainen, E. 2000b, MNRAS, 319, 393
Scoville, N. Z., Polletta, M., Ewald, S., Stolovy, S. R., Thompson, R., & Rieke, M. 2001, AJ,
112, 3017
Seigar, M. S., Block, D. L., Puerari, I., Chorney, N. E., & James, P. A. 2005, MNRAS, 359,
1065
Seigar, M. S., Bullock, J. S., Barth, A. J., & Ho, L. C. 2006, ApJ, 645, 1012
Sellwood, J. A. 1985, MNRAS, 217, 127
Sellwood, J. A. 1989, in Dynamics of Astrophysical Disks, 155, ed. J. A. Sellwood (Cambridge
Univ. Press), 155
Sellwood, J. A., & Carlberg, R. G. 1984, ApJ, 282, 61
Sellwood, J. A., & Athanassoula, E. 1986, MNRAS, 221, 195
Shetty, R., & Ostriker, E. C. 2006, ApJ, 647, 997
Shetty, R., Vogel, S. N., & Ostriker, E. C. 2007, ApJ, 665, 1138
Shu, F. H., Milione, V., & Roberts, W. W. 1973, ApJ, 183, 819
Smith, J., Gehrz, R. D., Grasdalen, G. L., Hackwell, J. A., Dietz, R. D., & Friedman, S. D.
1990, ApJ, 362, 455
Springel, V., Yoshida, N., & White, S. D. M. 2001, New Astronomy, 6, 79
Sundelius, B., Thomasson, M., Valtonen, M. J., & Byrd, G. G. 1987, A&A, 174, 67
Sundin, M. 1989, in Dynamics of Astrophysical Discs, ed. J. A. Sellwood (Cambridge Univ.
Press: Cambridge), 215
Toomre, A. 1964, ApJ, 139, 1217
Toomre, A. 1969, ApJ, 158, 899
Toomre, A. 1981, in Structure and Evolution of Normal Galaxies, eds. S. M. Fall & D.
Lynden-Bell (Cambridge:Cambridge Univ. Press), 111
Toomre, A., & Toomre, J. 1972, ApJ, 178, 623
– 31 –
Tremaine, S., & Weinberg, M. D. 1984, ApJ, 282, L5
Vandervoort, P. O. 1971, ApJ, 166, 37
Vogel, S. N., Kulkarni, S. R., & Scoville, N. Z. 1988, Nature, 334, 402
Wetzstein, M., Naab, T., & Burkert, A. 2007, MNRAS, 375, 805
White, R. L. 1988, ApJ, 330, 26
Willner, S. P., et al. 2004, ApJS, 154, 222
Woodward, P. R. 1975, ApJ, 195, 61
Zimmer, P., Rand, R. J., McGraw, J. T. 2004, ApJ, 607, 285
This preprint was prepared with the AAS LATEX macros v5.2.
Table 1. Summary of model parameters and simulation results
Modela
(1)
Mp/Mg
(2)
Rperi(kpc)
(3)
A1
A2
A2*
A3
B1
B2
B3
C1
C2
C3
0.4
0.4
0.4
0.4
0.2
0.2
0.2
0.1
0.1
0.1
25
35
35
45
25
35
45
25
35
45
Sb
(4)
0.250
0.151
0.151
0.103
0.135
0.081
0.056
0.070
0.043
0.029
Mcap/Md
(5)
c (%)
Mesc/Md
(6)
d (%)
4.99
1.99
1.84
0.46
1.95
0.79
0.07
0.73
0.22
· · ·
1.97
0.08
0.08
0.006
1.33
0.03
0.002
0.67
0.01
· · ·
ttail
(7)
0.14
0.18
0.19
0.25
0.18
0.24
· · ·
· · ·
· · ·
· · ·
tan itail
(8)
0.546
0.511
0.707
0.499
0.457
0.411
· · ·
· · ·
· · ·
· · ·
Σtail/Σ20
e
(9)
29.2
22.0
11.6
15.3
17.4
13.4
· · ·
· · ·
· · ·
· · ·
Fmax
(10)
0.22
0.18
0.11
0.09
0.17
0.10
0.06
0.11
0.06
0.04
aModel A2* is identical to model A2 except that the former neglects the self-gravity of the perturbed density.
bS ≡ (Mp/Mg)(Rg /Rperi)3(∆T /T ) = 0.738(Mp/Mg)(Rg /Rp)3/2[1 + (Mp/Mg)]−1/2 is the dimensionless tidal strength parameter.
cMcap is the total mass of the captured particles by the perturbing companion.
dMesc is the total mass of the non-captured, escaping particles from the whole system.
eΣ20 = Σd(20 kpc) is the surface density of the initial unperturbed disk at R = 20 kpc.
–
3
2
–
– 33 –
Fig. 1.— Snapshots of the particle distributions in model A2 in the x-y plane. Only 20% of
the particles are shown to reduce crowding. The elapsed time is shown at the upper right
corner of each panel. The arrow and the associated number give the direction and distance
(in kpc) to the perturber that passes through the pericenter (x, y)=(35 kpc, 0) at t = 0 in
the counterclockwise direction. See text for details.
– 34 –
Fig. 2.— Distributions of the perturbed surface density δΣ/Σ of model A2 in the polar
coordinates. When t <∼ 0.2, δΣ is dominated by the extended-disk structures such as bridge
and tail at R >∼ 15 kpc, which become loose and spread widely after t ∼ 0.3. The spiral arms
at R <∼ 15 kpc appear straight in the φ − ln R plane, with a slope becoming progressively
smaller with time.
– 35 –
Fig. 3.— Dispersions in the radial departures of particles from the initial locations over the
course of tidal interactions. Various symbols indicate the numerical results of models A2, B2,
and C2. The thin solid, dotted, dashed curves draw the analytic estimates (eq. [C2]) based
on the impulse approximation. The thick solid line corresponds to the unperturbed disk in
which δR is purely due to the epicycle orbits associated with the initial velocity dispersions.
– 36 –
Fig. 4.— Azimuthal variations of the radial (vR) and azimuthal (vφ) velocities of the particles
at R = 20 kpc in model A2 at early epochs of tidal interactions. The azimuthal phase in the
abscissa is repeated for clarity. In each panel, vertical dotted lines mark the phase angles of
the perturber.
– 37 –
Fig. 5.— Spatial distributions of selected particles in model A2 at some time epochs when
the perturber is close to the pericenter. Blue, green, yellow, orange, and red colors represent
the particles originally located in annuli with R0 = 6 − 7, 10 − 11, 14 − 15, 18 − 20, and
22− 24 kpc, respectively. The black circle in each panel has a radius of 20 kpc and the arrow
indicates the direction to the perturber. The black and cyan dots denote the groups of the
near-side and far-side particles, respectively, in the R0 = 22 − 24 kpc ring, which merge at
the far side to form a tail at t = 0.2.
– 38 –
Fig. 6.— Dependences on the the tidal strength parameter S of the formation epoch ttail,
the pitch angle itail at t = ttail, and the surface density Σtail at R = 20 kpc and t = ttail of
tidal tails in various models. The dotted line in each panel gives the best fit to the numerical
results: ttail = 0.07S −0.54, tan itail = 0.75S0.22, and Σtail/Σ20 = 79.0S0.72, where Σ20 denotes
the surface density of the initial disk at R = 20 kpc.
– 39 –
Fig. 7.— Close-up views of the distributions of stellar surface density in model A2, with
the color bar labeling (Σ/104 M⊙ pc−2)1/2. The two-armed spiral density waves excited by
the tidal forcing achieve the maximum strength at t ∼ 0.3, and then gradually weaken.
– 40 –
Fig. 8.— Evolution of the Fourier amplitudes with wavenumber p, defined by equation (4), of
the m = 2 logarithmic spirals in model A2. The modal growth is due to swing amplification
at early time (t <∼ 0.04), which becomes soon dominated by the kinematic overlapping of the
perturbed epicycle orbits.
– 41 –
Fig. 9.— Temporal changes of the pitch angle of the spiral arms located in the R = 5 − 10,
8.0− 8.5, 8− 13, and 11− 16 kpc regions for model A2, A2*, B2, and C2, respectively. In the
self-gravitating models A2, B2, and C2, the pitch angle deceases as tan i ∝ t−0.5∼−0.6, with
weaker arms decaying slightly more rapidly, whereas tan i ∝ t−1 for the non-self-gravitating
model A2*.
– 42 –
Fig. 10.— Azimuthal distributions of the perturbed density δΣ (top) as black curves and
the radial velocity vR (middle) and perturbed circular velocity δvφ = vφ − ¯vφ (bottom) as
dots in the disk at R = 10 kpc when t = 0.4 for model A2 (left) and when t = 0.5 for model
C2 (right). Note that the vertical scales are different from the left and right panels. In the
top panels, the red lines give the m = 2 Fourier modes δΣm=2 of the perturbed density. In
the middle and bottom panels, the blue curves give the average values of vR and δvφ, while
the red curves plot the predictions of the linear density wave theory corresponding to δΣm=2.
– 43 –
Fig. 11.— Arm strength F averaged over the time interval ∆t = 0.4 centered at the time of
the peak strength as a function of radius. The induced spiral arms for stronger encounter
models A1, A2, B1, and C1 peak at Rmax ∼ 5 − 10 kpc, while weaker encounter models B3,
C2, and C3 produce spiral arms at Rmax ∼ 11 − 16 kpc.
– 44 –
Fig. 12.— Time evolution of the arm strength F averaged over 5 kpc <∼ R <∼ 10 kpc for
models A1, A2, B1, and C1, over 8 kpc <∼ R <∼ 13 kpc for models A2*, A3 and B2, and over
11 kpc <∼ R <∼ 16 kpc for models B3, C2, and C3. In each self-gravitating model, it takes
about one or two rotational periods at R = Rmax for the arms to reach the maximum value.
After the peak, F decays as ∼ exp(−t/1 Gyr) due largely to large dispersions in the particle
velocities. Fluctuations of F at t >∼ 1 arise as the particles once pertaining to the bridge and
tail move in and out the arms. Without disk heating, the decay of the spiral arms in the
non-self-gravitating model A2* is quite slow.
– 45 –
Fig. 13.— Dependence of the peak arm strength Fmax on S. The dotted line Fmax = 0.95S0.86
is the best fit to our numerical results.
– 46 –
Fig. 14.— Contours of the cross correlation of the normalized surface density in the radius
– frequency domain for models A2 (left) and A2* (right). Smooth curves draw Ω (solid )
and Ω ± κ/2 (dashed ) from the initial disk rotation. The dotted line in each of the left
panels plots the theoretical patten speed calculated from the linear dispersion relation for
the background parameters equal to the azimuthally-averaged disk values obtained from the
simulation. Nearly constant Ωp ∼ 9.5 km s−1 kpc−1 at R >∼ 17 kpc for t <∼ 0.2 traces the
tidal bridge and tail, which instantaneously corotate with the perturber. The loci of the
maximum cross correlation for the arms match well with the Ω − κ/2 curve in model A2 for
t >∼ 0.6 and in model A2* for t >∼ 0.3.
– 47 –
Fig. 15.— Circular speed vc(R) (solid ) and mean streaming velocity ¯vφ(R) (dotted ) of our
model galaxy as functions of the galactocentric radius R. Contributions to vc from disk,
bulge, and halo are plotted as dashed lines.
– 48 –
Fig. 16.— Evolution of our model disk in isolation. Time is expressed in units of 109 yr. Only
104 particles are plotted to reduce crowding. The disk is rotating in the counterclockwise
direction. No notable change in appearance is found, indicating that the disk is globally
stable.
– 49 –
Fig. 17.— Radial distributions of (a) the disk surface density Σd, (b) mean streaming
velocity ¯vφ, (c) radial velocity dispersion σR and (d ) Toomre stability parameter Q from an
isolated disk evolution at t = 0 (dotted ), 1 Gyr (solid ), and 2 Gyr (dashed ).
|
astro-ph/0510347 | 2 | 0510 | 2006-04-14T05:27:43 | Ly-alpha Radiative Transfer in Cosmological Simulations and Application to a z~8 Emitter | [
"astro-ph"
] | We develop a Ly-alpha radiative transfer (RT) Monte Carlo code for cosmological simulations.High resolution,along with appropriately treated cooling can result in simulated environments with very high optical depths.Thus,solving the Ly-alpha RT problem in cosmological simulations can take an unrealistically long time.For this reason,we develop methods to speed up the Ly-alpha RT.With these accelerating methods,along with the parallelization of the code,we make the problem of Ly-alpha RT in the complex environments of cosmological simulations tractable.We test the RT code against simple Ly-alpha emitter models,and then we apply it to the brightest Ly-alpha emitter of a gasdynamics+N-body Adaptive Refinement Tree (ART) simulation at z~8.We find that recombination rather than cooling radiation Ly-alpha photons is the dominant contribution to the intrinsic Ly-alpha luminosity of the emitter,which is ~4.8x10e43 ergs/s.The size of the emitter is pretty small,making it unresolved for currently available instruments.Its spectrum before adding the Ly-alpha Gunn-Peterson absorption (GP) resembles that of static media,despite some net inward radial peculiar motion.This is because for such high optical depths as those in ART simulations,velocities of order some hundreds km/s are not important.We add the GP in two ways.First we assume no damping wing,corresponding to the situation where the emitter lies within the HII region of a very bright quasar,and second we allow for the damping wing.Including the damping wing leads to a maximum line brightness suppression by roughly a factor of ~62.The line fluxes,even though quite faint for current ground-based telescopes,should be within reach for JWST. | astro-ph | astro-ph | draft version; September 23, 2018
Preprint typeset using LATEX style emulateapj v. 11/12/01
6
0
0
2
r
p
A
4
1
2
v
7
4
3
0
1
5
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
Ly-α RADIATIVE TRANSFER IN COSMOLOGICAL SIMULATIONS AND APPLICATION TO A
z ≃ 8 Ly-α EMITTER
Argyro Tasitsiomi1,2
draft version; September 23, 2018
ABSTRACT
We develop a Ly-α radiative transfer (RT) Monte Carlo code for cosmological simulations. High res-
olution, along with appropriately treated cooling, can result in simulated environments with very high
optical depths. Thus, solving the Ly-α RT problem in cosmological simulations can take an unrealisti-
cally long time. For this reason, we develop methods to speed up the Ly-α RT. With these accelerating
methods, along with the parallelization of the code, we make the problem of Ly-α RT in the complex
environments of cosmological simulations tractable. We test the RT code against simple Ly-α emitter
models, and then we apply it to the brightest Ly-α emitter of a gasdynamics+N-body Adaptive Refine-
ment Tree (ART) simulation at z ≃ 8. We find that recombination rather than cooling radiation Ly-α
photons is the dominant contribution to the intrinsic Ly-α luminosity of the emitter, which is ≃ 4.8×1043
ergs/s. The size of the emitter is pretty small, making it unresolved for currently available instruments.
Its spectrum before adding the Ly-α Gunn-Peterson absorption (GP) resembles that of static media,
despite some net inward radial peculiar motion. This is because for such high optical depths as those
in ART simulations, velocities of order some hundreds km/s are not important. We add the GP in two
ways. First we assume no damping wing, corresponding to the situation where the emitter lies within the
HII region of a very bright quasar, and second we allow for the damping wing. Including the damping
wing leads to a maximum line brightness suppression by roughly a factor of ∼ 62. The line fluxes, even
though quite faint for current ground-based telescopes, should be within reach for JWST.
Subject headings: cosmology: theory -- diffuse radiation -- galaxies: formation -- intergalactic
medium -- radiative transfer -- line: formation -- radiative transfer -- resonant --
polarization
1.
introduction
Since the classic paper by Partridge & Peebles (1967),
intense observational efforts have focused on the search
for Ly-α emitters at high redshifts. Although most of
the early attempts ended in negative results before the
mid 1990s, recent observational advances enabled us to
identify star forming galaxies at ever increasing redshifts.
Currently, several observational projects, such as LALA
(e.g., Rhoads et al. 2003), CADIS (e.g., Maier 2002), the
Subaru Deep Field Project (e.g., Taniguchi et al. 2005),
etc., spectroscopic surveys that use lensing magnification
from clusters (e.g., Santos et al. 2003), surveys that com-
bine Subaru (e.g., Hu et al. 2004) or HST/ACS/NICMOS
imaging (e.g., Stanway et al. 2004; Dickinson et al. 2004,
for the GOODS survey) with Keck spectroscopy, etc., focus
on finding high-z starforming galaxies. Surveys currently
reach up to z ≃ 7− 8 (e.g., see Bouwens et al. 2004, for re-
cent results from the NICMOS observations of the HUDF)
and will likely reach higher redshifts in the coming years
(e.g., via JWST).
The hydrogen Ly-α line is a very promising way to probe
the high-redshift universe. Besides yielding redshifts, the
shape, equivalent width and offset of the Ly-α line from
other emission/absorption lines potentially convey valu-
able information about the geometry, kinematics, and un-
derlying stellar population of the host galaxy. Further-
more, after escaping the environment of the host galaxy,
1 Dept. of Astronomy and Astrophysics, Kavli Institute for Cosmo-
logical Physics, The University of Chicago, Chicago, IL 60637
2 Current address: Dept.
of Astrophysical Sciences, Prince-
ton University, Peyton Hall-Ivy Lane, Princeton, NJ 08544;
[email protected]
Ly-α photons are scattered in the surrounding IGM. The
presence or absence of observed Ly-α emission can be used
to place constraints on the state of the IGM, useful in con-
straining for example the epoch and topology of reioniza-
tion. Because of the numerous factors that contribute to
the final Ly-α emission, the interpretation of such features
can be very complex. To use all the currently available
and future observations in the most effective way possi-
ble we need to improve our theoretical understanding of
Ly-α emission from high-redshift objects. To this end we
develop a general Ly-α radiative transfer (RT) scheme for
cosmological simulations. As an example, we apply the
RT scheme to gasdynamics+N-body Adaptive Refinement
Tree (ART; Kravtsov 2003) simulations of galaxy forma-
tion.
There are quite a few studies of Ly-α emission from
high-z objects (e.g., Gould & Weinberg 1996; Haiman &
Spaans 1999; Loeb & Rybicki 1999; Ahn et al. 2001, 2002;
Zheng & Miralda-Escud´e 2002; Santos 2004; Dijkstra et al.
2005a,b). The problem these studies address is highly
complex and has many unknowns.
Inevitably, most of
these studies had to make at some point some simplify-
ing assumptions. Usually, a high degree of symmetry for
the emitting source, and its density, temperature, and ve-
locity field is assumed. The same is the case with respect
to the processes that are responsible for the production of
Ly-α photons. On the other hand, cosmological simula-
tions hopefully capture most of the basic elements, lifting
thus practical constraints that existed in these previous
studies.
There is a small number of related studies using cos-
mological simulations (Fardal et al. 2001; Furlanetto et al.
2003; Barton et al. 2004; Furlanetto et al. 2005; Cantalupo
1
2
TASITSIOMI
et al. 2005; Delliou et al. 2005a,b). Some of these simula-
tions are lacking crucial processes such as radiative cooling
of the gas and consistent RT, the various sources of Ly-α
photons, and/or sufficient resolution in order to resolve the
clumpiness of the gas. Furthermore, most of these studies
do not perform Ly-α RT, but rather they assume that the
observer sees whatever is being emitted initially, simply
modified by e−τ with τ the optical depth for Ly-α scat-
tering due to neutral hydrogen between the emission point
and the observer. Namely, in most cases Ly-α spectra from
simulations are treated as absorption spectra when, in real-
ity, they are scattering spectra (see, e.g., Gnedin & Prada
2004). For gas well outside the source of emission this
is an appropriate approximation since scattering off the
direction of viewing removes the photons that could be
observed and thus appears as effective absorption. This
is no longer true for the source of emission itself, since
photons that were originally emitted in directions differ-
ent from the direction of observation may scatter into this
direction.
It is important that the difficulty of implementing a Ly-
α RT scheme for cosmological simulations become clear.
The classical problem of resonance RT, relevant to a wide
range of applications from planetary atmospheres to ac-
cretion disks, has been quite extensively studied in the lit-
erature (e.g., Zanstra 1949; Hummer 1962; Auer 1968; Av-
ery & House 1968; Adams 1972; Harrington 1973; Neufeld
1990; Ahn et al. 2001, 2002). However, analytical solu-
tions derived in the past are applicable only to certain
specific conditions. On the other hand, the slow conver-
gence of the numerical techniques used limited the nu-
merical studies at optical thicknesses that are relatively
low compared to those encountered in high-redshift galax-
ies (and cosmological simulations of high-redshift galaxies,
as we will show). Thus, unlike previous studies, most of
which focused on the classical problem of resonant RT in
a semi-infinite slab, in cosmological simulations one has to
solve simultaneously thousands or even millions of these
problems.3Furthermore, having in mind existing and fu-
ture cosmological simulations that can achieve sufficiently
high resolution to resolve the gas clumpiness and that treat
cooling appropriately, we anticipate column densities that
are orders of magnitude higher than those found in lower
resolution simulations without cooling.
In this case we
need a RT algorithm much faster than the more standard
direct Monte Carlo approach [which, however, is our start-
ing point] of previous studies. Thus, we must develop RT
acceleration methods that, along with the highly parallel
nature of the RT problem that enables us to make use of
many parallel machines, can make the Ly-α RT problem
tractable.
The paper is organized as follows. In § 2 we discuss the
RT scheme. More specifically, in § 2.1 we present the ba-
sic Monte Carlo algorithm, in § 2.2 we present tests of the
basic algorithm, in § 2.3 we discuss the acceleration meth-
ods we use to speed up the RT, and in § 2.4 we present
the method images and spectra are constructed.
In § 3
we discuss in detail an application of the Ly-α RT code to
ART simulations. More specifically, in § 3.1 we briefly give
3 For example, in the case of Adaptive Mesh Refinement (AMR)
codes, each time a photon enters a simulation cell one has the equiv-
alent of a new slab RT problem.
some information about the ART simulations. In § 3.2 we
discuss the intrinsic Ly-α emission of the specific simu-
lated Ly-α emitter we focus on. In § 3.3 we present results
on the emitter before RT. In § 3.4 we discuss results after
performing the RT, and with/without the Gunn-Peterson
(GP) absorption, as well as with/without the red damp-
ing wing of the GP absorption.
In § 4 we discuss and
summarize our results and conclusions.
2. the Ly-α radiative transfer
2.1. The basic Monte Carlo code
The following discussion assumes in various places a cell
structure for the simulation outputs, as is inherently the
case in AMR codes. However, the Ly-α RT code we dis-
cuss is applicable to outputs from all kinds of cosmological
simulation codes, since one can always create an effective
mesh by interpolating the values of the various physical
parameters. The size of the mesh cell can be motivated by
resolution related scales (e.g., the softening scale, or larger
if convergence tests with respect to the Ly-α RT justify a
larger scale). Thus, in what follows we refer to simulation
cells either the direct output of the cosmological simula-
tion code has a cell structure or not.
The initial emission characteristics (simulation cell, fre-
quency, etc.) of each photon depend on the specific phys-
ical conditions, thus we defer this discussion for §3 where
an application to a Ly-α emitter produced in ART cos-
mological simulations is presented. After determining the
initial characteristics for each photon, we follow a series
of scatterings up to a certain scale where the detailed RT
stops. This scale is to be determined via a convergence
study.
In this subsection we describe the basic steps of
the algorithm.
2.1.1. Propagating the photon
For every scattering we generate the optical depth, which
determines the spatial displacement of the photon, by sam-
pling the probability distribution function e−τ
(1)
with R a uniformly distributed random number. This op-
tical depth is equal to
τ = − ln(R) ,
τ =
l
Z0
∞
Z−∞
dldupσL(ν(1−up/c))r mp
2πkBT
nHI exp −
mpu2
p
2kBT! ,
(2)
with nHI the number density of neutral hydrogen. The
function σL is the scattering cross section of Ly-α photons
as a function of frequency, defined in the rest frame of the
hydrogen atom as
σL(ν) = f12
πe2
mec
∆νL/2π
(ν − ν0)2 + (∆νL/2)2 ,
(3)
where f12 = 0.4162 is the Ly-α oscillator strength, ν0 =
2.466×1015 Hz is the line center frequency, ∆νL = 9.936×
107 Hz is the natural width of the line, and other symbols
have their usual meaning. In equation (2) the fact that the
photons are encountering atoms with a Maxwellian distri-
bution of thermal velocities has been taken into account.
Integrating over the distribution of velocities, the resulting
cross section in the observer's frame is
σ(x) = f12
H(α, x)
(4)
√πe2
mec∆νD
Ly-α radiative transfer in cosmological simulations
3
where
H(α, x) =
α
π Z ∞
−∞
e−y2
(x − y)2 + α2 dy
(5)
is the Voigt function, x = (ν − ν0)/∆νD is the relative
frequency of the incident photon in the observer's frame
with ∆νD = p2kBT /(mpc2)ν0 the Doppler width, and
α = ∆νL/2∆νD with ∆νL the natural line width. Assum-
ing that σ is independent of l, the optical depth is given
by
τ = nHI σ(x)l .
(6)
When applied to cosmological simulations, equation (6) is
substituted by a sum of terms similar to the r.h.s. This
sum is over the different cells (=different physical condi-
tions such as neutral hydrogen density, temperature, etc.)
that the photon crosses until it reaches τ and gets scat-
tered.
For the Voigt function we use the following analytic fit,
which is a good approximation to better than 1% for tem-
peratures T > 2K (N. Gnedin, personal communication)
V (α, ν) ≡
=
1
√π∆νD
∆νD (cid:20)q +
1
H(α, x) =
1
∆νD
φ(x)
e−x
1.77245385(cid:21)
where x = x2, and q = 0 if z = (x − 0.855)/(x + 3.42) ≤ 0
and
(7)
(8)
α
21
q = z(cid:18)1 +
× {0.1117 + z [4.421 + z(−9.207 + 5.674z)]}
x (cid:19)
π(x + 1)
if z > 0. The definition in terms of the function φ(x) is
also given since the latter has been used in many previous
studies, and we also use it in what follows. If in addition to
the thermal motion of the atoms there is bulk motion, such
as peculiar or Hubble flow velocities, in equation (4) we use
xf = x − (vf z/c)ν0/∆νD, where vf z is the component of
the fluid bulk velocity along the direction of the incident
photon.
In equation (2) the cross section σ becomes l-dependent
when Hubble expansion is taken into account. In this case
the equation is an integral and does not reduce to the sim-
ple algebraic equation (6). To propagate the photon one
must solve for the step which is the upper limit of the in-
tegral.
In the simple examples discussed in §2.2, things
are relatively simple even when the Hubble expansion is
included, since in these cases there is homogeneity and
isothermality and no sum over cells is required. In those
cases, Hubble expansion is included as follows: 1.) we
make a first guess for l using the Hubble velocity at the
current point, 2.) we use as a step for the photon a certain
fraction of l, 3.) for a specified tolerance with which we
want to achieve τ , we refine the step as necessary. Note
that the simple tests of the code presented in §2.2 do not
In the actual simulations the
include peculiar motions.
peculiar velocities rather than the Hubble expansion are
dominant on the relevant scales (e.g., for the emitter we
focus on, the mean radial component of the peculiar mo-
tion dominates over the Hubble expansion up to about 80
physical kpc). In the detailed RT which we perform within
such distances, we approximate the subdominant Hubble
expansion velocity within a certain cell by the expansion
velocity that corresponds to the center of that cell. This
is calculated to have a negligible effect on the results.
The n = 2 state of atomic hydrogen consists of the
2S1/2, 2P1/2 and 2P3/2 substates, whereas the n = 1 state
consists of 1S1/2. According to the electric dipole selec-
tion rules, the allowed transitions are 2P1/2 to 1S1/2 and
2P3/2 to 1S1/2, whereas 2S1/2 corresponds to destruction
of the initially absorbed Ly-α photon, since this state de-
excites through the emission of two continuum photons.
The multiplicity of each of these states is 2J + 1. Thus
the probabilities for the 2P states the atom can be found
in when absorbing the Ly-α photons 2P1/2 : 2P3/2 are
1 : 2. Collisions can potentially cause the 2P → 2S tran-
sition in which case the photon gets destroyed. A similar
destruction effect can be caused by the existence of dust.
Both these destruction mechanisms are briefly discussed in
the context of the ART simulations in §3.5.1 and §3.5.2,
respectively. Considering the 2P1/2 and 2P3/2 cases sepa-
rately, one would have to modify both the Voigt function
and the velocity distribution of the scattering atom dis-
cussed in the next section (see, e.g., Ahn et al. 2001).
However, the level splitting between the two 2P states is
small, just 10 GHz. This corresponds to a velocity width
of ∼ 1 km/s, much smaller than the width due to ther-
mal velocities in media with roughly T > 100 K. In addi-
tion, even for lower temperatures, this level splitting is still
small for high optical depths. In our case, the thermal, pe-
culiar, and Hubble velocities are all more important than
the splitting, and combined with the fact that we have high
optical depths, we do not make the distinction between the
two sublevels. As discussed below, however, the different
fine structure levels are taken into account when choos-
ing scattering phase functions, important for polarization
calculations that we will present in a future paper.
2.1.2. The scattering
After determining the point in space where the pho-
ton will be scattered next, we choose the thermal velocity
components of the scattering atom. In the two directions
perpendicular to the direction of the incident photon the
components are drawn from a (1-D) Gaussian distribution
. The component up of the
thermal velocity of the atom along the direction of the
incident photon is drawn from the distribution
with dispersion equal to q kB T
mp
f (vp) =
a
π
e−v2
p
(x − vp)2 + a2 H−1(a, x) ,
(9)
with vp = up(mp/2kT )1/2. To draw numbers that follow
this distribution we use the method of Zheng & Miralda-
Escud´e (2002).
After each scattering we need to assign a new frequency
(in the observer's frame) and direction to the photon. To
this end we perform a Lorentz transformation of the fre-
quency and direction of the incident photon from the ob-
server to the atom rest frame, using the velocity of the
atom chosen as described previously.
Although the code ignores the level splitting with re-
spect to the scattering cross section and the velocity dis-
tribution, it takes into account the different phase distribu-
tions for core versus wing scatterings, as well as for 2P1/2
4
TASITSIOMI
versus 2P3/2 scatterings. For resonant scattering, it is the
angular momenta of the three states involved and the mul-
tipole order of the emitted radiation that determines the
scattering phase function. Hamilton (1940) found that
the transition from 2P1/2 gives totally unpolarized photons
and is characterized by an isotropic angular distribution
function, whereas that from the 2P3/2 state corresponds
to a maximum degree of polarization of 3/7 for a 90◦ scat-
tering (also see Chandrasekhar 1960). More specifically,
the scattering phase function for dipole transition can be
written as (Hamilton 1940)
W (θ) ∝ 1 +
R
Q
cos2 θ
(10)
with R/Q the degree of polarization for a 90◦ scattering
and equal to
R/Q =
(J + 1)(2J + 3)
26J 2 − 15J − 1
(11)
for the 2P3/2 → 1S1/2 transition since ∆J = −1, ∆j =
1, J = 3/2 according to Hamilton's conventions, and
R/Q =
(2J − 1)(2J + 3)
12J 2 + 12J + 1
(12)
for the 2P1/2 → 1S1/2 transition with ∆J = ∆j = 0, J =
1/2. In both equations, J is the total angular momentum
at the excited (n = 2) state. Thus, W (θ) is constant
(isotropic) for 2P1/2 as the excited state, whereas it equals
(13)
with maximum polarization degree of 3/7 at a 90◦ scat-
tering.
W (θ) ∝ 1 + 3/7 cos2 θ
On the other hand, Stenflo (1980) showed that at high
frequency shifts (i.e., at the line wings) quantum mechani-
cal interference between the two lines acts in such a way as
to give a scattering behavior identical to that of a classical
oscillator, namely pure Rayleigh scattering. Then the di-
rection follows a dipole angular distribution with Rayleigh
polarization 100% at 900 scattering, namely
W (θ) ∝ 1 + cos2 θ .
(14)
Lastly, the frequency of the photon before and after scat-
tering in the rest frame of the atom differs only by the
recoil effect. Hence,
ν
ν =
1 + hν
mpc2 (1 − cos θ)
(15)
where ν, ν are the frequency of the incident and scattered
photon in the atom rest frame, respectively, the latter
modified due to the recoil effect. This effect is negligible
for the environments produced in the simulations.
After determining the new direction and frequency of
the scattered photon in the atom's rest frame we trans-
form back to the observer's frame, and repeat the whole
scattering procedure.
2.2. Testing the basic scheme
Here we present some of the tests of the RT code we
performed against analytical solutions, as well as other nu-
merical results that exist in the literature. In addition to
showing the good performance of the code, these tests are
presented here as relevant to either Ly-α emitters and/or
the way we accelerate the code when applied in cosmolog-
ical simulations (see §2.3).
2.2.1. Neufeld (1990) test
Neufeld (1990) derived an analytic solution in the limit
of large optical depth for a source radiating resonance line
photons in a thick, plane-parallel, isothermal semi-infinite
slab of uniform density. The analytic emergent spectrum
as a function of frequency shift for a midplane source is
√6
24
x2
ατ0
1
cosh[(π4/54)0.5(x3 − x3
J(±τ0, x) =
i /ατ0)]
with x = (ν − ν0)/∆νD, ∆νD = ν0p2kBT /(mpc2) the
thermal Doppler width, xi the injection frequency shift
(zero for injection at line center), and α the ratio of the
natural to two times the thermal Doppler width. The
quantity τ0 is the optical depth from midplane to one
boundary of the slab at the line center.4
(16)
This analytical solution is valid in the very optically
thick limit, with the latter being defined according to Neufeld
(1990) as τ0 ≥ 103/(√πα). This corresponds to τ0 ≥
3.8 × 104 approximately for a temperature T=10 K as-
sumed in the tests we present here. In deriving equation
(16) the scattering was assumed to be isotropic.
In ad-
dition, it was assumed coherence in the rest frame of the
atom, an assumption that makes the solution valid at the
low density limit only, as well as approximations under
the assumption that wing scatterings dominate were done
(hence the solution is valid at high optical depths). Fur-
thermore, note that the classical slab problem is indepen-
dent of the real size of the slab (all quantities depend on
l/l0 with l0 the actual size of the finite dimension). Lastly,
for this solution it is assumed that the source has unit
strength and is isotropic, namely it emits 1 photon per
unit time or 1/4π photons per unit time and steradian.
For center-of-line injection frequencies the emerging spec-
trum has maximum at x ≃ ±0.88119(α√πτ0)1/3, and an
average number of scatterings N ≃ 0.909316√πτ0 (Har-
rington 1973, with τ0 in these expressions defined using
our conventions rather than Neufeld's). This scaling of
the mean number of scatterings with optical depth in the
case of resonant-line RT in extremely optically thick me-
dia was first explained by Adams (1972), who understood
that photons escape the medium after a series of excur-
sions to the wings. Before this study it was believed that
the number of scatterings scales with τ 2
0 , as would be pre-
dicted by plain spatial random walk arguments (Oster-
brock 1962). We briefly review the interpretation given
by Adams (1972) with respect to the linear scaling of the
mean number of scatterings with optical depth, since we
refer to it extensively in the following sections.
The mean number of scatterings is the inverse of the
escape probability per scattering. The escape probability
per scattering is the integral of the probability per scat-
tering that a photon is scattered beyond certain frequency
shift x∗. Adams (1972) identified this frequency as the fre-
4 Neufeld's definition, used in equation (16), is such that the opti-
cal depth at frequency shift x is given as τx = τ0φ(x). Note that
throughout this section, with the exception of equation (16), our
definition of τ0 is such that the optical depth at frequency shift x is
given as τx = τ0H(α, x). This definition was chosen following recent
studies (e.g. Ahn et al. 2002; Zheng & Miralda-Escud´e 2002) so that
comparisons with these studies be easier. Since φ(x) = H(α, x)/√π,
our τ0 is smaller than Neufeld's by a factor of √π. Note though that
in the following sections we return to the Neufeld (1990) definition
of τ0.
Ly-α radiative transfer in cosmological simulations
5
Fig. 1. -- Left panel: Emergent spectra from the Monte Carlo RT (solid histograms) and as predicted analytically by Neufeld (1990) (dotted
lines) for 3 different center -- of -- line optical depths. The agreement between Monte Carlo and analytical result becomes better with increasing
optical depth, as expected since the analytical solution is valid for very optically thick media. Right panel: The same as in left panel but in
this case the Monte Carlo results are derived with recoil being included, whereas the analytic solution does not include recoil. The dashed
line in the case of τ0 = 106 is obtained by modifying the spectrum obtained from the Monte Carlo RT without recoil by the factor correcting
for recoil (see text for details).
′
quency where the photon, while performing an excursion
to the wings, and before returning back to the core, trav-
els an rms distance comparable to the size of the medium.
Note that this is in fact an essential difference in the un-
derstanding of resonant-line RT in extremely thick media
compared to the spatial random walk approach. The latter
approach assumes that during an excursion to the wings
the photon travels an rms distance much smaller than the
size of the medium. Thus, the first step is to determine
x∗. Using the redistribution function (i.e., the function
that gives the probability that a photon with certain fre-
quency shift x before scattering will have a frequency shift
after scattering) one can calculate both the rms fre-
x
quency shift and the mean frequency shift of a photon
which is scattered repeatedly. For a photon initially in the
wings with a frequency shift x Osterbrock (1962) found
that the rms shift is 1 and the mean frequency shift is
−1/x. For x ≫ 1, the mean shift is much smaller than
the rms and the photon is undergoing a random walk in
frequency with mean number of scatterings ∼ x2. In real
space, the rms distance traveled is equal to the square
root of the mean number of scatterings times the mean
free path. In the wings, the Voigt profile varies relatively
slowly and the mean free path is ∼ 1/φ(x) ∼ x2/α line
center optical depths (we only focus on the scalings here,
hence constants of order unity are dropped). Thus, the
distance traveled is x/φ(x) ∼ x3/α. Setting this rms dis-
tance equal to τ0 we get x∗ ∼ (ατ0)1/3, which is in fact
the scaling of the frequency shift where the emergent spec-
trum takes its maximum value. Thus, going back to the
mean number of scatterings, the escape probability per
A(x)dx with A(x) a function to
scattering will be ∼ R ∞
x∗
be determined. According to the previous discussion x∗ is
the minimum frequency shift for which the photon during
an excursion to the wings can travel an rms distance at
least equal to the size of the medium. If, for simplicity,
one assumes complete redistribution the probability that
a photon is found after scattering with a shift between
x and x + dx is φ(x)dx. However, this is not the prob-
ability per scattering, since the photon will scatter ∼ x2
before returning to the core. Thus, A(x) is φ(x)/x2, and
Nsc ∼hR ∞
, with φ(x) ∼ α/x2 in the wings.
Using the above expression for x∗ one obtains Nsc ∼ τ0,
with the constant of proportionality being of the order of
unity.
φ(x)/x2dxi−1
x∗
The emerging spectra without and with recoil included
are shown in the left and right panel of Figure 1, respec-
tively. A convergence test indicates that these results are
robust if more than of order 103 photons are used. Refer-
ring to the left panel of the figure, the agreement between
the results obtained with the code and the analytic so-
lution gets better at higher optical depths. As has been
already mentioned, the analytic solution is derived after a
series of approximations done on the assumption of opti-
cally thick media. For example, when deriving the analytic
solution the Voigt function is set equal to α/πx2. Set-
ting the Voigt function equal to this approximation in the
code makes the agreement even better. The way the spec-
trum behaves for different ατ0 is expected qualitatively:
the higher the optical depth or the lower the temperature
(the higher the α), the more difficult it is for the pho-
tons to exit the medium and the photon frequencies must
move further away from resonance to escape. Hence, the
6
TASITSIOMI
peaks of the emerging spectrum occur at higher frequency
shifts, and the separation between the two peaks becomes
larger. The width of the peaks gets larger with larger ατ0
in agreement with the dependence of the optical depth on
frequency (i.e., when in the less optically thick regime core
photons are relevant and the optical depth goes as e−x2
,
whereas in the more optically thick regime wing photons
are more relevant, and there the optical depth scales as
1/x2).
In the right panel of Figure 1 we present numerical re-
sults when recoil is included, along with the analytical
solution (as a guide) that does not include recoil. As
expected, including recoil shifts more photons to smaller
(more red) frequencies. The magnitude of the effect can
be understood as reflecting the thermalization of photons
around frequency ν0 (Wouthuysen 1952; Field 1959). This
process modifies the photon abundance by exp(−x/xT ),
with xT = kBT /h∆νD. Indeed, in the right panel of Fig-
ure 1 the dashed line for τ0 = 106 is obtained by modifying
by exp(−x/xT ) the emerging spectrum obtained from the
simulation when no recoil is included. These results are in
agreement with the results and interpretations by Zheng
& Miralda-Escud´e (2002).
2.2.2. Loeb & Rybicki (1999) test
Loeb & Rybicki (1999) address the RT problem in a
spherically symmetric, uniform, radially expanding neu-
tral hydrogen cloud surrounding a central point source of
Ly-α photons. No thermal motions are included (T = 0
K). They find that the mean intensity J(r, ν) as a function
of distance from the source r and frequency shift ν in the
diffusion (high optical depth) limit is given by
J =
1
4π (cid:18) 9
4πν3(cid:19)3/2
exp(cid:18)−
9r2
4ν3(cid:19)
(17)
with ν = ν/ν⋆, ν = ν0 − νphoton, ν0 the Ly-α resonance
frequency, and ν⋆ the frequency where the optical depth
becomes unity. The scaled radius, r is equal to r/r⋆, with
r⋆ the physical distance where the frequency shift due to
the Hubble-like expansion of the hydrogen cloud equals
the frequency shift that corresponds to unit optical depth
(= ν⋆). A comparison of the results from the code with
the analytic solution is shown in Figure 2. The analytic
solution becomes progressively more accurate the higher
the optical depth (or the smaller the frequency shift in the
way this problem is parameterized, so that we are still
in the core of the line).
it deviates more
and more from the (exact) simulation result at larger r,
since the larger the r the more optically thin the medium
and thus the further away we are from the assumption of
an optically thick medium made by the analytic solution.
Thus, the disagreement at high r is real and not an arti-
fact caused, e.g., by small number of photons that would
be inadequate to sample the low intensities at large r.
In addition,
2.2.3. Simple models of Ly-α emitters: Spherical clouds
of uniform density and temperature
Here we develop some simple models of Ly-α emitters.
Even though there are no analytic solutions for these cases,
one could compare our results with the published results
of Zheng & Miralda-Escud´e (2002). More speficically, in
Fig. 2. -- Mean intensity as a function of radius for certain fre-
quency shifts. Solid lines show the results from the Monte Carlo
code and dotted lines show the analytic solution of Loeb & Rybicki
(1999), appropriate in the diffusion limit. The specific frequency
shifts plotted were chosen based on the fact that the diffusion limit
is the right limit for ν << 1 (for details and definitions see text).
this section, following these authors we model spherical
neutral hydrogen clouds. We consider two different cases
as far as the emission is concerned. In the first case it is as-
sumed that we have a spherical cloud with a Ly-α emitting
point source at its center. In the second case we assume
uniform emissivity, namely a photon is equally likely to
be emitted from any point within the cloud. For each one
of these two cases we make runs assuming the cloud is
static, contracting and expanding. In the latter two cases
the contraction/expansion is assumed to be Hubble-like,
namely the velocity of the neutral hydrogen atoms scales
linearly with the radius measured from the center of the
cloud. This velocity is set equal to 200 km/s at the edge
of the system (and is negative/positive in the case of con-
traction/expansion). For each case we perform two runs,
one with column density equal to 2× 1018cm−2, typical for
Lyman limit systems, and one with column density equal
to 2 × 1020cm−2, typical of Damped Ly-α systems (or line
center optical depths equal to 8.3 × 104 and 8.3 × 106, re-
In all cases the temperature is set equal to
spectively).
2 × 104K. The initial photon frequency is assumed to be
at the line center in the rest frame of the atom.
In all
results shown, the effect of recoil is included. Lastly, 1000
photons were used in all runs.
The results for the optically thin case are shown in Fig-
ure 3 and that for the optically thick configuration are
shown in Figure 4. In both case the agreement with the
results obtained by Zheng & Miralda-Escud´e (2002) is very
good. These spectra can be understood qualitatively us-
ing the way the Neufeld solution behaves depending on
the optical thickness. In the case of an expanding cloud,
Ly-α radiative transfer in cosmological simulations
7
Fig. 3. -- Top panel: frequency distribution of emergent Ly-α pho-
tons in the case of a static (solid histograms ), a contracting (dot-
ted histograms), and an expanding (dashed histograms), isothermal,
spherically symmetric neutral hydrogen cloud with column density
NHI = 2×1018cm−2 and uniform emissivity. Bottom panel: same as
left panel but the Ly-α photons in this case originate from a central
point source.
Fig. 4. -- Top panel: frequency distribution of emergent Ly-α pho-
tons in the case of a static (solid histograms ), a contracting (dot-
ted histograms), and an expanding (dashed histograms), isothermal,
spherically symmetric neutral hydrogen cloud with column density
NHI = 2×1020cm−2 and uniform emissivity. Bottom panel: same as
left panel but the Ly-α photons in this case originate from a central
point source.
the photons will escape on average with a redshift be-
cause they are doing work on the expansion of the cloud
as they are scattered. Photons with negative frequency
shifts (redshifted) can escape, but those with positive fre-
quency shift (blueshifted) will be scattered at some point
and they have to undergo a series of many positive shift
scatterings to escape. Hence, the blue part of the spec-
trum is suppressed. The situation is reversed in the case
of a contracting cloud.
It is important to keep in mind
that the degree of suppression of one of the two peaks
due to bulk motions depends on factors such as the op-
tical depth and the temperature. In the case of uniform
emissivity and expansion/contraction all spectra become
broader because of the different velocities of the emission
In addition, when the cloud is ex-
sites of the photons.
panding (contracting), the blue (red) part of the spectrum
is not suppressed as much as in the central point source
case because, at least, photons initially emitted close to
the edge of the system have some chance of escaping even
if they are blue (red).
In the optically thicker cloud, as
soon as the photon reaches a sufficiently large x it is not
likely that it will be scattered by an atom with the right
velocity to bring the photon into the line center. Rather,
the photon will get another random shift in frequency and
will follow an excursion in frequency while at the same
time it diffuses spatially. This along with the fact that
the optical depth has a power law rather than Gaussian
dependence on x broadens the peaks compared to those of
the optically thin case, exactly as discussed for behavior of
the Neufeld solution. In addition, the emission peaks move
8
TASITSIOMI
further away from the center compared to those from the
optically thin case, since the photons have to be further
away from the center of the line in order to escape when
the medium is optically thick.
2.3. Accelerating the RT
c = e−x2
c/√π, where xc = (νc − ν0)/∆νD and α
The previous tests demonstrated that our basic Monte
Carlo scheme works well for the simple test cases. When
using it in its simplest form in high resolution cosmological
simulations, such as the ART simulations (see §3), it takes
unrealistic running times in order to produce results with
sufficient numbers of photons. This is because in the case
of resonant RT in extremely optically thick media, a sig-
nificant amount of time is spent on the relatively insignifi-
cant core scatterings. If we define the core through the fre-
quency range where the Doppler profile dominates over the
Lorentzian wings, then roughly speaking the core is given
by α/πx2
is as defined previously. For a temperature of 105 K, the
core is roughly xc = 3.5. Also, assuming complete redis-
tribution,5 the probability per scattering for a core photon
to exit the core is I/(I + erf(xc)) with I = 2α/(πxc) and
erf(xc) = 2
dt. That is, roughly, the photon will
have to scatter (I + erf(xc))/I = 1 + erf(xc)/I ≃ 1 + I−1 ≃
I−1 times before exiting the core. Using the above core
definition, one finds that this is equal to √πex2
c /(2xc)
and keeping only the dominant dependence on xc, this
is roughly ex2
c or ∼ 105 scatterings. These scatterings are
insignificant in the sense that they happen in such copious
amounts, without being accompanied by significant spa-
tial diffusion, since the latter occurs mostly through the
wings.
0 e−t2
√π R xc
One way to advance photons in very high optical depths
is to use the technique of the prejudiced first scattering
(Cashwell & Everett 1959). With this technique one bi-
ases the τ values toward larger values than the ones that
would be drawn from equation(1). More specifically, τ is
chosen to be uniformly distributed in [0, τesc], with τesc
the optical depth for escape. Then one weights the pho-
tons by τesce−τ to correct for the fact that τ (i) is limited
to be less than or equal to τesc, and (ii) is assumed to be
uniformly distributed in the < τesc range. Using this tech-
nique however does not improve run time requirements to
the extent we need and clearly more drastic acceleration
methods are needed.
Exiting the core does not in general guarantee that the
photons escape. In fact, the photons may return back to
the core many times before escaping. This is not surpris-
ing since, as we will discuss, the maximum core frequen-
cies that can be used are much smaller than x∗ discussed
previously. Especially for extremely optically thick me-
dia (ατ0 > 103), this in -- and out -- of -- the -- core procedure
is still very expensive to follow. Hence, we accelerate our
5 In other words, assuming that the frequency distribution after scat-
tering is independent of the frequency before scattering and is given
by the line profile (i.e., the source function is independent of fre-
quency). The assumption of complete redistribution was found to
be pretty accurate for core photons (Unno 1952; Jefferies & White
1960). This is intuitively expected since, when in the core, the pho-
ton frequency shift is small or comparable to the thermal velocities
of the atoms. Thus, the latter can have a significant impact on the
frequency of the photon and in effect they redistribute it after each
scattering according to the line profile.
RT scheme by implementing two different methods, de-
pending on the center-of-line optical thickness of the cell
a photon finds itself in (τ0), as well as on the thickness of
the cell for the specific frequency shift of the incident pho-
ton (= τ0φ(xi) with xi the frequency shift of the incident
photon). In fact we parameterize the optical thickness of
a cell not only via τ0, the line-of-center optical depth from
the center of the cell to one of its edges, but rather via the
product of α and τ0, motivated by the Neufeld solution.
This parameterization turns out to be very good for media
less optically thick than those the Neufeld solution applies
to. We discuss these two acceleration methods, as well as
some additional acceleration techniques in the following
subsections.
2.3.1. Extremely optically thick cells: Controlled Monte
Carlo motivated by the Neufeld solution
This acceleration scheme is based on controlled Monte
Carlo simulations of resonance RT in cells (cubes) with
several physical conditions, representative of the extremely
optically thick cells in the simulations. The idea is to ob-
tain trends and best -- fit functional forms for the spectra
emerging from thick cells. These spectra can then be used
when running the code so that instead of following the
scattering of the photons in detail, we can draw the fre-
quency of the photon emerging from a thick cell using the
pre-calculated spectrum appropriate for the physical con-
ditions in this cell. In principle, controlled Monte Carlo
simulations can be used for any range of optical thick-
nesses. We use it only in the extremely optically thick
cells where (ατ0)ef f > 2 × 103 and (τ0φ(xi))ef f ≫ 1 with
xi the frequency shift of the incident photon.6 We do that
because we are motivating this method by the Neufeld so-
lution which is applicable only at the diffusion limit. The
inherent cell structure of the AMR simulation outputs or
the cell structure that can be generated for other cosmolog-
ical codes, along with the resolution imposed isothermality
and uniformity of each cell, are conducive to some kind of
modification of the Neufeld solution. In some sense, with
the advent of cosmological simulations, the contemporary
analogue of the extensively studied classical slab problem
is the completely unexplored problem of resonance RT in
a cube. This motivated a detailed study of the resonance
RT problem in cubes where the reader is referred to for
more details and results (Tasitsiomi 2006). Here we only
summarize briefly some key results relevant to the current
study.
As discussed in §2.2.1, the Neufeld solution was obtained
under some assumptions. To fit the controlled Monte
Carlo spectra with a Neufeld type spectrum we have to in-
vestigate how sensitively the solution depends on these as-
sumptions, as well as whether these assumptions are valid
in cosmological simulations. This is done in the following
paragraphs.
6 We have used the index eff because, as is discussed later in this
section having in mind an implementation of the RT code for AMR
simulations, to decide whether this method is applicable or not we
create a mesh on top of the simulation mesh. In this new mesh, the
photon is always at the center of a cell. Then it is the 'effective'
physical conditions in this new cell that are relevant when decid-
ing if the acceleration method at hand is applicable or not. In the
case of simulations without a cell structure, the index eff becomes
redundant, since there is no initial mesh to begin with.
Ly-α radiative transfer in cosmological simulations
9
Choosing the exiting frequency
The exiting frequency of a photon entering an extremely
optically thick cell is drawn by an emerging frequency dis-
tribution similar to the Neufeld solution (equation 16).
However, the Neufeld solution is derived for a semi-infinite
slab, whereas the simulation cells are finite cubes. Further-
more, the solution assumes isotropic scattering, no recoil,
which anyway is negligible in the simulations, and does
not include velocities such as those associated with pecu-
liar motions or the Hubble expansion. Lastly, it assumes
that the source of the radiation lies within the slab,7 and
is valid for optically thick frequencies (τ0φ(xi) ≫ 1).
Starting from the point on bulk velocities, we use the
Neufeld solution -- applicable for an observer moving with
the bulk flow of the fluid -- by taking into account the
way the specific intensity transforms between two iner-
tial observers moving at a certain speed with respect to
each other (i.e., Iν/ν2 is invariant, where Iν is the number
of photons rather than the energy intensity. In the latter
case, the quantity that would be invariant would be Iν /ν3).
The second point we address has to to do with the slab
versus cube difference between the analytical solution and
the simulations. As discussed in §2.2.1, Neufeld's solution
depends on one parameter, ατ0. Qualitatively, one expects
that the spectrum emerging from a cube rather than a slab
be well described by the same solution but for an effective
ατ0 smaller than the actual ατ0 of the cell. The reason for
this is that when for example observing the emergent flux
from the z-direction in a cube, we lose all photons that
in the case of the slab would wander, scatter many times
along the infinite dimensions and finally find their way out
from the z-plane. In the case of the cube these photons
would not be counted simply because they have exited the
cube from planes other than the z-plane. This would be
equivalent to solving the problem that Neufeld solved but
this time including losses of photons (or, more appropri-
ately, by generalizing the 2-dimensional diffusion equation
derived by Neufeld into a four dimensional one -- instead
of τ, ν now the intensity will be a function of τx, τy, τz and
ν). Numerical experimentation of RT in cubes and slabs of
the same physical conditions, verified that the above guess
is correct. In fact, the cube spectrum is well described by
the Neufeld solution for a slab if 2/3 of the ατ0 of the cube
are used as input parameter to the slab analytic solution.
This is shown in Figure 5 (also see Tasitsiomi 2006).
Furthermore, the Neufeld solution assumes that the source
of radiation lies within the slab (or cube in our case). In
fact, the version of the solution we have been discussing so
far (equation 16) assumes that the source is at the center
of the slab. However, in the case of mesh -- based codes, as
photons cross from one cell to the other, in general the
source is not at the center of the cell. For codes without
an inherent cell structure the obvious solution to this is
to create a cell and have the photon at each instant at
the center of the cell. As is discussed in what follows, this
turns out to be the most efficient solution in the case of
7 More specifically, equation 16 assumes that the source is a plane
source in the middle of the slab. Due to symmetry arguments, a
plane source located at the middle of a slab is equivalent with respect
to the spectrum of the emergent radiation to a central point source.
Neufeld (1990) provides a more general expression for different source
positions.
Fig. 5. -- Comparison of the emergent spectra from a semi-infinite
slab (dotted histograms) and a cube (dashed histograms) of the same
physical conditions (ατ0 = 105). Also shown is the analytical so-
lution derived by Neufeld (1990) (solid line) for the emergent spec-
trum from a semi-infinite slab. Note that the analytical solution is
for ατ0 = 2× 105/3, which is the 'effective' ατ0 one has to use in the
analytic solution obtained for a semi-infinite slab, for the solution to
give the spectrum from a finite cube of the same physical conditions
as the slab.
mesh -- based codes as well.
Neufeld provides a more general expression for various
source positions within the slab, as well as for the trans-
mission and reflection coefficients assuming an external
source. Using either option for mesh -- based codes, trying
to take advantage of the already existing mesh structure,
creates complications: in the case of a non-central but in-
ternal source, the equivalence of a point or infinite plane
source -- necessary for all the above discussion to be valid --
breaks down if the source is not located at the center of the
slab. And using the reflection/transmission probabilities
makes the algorithm more complicated. But most impor-
tantly, there is an intrinsic limitation in the simulations
due to finite resolution: it is not clear how meaningful it
is to be discussing differences in position less than the cell
size (i.e., if one can really tell the edge from the center of
the cube). Instead, at every point the photon is found we
create a new mesh on top of the simulation mesh. The pho-
ton is always found at the center of a cell whose physical
parameters are calculated using the cloud-in-cell weighting
scheme. Each time the size of the cell is set to the simula-
tion cell size the photon is in. Note that it is the physical
parameters of this effective cell that determine the way the
code proceeds (i.e., if the effective cell ατ0 is larger than
2 × 103 and τ0φ(xi) ≫ 1 then the controlled Monte Carlo
results are used. If one of these two conditions (or both)
is not satisfied in the effective cell then the code returns
to the original cell. Depending on the original cell phys-
ical conditions and the photon frequency either the exact
10
TASITSIOMI
Monte Carlo or the method described in §2.3.2 is used).
In the Neufeld solution the condition τ0φ(xi) ≫ 1 allows
him to truncate a series appearing in the solution process
by keeping up to first order terms in 1/(τ0φ(xi)). Thus,
the solution is valid only for optically thick injection fre-
quencies. We find that the higher order corrections are
pretty small. However, for a certain tolerance, one must
decide how thick is thick enough for the Neufeld spectrum
to be applicable. We take that the spectrum from a slab is
satisfactorily predicted by the analytical solution for fre-
quency shifts for which τ0φ(xi) ≥ 10.
As has been shown in §2.2.1 the recoil effect can be eas-
ily accounted for multiplying the Neufeld solution by the
appropriate factor. In any event, the recoil effect for our
conditions is negligible and hence is dropped in the simula-
tion calculations. To see this, the recoil effect corresponds
to a frequency shift that would be caused by a velocity
≃ hν/mpc = 3 m/s. This velocity is negligible compared
to the thermal velocities expected in cosmological simula-
tions, and given the peculiar and Hubble flow velocities,
the small non-coherence in the atom's rest frame intro-
duced by recoil will be totally unobservable. Hence, the
Neufeld approximation is good in that respect as well.
Choosing the exiting direction and point
Referring to µ, the cosine of the angle with which the pho-
ton is exiting a cell, measured with respect to the normal
to the exiting surface, we draw its value from the following
cumulative probability distribution function (cpdf) (Tasit-
siomi 2006)
P (< µ) =
µ2
7
(3 + 4µ) .
(18)
This cpdf is found to be an excellent description of the
directionality of the emergent spectrum and clearly devi-
ates from isotropy. In fact, it verifies the findings of other
studies that in optically thick media photons tend to exit
in directions perpendicular to the exiting surface (see, e.g.,
Chandrasekhar 1960; Phillips & Meszaros 1986; Ahn et al.
2002). In the case of RT in accretion disks this has been
identified as an expected limb darkening (or 'beaming';
Phillips & Meszaros 1986) of the disk (i.e., the disk is very
bright when observed face on and less bright when ob-
served edge on). In cases of very optically thick media, the
emerging radiation directionality approaches the Thomson
scattered radiation emergent from a Thomson-thick elec-
tron medium. This Thomson limit obtained initially by
Chandrasekhar (1960), was confirmed later numerically by
Phillips & Meszaros (1986).
It has been implied by some authors (Ahn et al. 2002)
that the fact that in optically thick media RT occurs mostly
via wing photons with the latter being described by a
dipole phase function (see §2.1.2), and the fact that Thom-
son scattering is also described by a Rayleigh (dipole)
scattering phase function, explains why the resulting µ
probability distributions are similar. However, we find the
same cpdf when the scattering is taken to follow either an
isotropic or a dipole distribution. For such optical thick-
nesses the details of the exact phase function do not mat-
ter, at least not with respect to the exiting angle cpdf. All
the phase functions involved in Ly-α scattering are only
mildly anisotropic and they simply enhance a little bit the
coherence of the scattering at the observer's frame com-
pared to the isotropic scattering case. So the fact that
the exiting angle cpdf in extremely optically thick slabs
(cubes) does not depend crucially on the assumptions on
the phase functions does not come as a surprise. The un-
derlying physics is simply that in extremely thick media
most of the photons escape along the normal to the slab
where the opacity is smaller. The azimuthal angle φ with
which the photon exits a cell is distributed fairly uniformly
in [0, 2π] (for more details see Tasitsiomi 2006).
Referring to the distribution of exit points, one can ar-
gue that trying to specify the exact coordinates of the exit
point of a photon from a simulation cell is, in some sense,
superfluous since there is always the resolution limitation.
Thus, we assume that the exiting points are distributed
uniformly. The deviations of the exiting points from uni-
formity are relatively small (Tasitsiomi 2006). Similarly,
resolution limitations make us focus on total distribution
functions of photon properties -- where total here means
distributions averaged over an entire cube side -- without
regards to a possible dependence of these distribution func-
tions on the photon exit point.
Lastly, we have checked whether the emergent photon
parameters can be drawn independently. We found no
significant correlations among them (e.g., we checked for
correlations between emergent frequency shift and (pre-
ferred) range of exiting directions). Thus, drawing them
independently is correct.
2.3.2. Moderately optically thick cells: Skipping the core
scatterings
This acceleration scheme is used if the cell the photon
is in has 1 ≤ ατ0 ≤ 2 × 103. It is also used in the case
of cosmological simulation codes with a pre-existing mesh
when the cell the photon is in has ατ0 > 2 × 103, but
the effective cell (see §2.3.1) has 1 ≤ ατ0 ≤ 2 × 103, and
thus the previous acceleration scheme (discussed in §2.3.1)
is not applicable. The scheme is based on the idea that
if a photon is within a certain core (to be determined),
we can skip all the core scatterings and go directly to the
scattering with a rapidly moving atom that can bring the
photon out of the core (for some first implementations of
this idea see Avery & House 1968; Ahn et al. 2002). As
soon as this happens, the initial detailed transfer resumes
until either the photon escapes or re-enters the core. The
scheme's validity relies upon the correct choice of the core
value, so that while in the core the photon does not diffuse
significantly in space, whereas significant diffusion occurs
when the photon exits the core.
To achieve the scattering that brings the photon out-
side the core we choose thermal velocities (in units of
p2kT /m) from the distribution (Avery & House 1968;
Ahn et al. 2002)
(19)
p(v) =
1
√π
e−v2
and in the range [vmin, vmax]. The lower limit vmin is
the minimum velocity necessary for the photon to just
make it to the core xc. The upper limit is formally in-
finite, but for any practical realization it can be set to a
c + 10). For a scattering to
bring the photon to just xc from the center, independent
large enough number (e.g., px2
Ly-α radiative transfer in cosmological simulations
11
of the directions of incident and outgoing photon, and un-
der the assumptions of coherence in the rest frame of the
atom, isotropic scattering phase function, and zero radia-
tion damping, it can be shown that vmin = max(x,xc)
(Hummer 1962), with x the initial frequency shift (as usual
in units of the thermal Doppler width). In our case it is
always vmin = xc since the photon is inside the core.
We checked and verified that the assumptions under which
vmin is derived are good for cosmological simulations. This
is not surprising since, e.g., the assumption of an isotropic
phase function is not very crucial. As discussed already,
none of the relevant phase functions is strongly anisotropic.
Those that are anisotropic simply tend to favor slightly
smaller frequency shifts (since they favor post-scattering
directions close to pre-scattering directions) and hence in-
crease a little bit the coherence in the observer's frame
from scattering to scattering. At the limit of many scat-
terings (and while still at the optically thick regime) this
is not a significant effect (for the tiny differences in the fre-
quency redistribution function with isotropic versus dipole
phase function see Figure I of Hummer 1962). Or, the as-
sumption of coherence in the rest frame of the atom is
also expected to be a pretty good assumption for the me-
dia in the simulations from the point of view of the recoil
effect, as we discuss in §2.3.1, and from the point of view
of collisions as we discuss in §3.5.1.
To motivate the core values we can use (i.e., the maxi-
mum frequency shifts for which we can ignore the repeated
scatterings without biasing the results) we must take into
account the different physics of resonant RT in the two
different regimes, 1 ≤ ατ0 ≤ 2× 103 and ατ0 > 2× 103. In
the first regime photons escape on a single longest flight
(Adams 1972) in accordance with the understanding of
resonant RT in moderately thick media developed by Os-
terbrock (1962).
In this thickness regime the important
frequency is the frequency where the optical depth be-
comes unity. Photons within this frequency shift barely
diffuse in space, whereas as soon as they exit this fre-
quency shift they escape while taking their longest spa-
tial step (flight). In the second, extremely optically thick
regime (ατ0 > 2 × 103) as Adams (1972) suggested, pho-
tons escape during a single longest excursion rather than
flight.
In this case the important frequency is the fre-
quency with the following property:
if a photon is given
this frequency and is left to slowly return to the center of
the line (by performing a double random walk, in space
and frequency), the overall rms distance that it will travel
in real space while returning to the line center equals the
size of the medium (i.e., the important frequency shift in
this case is the shift x∗ discussed in §2.2.1). This physics
motivates our cores, i.e., for moderately thick media the
core must be safely optically thick, whereas for extremely
optically thick media the core must be safely smaller than
x∗. Then using numerical experimentation we find the ex-
act maximum possible core values that can be used in each
case.
A comparison of the exact Monte Carlo and the core
acceleration scheme applied to moderately thick media is
shown in Figure 6. Note that these spectra are one cell
runs, and are not the final results of the RT around the Ly-
α emitter (which are discussed in a later section). In the
top panel, we present the exact emergent spectrum from
Fig. 6. -- Top panel: Comparison of the exact Monte Carlo results
('ex.', solid line) and the results obtained using the core acceleration
method ('ap.', dotted line) for the minimum cell ατ = 1 for which
this acceleration method is used in the Ly-α RT code. Also shown is
a larger core frequency, xc = 0.2, which shows the way the emergent
spectrum is biased if one uses a higher core frequency. Bottom panel:
Same as in top panel but for more optically thick cells, ατ0 = 102.
In this case a core frequency xc = 0.8 can be used.
In addition,
we show exact results for a different pair of temperature and optical
depth (dashed line), that however correspond to ατ0 = 102. Clearly,
ατ0 is a good way to parameterize the problem at these moderate
optical thicknesses.
a cube with ατ0 = 1, as well as the spectrum obtained if
a core xc = 0.02 is used. Despite it being a pretty small
core, it improves the speed of the algorithm by orders of
magnitude.8 Also shown is what the bias would be if one
used a higher core frequency (xc = 0.2): photons would
be artificially shifted at higher (absolute) frequency shifts.
8 The exact improvement factor depends on optical thickness, and
is higher for thinner cells. Furthermore, the improvement factor is
different for the same ατ0 but different temperatures and optical
depths. More specifically, it is higher for lower optical depths and
temperatures.
12
TASITSIOMI
To find the maximum core that can be used without this
biasing, we made runs with successively higher cores. We
use as cores: 0.02 for 1 ≤ ατ0 < 10, 0.1 for 10 ≤ ατ0 < 102
and 0.8 for 102 ≤ ατ0 < 2 × 103. One can easily verify
that for a wide temperature range these cores are safely
within the optically thick regime.
The comparison between the exact Monte Carlo and
the accelerated scheme for optically thicker cells (but still
at the moderately thick regime) is shown at the bottom
panel of Figure 6. We have seen via the Neufeld solution
that characterizing a slab -- or a cube in our case -- us-
ing ατ0 is very good in the case of very optically thick
media (ατ0 ≥ 103). In the bottom panel of Figure 6 we
present two different sets of temperature and τ0, which
nevertheless correspond to the same ατ0 (and smaller than
that for which the Neufeld solution is applicable). Clearly,
ατ0 parameterizes nicely enough these emergent spectra
as well. This fact justifies our classification of simulation
cells with respect to their ατ0 value. Note that the fact
that the emergent spectrum for these physical conditions
seems to depend on ατ0 is not trivial, and was checked
only for ranges of temperature and optical depth that are
anticipated to be relevant to cosmological simulation en-
vironments. A simple way to see why this may not be a
general statement comes from the physics of RT in mod-
erately thick media. As discussed in such media photons
escape roughly when they reach the frequency where the
optical depth is unity. If, for example, the frequency shift
x where the optical depth becomes unity is within the
Doppler core (as anticipated) then this frequency shift is
defined through τ0e−x2
= 1 and clearly depends only on
τ0 and not on temperature. This is in contrast to ex-
tremely optically thick media where the frequency shift
relevant for escape through the single longest excursion is
x∗ ∼ (ατ0)1/3 (see §2.2.1), namely it depends on ατ0.
In the case of extremely thick media we find roughly
the following maximum possible cores: 3 for 2 × 103 ≤
ατ0 < 104, 5 for 104 ≤ ατ0 < 105, 7 for 105 ≤ ατ0 < 106,
17 for 106 ≤ ατ0 < 107, 30 for 107 ≤ ατ0 < 108, and
80 for ατ0 ≥ 108. As an example, in Figure 7 we show
the Neufeld prediction for the emergent spectrum from a
slab with ατ0 = 107 and the results of our acceleration
scheme using a core xc = 30. This is a quite large core
frequency, and still the acceleration scheme gives a very
accurate emergent spectrum. The core values we find scale
with x∗ roughly as xc ≃ 0.15x∗.
The ατ0-dependent core frequencies that we motivate
here based on the different physics for different ατ0 regimes
is a quite new approach. Previous studies (e.g., Hansen &
Oh 2005) define the core frequency as the frequency where
the wings start dominating over the Doppler core. Clearly,
to achieve the best efficiency of the acceleration scheme,
which is highly desirable in our applications due to the very
complex environments, we have to use a depth -- dependent
core definition. Other authors who considered variation
of the core frequency with temperature and optical depth
(Ahn et al. 2002) find a bit different values than ours, at
least for the low ατ0 range that they worked with: they
find that a core frequency of about √π can be used for
ατ0 > 103, with slightly higher values permitted for even
larger τ0. However, we find that this value is a bit large for
ατ0 ≃ 103, and that significantly higher core values can be
Fig. 7. -- Comparison of the analytic solution obtained by Neufeld
(1990, 'ex.', solid line) and the results obtained using the approxi-
mate core acceleration method ('ap.', dotted line) for ατ = 107 and
xc = 30. The point of the figure is that at extremely high optical
depths the core values one can use can be pretty high.
used for higher τ0. The reasons for our disagreement with
Ahn et al. (2002) are not clear.
The discussion with respect to the validity of this ac-
celeration scheme has been limited so far to the emergent
spectrum of radiation. One would expect to see that in-
deed the assumption that the photons do not move signif-
icantly in space during the multiple core scatterings that
are skipped is true. And, that all other quantities, such as
exit point and exit angle distributions remain the same, in
addition to the emergent spectrum. The latter has been
tested and found true. Furthermore, note that the an-
gle information is relevant mostly when the photon is at
the optically thin regime, where anyway we use the exact
transfer scheme. With respect to the exit points, or dis-
tances that the photons move while in the core, since these
are not larger than one cell size, limitations due to the fi-
nite simulation resolution render these concerns moot. To
get an idea, following an argument similar to that pre-
sented in §2.2.1 leading to x∗ ≃ (ατ0)1/3, and using the
scaling xc ≃ 0.15x∗ one finds that by ignoring the scat-
terings within the core for extremely optically thick cells
roughly one ignores a spatial diffusion of the photons of
order 10−3 of the size of a simulation cell.
In summary, each time a photon enters a simulation cell,
there are the following three possibilities:
1. If the cell has ατ0 < 1, the exact Monte Carlo RT
is used.
2. If the cell has 1 ≤ ατ0 ≤ 2 × 103 and the photon
frequency shift is x ≤ xc, then we skip the core
scatterings. If the photon frequency is outside the
core we use again the exact Monte Carlo RT.
Ly-α radiative transfer in cosmological simulations
13
3. If the cell has ατ0 > 2 × 103 then if there is no pre-
existing mesh structure of the cosmological simula-
tion then if the frequency of the photon is such that
τ0φ(x) ≫ 1 then the controlled Monte Carlo results
are used. If there is a pre-existing mesh structure,
then if ατ0 > 2 × 103 then the physical conditions
of the effective cell are calculated. If for the effec-
tive cell it is (i) ατ0 > 2 × 103 and the frequency
of the photon is such that (ii) τ0φ(x) ≫ 1, then we
use the controlled Monte Carlo motivated by the
Neufeld solution.
If either (i) or (ii) is not true,
then the first acceleration scheme is tried for the
original rather than the effective cell. And if it is
not applicable, then the exact Monte Carlo scheme
is used.
2.3.3. Calculating images and spectra
To construct images of the Ly-α emitters for various di-
rections of observation the code calculates the contribution
to the image along a certain direction at each scattering
(see, e.g. Yusef-Zadeh et al. 1984; Zheng & Miralda-Escud´e
2002). This contribution is e−τescP (φ, µ) where τesc is the
optical depth for escape from the current scattering po-
sition along the direction of observation to the observer,
µ is the cosine of the angle between the direction of the
incident photon and the direction of observation, φ is the
azimuthal angle, and P (φ, µ) is the normalized probability
distribution for the photon direction (in fact P is indepen-
dent of φ in our case).
This way of calculating images and spectra has the ad-
vantage of giving fairly good statistics for relatively small
numbers of photons. Thus, by lowering the number of
photons needed for the results to converge, it can poten-
tially speed up the calculations. It also converges rapidly
for the fainter parts of the source, hence it is very useful
for sources with high emissivity contrast. One disadvan-
tage is that due to computing resources limitations it lim-
its the calculations to only a small number of pre-chosen
directions of observations. In addition, for complicated ge-
ometries such as those produced in simulations one must
verify that running more photons is more expensive than
calculating τesc used in this method. We find that indeed
this is the case for the ART environments where the RT
code is applied in this study.
2.3.4. Parallelization
To reach high performance we implement the parallel
execution of the code. Our Monte Carlo scheme is partic-
ularly easy to parallelize, since each ray is independent of
others. The parallelization is done using the Message Pass-
ing Interface (MPI) library of routines. As every photon
ray is independent, communication requirements among
the different processes are minimal, and in essence MPI
distributes copies of the code which are run autonomously
in the different nodes used. However, each processor is as-
signed and runs photons from different emission regions.
To get an idea about the performance of the code (using
the above acceleration schemes), 107 photons 9 transfer to
10 physical kpc from the center of the ART Ly-α emitter
9 This number of photons is well above the minimum necessary for
the results to converge as will be discussed in a later section
we apply the code to in about 4 hours on 8 Intel Xeon 3.2
GHz processors on the Tungsten NCSA cluster .
2.4. Final images and spectra of simulated Ly-α emitters
The detailed Ly-α RT is carried out up to a certain
distance from the center of the source and then the Ly-α
GP absorption is added. This distance where the detailed
RT stops is determined through a convergence test. The
existence of such a scale is guaranteed given that the fur-
ther away a photon moves from the center of the object,
the most improbable it becomes for it to scatter back in
the direction of observation. Furthermore, the size of this
convergence radius can also be motivated observationally,
from the extent of Ly-α halos that have been observed.
The surface brightness of each pixel of the constructed
image is
SBp =
Σi,jFi,je−τesc,i,j P (φ, µ)
Ωpix
× e−τGP ,
(20)
where the sum is over the fluxes of all photons (i), and all
their scatterings (j) with scattering positions that project
onto the pixel; Ωpix is the angle subtended by the pixel to
the observer, and the factor e−τGP accounts for the dimin-
ishing of the brightness due to the hydrogen intervening
between the radius where the detailed RT stops and the
observer. To find the flux Fi,j carried by each photon at
each interaction, we first calculate the total luminosity,
Ltot, of the emitter through the sum of the luminosities of
the individual source cells. For N photons (or more ac-
curately wavepackets) used in the Monte Carlo, then each
photon carries a luminosity Fi,j (independent of photon
and scattering numbers i and j, respectively, in our case)
equal to
(21)
Fi,j =
Ltot
N
1
d2
L
where dL is the luminosity distance calculated for the
adopted cosmology. Note that there is no 1/4π factor.
This factor comes from P (φ, µ) -- in equation (20) -- which
is normalized to unity.
The GP absorption optical depth is calculated as de-
scribed in Hui et al. (1997). It is calculated for each pixel
separately, and the number of different lines of sight that
have to be used per pixel is determined by checking con-
vergence of the final result. For high enough image spatial
resolution (similar to the one used in this study) one line
of sight per pixel is enough, since the simulations them-
selves have finite spatial resolution. The characteristics of
the line emerging after the detailed RT (i.e., its width) and
before adding the GP absorption determine how far away
in distance one must go when calculating τGP , since one
needs to go up to the point where the shortest line wave-
length is redshifted at least to the Ly-α resonance because
of Hubble expansion. Often, this physical distance is larger
than the physical size of the cosmological simulation box.
In this case, we take advantage of the periodic boundary
conditions and use replicas of the same box making sure
we do not go through the same structures. This turns out
to be easily done as long as one does not have to use the
box too many times (more than ∼ 5). Furthermore, we
consider two distinct scenarios, one where the effect of the
red damping wing is taken into account and one where the
red damping wing is suppressed as would be the case if for
14
TASITSIOMI
example the Ly-α emitter was in the vicinity of a bright
quasar.
Lastly, spectra are obtained by collapsing the 3-D image
array (2 spatial dimensions+wavelength) along the spatial
dimensions.
3. application to cosmological simulations
3.1. The simulations
Here we present some basic information regarding the
cosmological simulations we use in what follows in order
to apply the Ly-α RT code in a cosmological setting.
The RT is carried out using outputs of the ART code for
the concordance flat ΛCDM model: Ω0 = 1 − ΩΛ = 0.3,
h = 0.7, where Ω0 and ΩΛ are the present-day matter and
vacuum densities, and h is the dimensionless Hubble con-
stant defined as H0 ≡ 100h km s−1 Mpc−1. For the power
spectrum normalization the value σ8 = 0.9 is used. This
model is consistent with recent observational constraints
(e.g., Spergel et al. 2003). The initial conditions of these
simulations are the same as those in Kravtsov (2003) and
Kravtsov & Gnedin (2005), leading to the formation of a
Milky Way sized galaxy at z = 0. However, these simu-
lations are different in that, in addition to dark matter,
gas dynamics, star formation and feedback, cooling, etc.,
they also include non-equilibrium ionization and thermal
balance of H, He, H2 and primordial chemistry, full RT
of ionizing radiation and optically thin line RT of Lyman-
Werner radiation. The continuum RT is modeled accord-
ing to the Optically Thin Variable Eddington Tensor ap-
proximation described in Gnedin & Abel (2001), whereas
cooling uses the abundances of species from the reaction
network, as well as corrections for cooling enhancement
due to metals.
The code reaches high force resolution by refining all
high-density regions with an automated refinement algo-
rithm. The criterion for refinement is the mass of dark
matter particles and gas per cell. Overall there are 9 re-
finement levels. The physical size of a cell of refinement
level l is 26.161 × 29−l pcs at z ≃ 8 (the redshift we fo-
cus on in this study). The dark matter particle mass at
the highest resolution region is 9.18× 105h−1M⊙, and the
box size for which results are presented in this paper is
6h−1Mpc.
For each simulation cell we have available information
such as the temperature, the peculiar velocity, the neutral
hydrogen density, the ionized hydrogen density, the metal-
licity, etc. With this information and using the mesh of the
ART code itself we follow how Ly-α photons are being ini-
tially emitted and subsequently getting scattered. As an
example of an application of the Ly-α RT code developed
for the ART code we focus on the most massive emitter
at z ≃ 8. This emitter is found within a highly ionized,
butterfly -- shaped bubble. Outside this bubble the Uni-
verse is highly neutral, whereas some dense neutral cores
associated with the forming galaxy exist within the bub-
ble. Results for more emitters, different redshifts, multiple
directions of observation, larger simulation boxes, etc., will
be presented in future papers.
3.2. Intrinsic Ly-α emission
There are a number of different mechanisms that can
produce Ly-α emission from high-redshift objects. Here
we classify them into recombination and collisional emis-
sion mechanisms. By recombination emission mechanisms
we refer to Ly-α photons that are the final result of the cas-
cading of recombination photons produced in ionized gas.
The gas may be ionized by the UV radiation of hot, young,
massive stars, from an AGN hosted by the galaxy, or by
the intergalactic UV background. By collisional emission
mechanisms we refer to photons that are produced by the
radiative decay of excited bound (neutral) hydrogen states,
with collisions being the mechanism by which these excited
states are being populated. This mechanism takes place
when gas within a dark matter halo is cooling and col-
lapsing to form a galaxy and radiates some of the gravita-
tional collapse energy by collisionally excited Ly-α emis-
sion, when gas is shock heated by galactic winds or by
jets in radio galaxies, and in supernova remnant cooling
shells. We underscore the fact that the states are bound
states, because in principle collisions can also cause ion-
ization in which case we would have production of Ly-α
photons under a recombination mechanism, according to
our definition conventions. With the exception of AGN
and jets, which are not included in ART simulations, as
well as the fluorescence emission due to the intergalactic
UV background which would be relevant at lower redshifts
than we focus on in this study, we will try to briefly assess
the importance of these separate Ly-α emission sources.
This is interesting in particular because, in addition to
the different dependence on the physical parameters (i.e.,
different temperature dependence and dependence on ion-
ized versus neutral hydrogen), these mechanisms may also
have a different spatial distribution. For example, shock
heated gas from gravitational collapse may be a spatially
more extended Ly-α source than the gas photoionized by
UV radiation of young stars at the relatively compact star
forming regions. The dominant source of Ly-α emission
may be what distinguishes most Ly-α emitters from the
more extended sources referred to in literature as Ly-α
blobs (Steidel et al. 2000; Haiman et al. 2000; Fardal et al.
2001; Bower et al. 2004).
Before discussing the different Ly-α emission mecha-
nisms, we should first mention that, due to practical lim-
itations (i.e., we can only use a relatively limited number
of photons), we use as source cells only the cells that con-
tribute significantly to the total luminosity of the object.
Hence, we set a threshold on the cell luminosity and use
as source cells only the cells whose luminosity exceeds this
threshold. Then by performing a convergence test, namely
by doing runs assuming different luminosity thresholds up
to the point where including lower luminosity source cells
does not change the results (within some pre-specified tol-
erance), we determine the minimum luminosity a simula-
tion cell must emit to be one of the cells where photons will
originate from. It is meaningful to consider a similar con-
vergence check with respect to the Ly-α RT results, and
this will be discussed in a later section. The convergence
test reveals that the luminosity of the object is dominated
by a few very luminous cells. To get an idea, the luminosi-
ties of cells within the virial extent roughly range from
1041 to several times 1054 photons/s. The total luminos-
ity of the object is the sum of the luminosities of the cells
considered. Even though most of the volume, say, within
the virial radius is in low to moderate luminosity cells,
Ly-α radiative transfer in cosmological simulations
15
the sum of the luminosities of these cells is not significant
enough compared to the less numerous high luminosity
cells. For the object at hand the convergence test suggests
that one can use as source cells only cells with luminosities
above ≃ 5 × 1050 photons s−1. This value determines the
relative importance of the different Ly-α emission mecha-
nisms discussed in what follows. With the aforementioned
luminosity threshold, the total luminosity of the emitter at
hand is roughly equal to 4.8× 1043 ergs/s. We sample the
emission region (i.e., the cells with luminosity above the
luminosity threshold discussed) by emitting equal weight
wave packets, but in numbers that reflect the relative lu-
minosities of the cells.
Note that this discussion on the various mechanisms,
emission rates, etc., should somehow be affected by the
limited simulation resolution, a factor that will be studied
in detail in the future. Furthermore, the approach adopted
in this section is an 'order-of-magnitude' one. We defer a
more thorough and statistical analysis of the Ly-α emis-
sion sources in high redshift galaxies to a future study,
where all factors will be taken into account. For exam-
ple, the discussion about the importance of the various
emission mechanisms must be extended to the after RT
results and after including dust. This is because it could,
for example, be the case that recombination Ly-α photons,
despite being more numerous as discussed below, may be
more likely to be absorbed than collisional Ly-α photons,
if one assumes that there is more dust in star forming re-
gions -- where recombination photons are generated -- than
in regions where collisional Ly-α photons originate from.
3.2.1. Ly-α photons from recombinations
The recombination rate of a cell is
r = nenpαBV
(22)
with ne, np the number density of electrons and protons,
respectively, and V the volume. In principle species other
than hydrogen may contribute to ne. Thus, ne in general
is not equal to np. In what follows, we take into account
electrons contributed by the ionization of He. Other BBN
predicted species such as Li, Be and B (with, anyway, tiny
abundances), and elements produced through stellar pro-
cessing such as C, N and O are not taken into account.
Recombination photons are converted with certain effi-
ciency into Ly-α photons. In particular, for a broad range
of temperatures centered on T = 104 K, roughly 38% of
recombinations go directly to the ground state. A fraction
∼ 1/3 (32%) of the recombinations that do not go to the
ground state go to 2S rather than 2P and then go to the
ground state via two continuum photon decay (cf. Table
9.1 of Spitzer 1978). Hence, only a fraction ∼ 40% of the
recombinations yield a Ly-α photon. The temperatures
of simulation cells within the virial extent of the emitter
are in the 102.4 − 106.3 K range, with most cells in the
104 − 106 K range. Due to the weak temperature depen-
dence of the various recombination coefficients the above
conversion efficiencies are roughly applicable throughout
this temperature range. Furthermore, if the gas is optically
thick, then photons that originate from recombinations to
the ground state will be immediately absorbed by another
neutral hydrogen atom and eventually they, as well, will
produce Ly-α photons. Assuming for now that this is the
case (as will be discussed later in this section), as well
as that the medium is thick in Lyman-series photons, so
that all higher Lyman-series photons are re-captured and
eventually yield Ly-α photons, we adopt case B recombi-
nation. For the recombination coefficient we use the fit
obtained by Hui & Gnedin (1997), accurate to 0.7% for
temperatures from 1 to 109K
αB = 2.753 × 10−14cm3s−1
λ1.5
(23)
h1 +(cid:0) λ
2.74(cid:1)0.407i2.242
with λ = 2Ti/T , and Ti = 157807 K the hydrogen ioniza-
tion threshold temperature. In agreement with the above
argument, the effective recombination coefficient at level
2P is approximately 2/3 of the case B recombination co-
efficient and that is what we use to convert recombination
rates into Ly-α photon emission rates. Thus we assume
that the conversion efficiency from recombination to Ly-α
photons is exactly the same for all simulation cells. This
is a good assumption since the conversion efficiency has a
very weak temperature dependence.
The exact conversion efficiency for each source cell also
depends on the rate at which collisions redistribute atoms
between the 2S and 2P state. Collisions with both elec-
trons and protons are relevant. To get an idea for the cross
sections involved, for a temperature of 104K and thermal
protons σ2S→2P ≃ 3 × 10−10cm2 (Osterbrock 1989). For
thermal protons and electrons the thermally averaged col-
lisional cross sections for the processes
(24)
and
H(2P ) + p → H(2S) + p
H(2P ) + e → H(2S) + e
(25)
are qp = 4.74 × 10−4cm3/s and qe = 5.70 × 10−5cm3/s,
respectively, for a temperature of 104K (cf. table 4.10 of
Osterbrock 1989). The 2P to 2S transition is relatively
important when the proton number densities are small (<
104cm−3), and in this case there is some probability that
the Ly-α photon gets destroyed through a two quantum
decay. For higher densities the opposite conversion (2S
to 2P ) becomes important, canceling out the destruction
effect (Osterbrock 1989). At the lower density regime,
which is applicable in the simulations since there np <
104 cm−3 everywhere (within the virial extent the proton
number density range is 10−4−102.5 cm−3, with most cells
in the range 10−3 − 1 cm−3), we can check how important
this process really is by comparing the radiative decay time
and the typical time between collisions,
qp(T )np + qe(T )ne
p =
≃ 8.5 × 10−13npT −0.17
A21
4
(26)
where the number densities of protons and electrons were
assumed to be roughly the same and in cm−3, A21 =
6.25 × 108s−1 is the spontaneous radiative decay for the
Ly-α transition, and temperature is measured in 104K
units. The temperature dependence of the collision rates is
taken from Neufeld (1990). For the temperature and pro-
ton/electron density ranges relevant to the source cell con-
ditions, the probability for a collisional 2P to 2S transition
is negligible, at least for the initial emissivity. We discuss
their effect during scattering of the photons in §3.5.1.
One assumption that we make is that the cascading
of the Lyman series photons, as well as the re-emission
16
TASITSIOMI
Fig. 8. -- Left panel: Cumulative probability distribution of center-of-line optical depths of ART simulation cells within the virial extent.
The three different lines correspond to the cell optical depth distribution in Ly-α (solid), Ly-δ (dotted) and Ly-limit (dashed) photons. Right
panel: Ly-α center-of-line optical depth of the simulation cells within the virial extent of the emitter plotted against the cell recombination
rate. Since only cells with the highest recombination rates (≥ 1051 s−1 or, equivalently, luminosities roughly ≥ 5 × 1050 photons s−1) need to
be used as source cells, and almost all of these cells have τ ≥ 103, roughly speaking our 'on-the-spot' approximation is satisfactory (see text
for details).
and re-absorption of photons from recombination to the
ground state, is done 'on-the-spot', namely, locally.
In
our case "locally" means within the same simulation cell.
This assumption is essential if one wants the Ly-α emis-
sivity of a cell to depend on its own recombination rate
only. If not, one faces the complicated situation where the
Ly-α emissivity of one cell depends on the recombination
rates and photon cascade processes that are happening in
other cells as well. The validity of our assumption de-
pends on the optical depth of Lyman series and ionizing
photons when traversing a typical cell in the simulation
(and should also be affected somewhat by resolution). In
the left panel of Figure 8 we show the optical depth prob-
ability distribution function for Ly-α, Ly-δ and Ly-limit
radiation. The distribution function has as independent
variable the optical depth of simulation cells within 10
physical kpc (≃ virial extent) from the center of the emit-
ter. These distributions are very similar, differing only
by the values of τ because of different oscillator strengths
and characteristic frequencies. Clearly, in all cases more
than half potential source cells are not optically thick, and
this is expected to get worse for ionizing radiation beyond
the Lyman limit. However, as shown in the right panel
of Figure 8 the optical depth of a cell correlates with its
recombination rate. In this figure the optical depth plot-
ted is that for Ly-α photons, but it is easy to see how
this scales approximately with optical thickness for other
Lyman-series photons. Since only cells with recombination
rates higher than 1051 s−1 (or equivalently with luminosi-
ties higher than roughly 5 × 1050 photons s−1) are used
as source cells, our 'on-the-spot' assumption seems pretty
satisfactory, if not always accurate. It becomes less and
less accurate the higher we go in the Lyman series, and of
course beyond the Lyman limit but for the time being we
content ourselves with this approximation, given the com-
plexities introduced when this assumption is not adopted.
We will investigate this point further in the future.
Lastly, to get an idea about the physical conditions of
the highest recombination rate (luminosity) source cells,
they consist of two classes with respect to temperature
and neutral hydrogen fraction: one class contains cold gas
elements (T ∼ 103K), with a neutral hydrogen fraction
> 0.9 (and high gas number density). The second class of
very luminous cells consist of warmer gas elements (T ∼
104 K and a bit higher). In the context of Ly-α cooling
radiation, discussed in the next section, the first class of
cells are unable to cool via atomic hydrogen cooling since
they are cold, whereas the second class of most luminous
cells could cool via atomic hydrogen cooling temperature-
wise, but that is not happening because these cells are
highly ionized.
3.2.2. Ly-α photons from collisional excitations
A collisional emission mechanism whose importance for
the simulated objects can be assessed relatively easily is
that of atomic hydrogen cooling. Using the expression by
Hui & Gnedin (1997) for the hydrogen cooling rate (used
in the ART simulations analyzed here), and assuming for
the moment that this energy is all emitted in the form
of Ly-α photons, we obtain for the luminosity (number of
Ly-α photons/s) emitted by a cell
Lcool = 4.6 × 10−8 e−1.18355/T5
1 + T 0.5
5
nenHI V
(27)
Ly-α radiative transfer in cosmological simulations
17
cooling shells, respectively. A thorough investigation of
the relative importance of SNR Ly-α emission with respect
to that from young stars photoionization has been carried
out by Charlot & Fall (1991, 1993). The general conclu-
sion reached is that for a broad range of physical conditions
and assumptions, the SNR contribution is at best a factor
of 2.5 less than that from stellar ionizing radiation. These
results make the effort to include the (anyway not resolved
in ART simulations) SNR contributions superfluous.
3.3. The Ly-α emitter before RT
To get an idea of the size of the emitting region, the
prevailing physical conditions, and for comparison with
results obtained later after including RT, in this subsec-
tion we briefly present the emission spectrum and image of
the emitter as they would appear to an observer at z = 0
if the Ly-α photons escaped without any scattering. An
image and a spectrum of the emitter along a certain di-
rection of observation is shown in the left and right panel,
respectively, of Figure 10.
The image is a surface brightness map (in units of ergs
s−1 cm−2 arcsec−2) of a roughly 1.4 × 1.4 arcsecs2 field
which corresponds to approximately one third of the virial
extent of the dark matter halo the emitter lives in (with
the virial extent ≃ 20 physical kpc in diameter). There are
two distinct emission regions, each one corresponding to
the two progenitors that merged and formed this object.
The color scale for the surface brightness is logarithmic.
Clearly, the emission region is very small (the largest of the
two structures is at most ≃ 2 − 2.5 physical kpc in diam-
eter, if one includes the faintest pixels), compared for ex-
ample to the virial extent of the dark halo. The resolution
of this image is 0.01 arcsecs (≃ 0.05 physical kpc), at least
10 times higher than the best resolution currently avail-
able. As discussed before, for these results only cells with
recombination rates higher or equal to 1051 s−1 are used.
The initial frequency is chosen according to a Voigt profile
that is sampled for each cell out to 10 Doppler widths and
shifted around the bulk (peculiar + Hubble) velocity com-
ponent along the direction of observation. The number of
photons used (3 × 105) has been determined after a con-
vergence study. Note that when we study the convergence
with respect to the number of photons we take into account
that this must be done in parallel with how far away in
the wings we go when sampling the emission Voigt profile
of each cell, since the higher the number of photons used
the better one can sample frequencies further away from
resonance. The convergence procedure gave the aforemen-
tioned number of photons and initial emission frequency
range (i.e., 10 thermal Doppler widths).
NSN
In the right panel of Figure 10, the frequency resolu-
tion is λ/∆λ ∼ 50000. The line shape has converged,
namely the peaks shown correspond to real velocity sub-
structure. For example, the most pronounced peak at
λ = 10952A corresponds to the component of the pecu-
liar velocity along the direction of observation of the most
luminous pixel of the image shown at the left panel (with
coordinates on the image (0.24,-0.42) arcsecs, roughly).
The dominant contribution to this pixel comes from the
highest recombination cell of the emitter with a recombi-
nation rate equal to ≃ 1.3 × 1055 s−1 and a peculiar ve-
locity component along the direction of observation equal
Fig.
9. -- Maximum cooling Ly-α luminosity, Lcool, plotted
against recombination Ly-α luminosity, Lrec, for all ART simula-
tion cells within the virial extent of a Ly-α emitter at z ≃ 8. The
cooling luminosity is the maximum possible Ly-α luminosity from
cooling because it is derived assuming that all cooling radiation is
emitted in Ly-α photons. The solid line shows the case where the
two luminosities are equal. Since, as discussed in the text, only cells
with luminosities roughly above 5 × 1050 photons s−1 contribute
significantly to the luminosity of the emitter, this figure shows that
recombination is dominant over cooling Ly-α radiation.
with T5 the temperature in units of 105 K. This is com-
pared with the recombination luminosity Lrec (≃ 0.68r)
in Figure 9. Taking into account the results of the conver-
gence test performed to specify what is the minimum cell
luminosity that needs to be taken into account (∼ 5× 1050
s−1), we see that the cells which are relevant are cells where
recombination processes dominate, as can be seen in Fig-
ure 9. Namely, similar to previous studies (e.g., Fardal
et al. 2001) we find that the cooling radiation Ly-α con-
tribution is subdominant compared to the recombination
contribution, hence in what follows we focus only on the
latter.
3.2.3. Supernovae Remnants (SNR)
A Ly-α source that yields Ly-α photons both from re-
combinations and collisional excitations is supernova rem-
nants (SNR). Shull & Silk (1979) have computed the time-
averaged Ly-α luminosity of a population of Type II SNR
using a radiative-shock code. They find that the Ly-α lu-
minosity of a galaxy due to SNR is LSN R = 3×1043n−0.5
0
ergs/s, with nH the ambient density in cm−3, E0 the typi-
cal supernova energy in units of 1051, and NSN the number
of supernova per year. Strictly speaking, this quantity also
depends on the assumptions on the IMF, and the lower
and upper stellar masses of the mass range over which the
IMF is to be integrated. This expression includes both
contributions, from recombination and collisional emission
mechanisms: from UV and X-ray ionization (coming from
the hot SNR interior) of the surrounding medium and from
H E0.75
18
TASITSIOMI
to 0.27 × 10−3 the speed of light. As mentioned, for each
emission cell the Voigt profile was used and sampled up
to 10 thermal Doppler widths. The total width of the line
however is dominated by the bulk velocity structure of the
emitter. The full width of the line at the minimum flux
level shown in the figure (10−22 ergs s−1 cm−2 A−1) is
roughly 15A (with the width if bulk velocities are set to
zero being less than half this). This width corresponds
to projected velocities along the direction of observation
roughly in the [−200, 200] km/s range (this is just approx-
imate, note however that the line is not symmetric around
the rest frame resonance). This velocity range is what
is expected given the peculiar velocities of the emitting
cells. Also shown is the spectrum of the smallest of the
two substructures (dotted line) of the image shown in the
left panel. One can easily infer what the spectrum of the
large Ly-α substructure looks like.
The results discussed in this section may be specific to
the emitter at hand, but the considerations themselves are
pretty general. The same kind of procedure must be re-
peated for each individual emitter identified in the simu-
lations.
3.4. The Ly-α emitter after RT
It is interesting to first treat the emitter as a finite con-
figuration. In this case, as soon as the photons exit this
configuration (whose size is taken to be roughly equal to
the virial extent of the object, namely 10 physical kpc)
they travel towards the observer. In other words at first
we ignore the effect of the GP absorption. This context is
pretty similar to that of §2.2 and §2.2.3. We focus on the
emergent spectrum shown with the solid line in the left
panel of Figure 11 .
The spectrum converges if 3 × 105 photons are used,
namely if the same number of photons are used as the
number of photons needed for the initial emission results
(discussed in §3.3) to converge. Of course, the higher the
number of photons the better one samples low intensity
wavelengths. We find that the number of photons used
affects wavelength ranges with flux less than about 10−22
ergs s−1 cm−2 A−1. The spectrum shown in Figure 11 is
produced using 107 photons. The spectral resolution used
in the figure is λ/∆λ ≃ 5000, whereas the spectrum is
identical if ten times better resolution is used. We have
performed a large set of convergence tests among which
the most interesting are for different (smaller) cores for the
acceleration scheme discussed in §2.3.2, and /or a larger
minimum τ0φ(xi) for which the acceleration scheme dis-
cussed in §2.3.1 is used. Our results are pretty robust,
as should following the discussion in §2.3.1 and 2.3.2 with
respect to the one cell convergence results.
Even though meant for a slab, it is interesting to check
if some predictions of the Neufeld solution, such as the fre-
quency where the spectrum has a maximum (≃ 0.9(ατ0)1/3),
are roughly in agreement with the spectrum of the simu-
lated emitter. Of course, the Ly-α emitter environment
is neither isothermal nor homogeneous, and it is not ob-
vious how to define an 'effective' temperature and opti-
cal depth for these purposes. Thus, focusing on order of
magnitude checks, setting the expression for the frequency
where the peak emission occurs in a slab equal to the fre-
quency where the spectrum of the emitter peaks (say in
red wavelengths) one finds that the 'effective' optical depth
and 'effective' temperature of the equivalent slab (i.e., the
slab that would give a spectrum with peak at the frequen-
cies where the emitter spectrum peaks) roughly satisfy the
relation τ0T ≃ 1.4 × 1010 with T measured in eV. The ef-
fective optical depth will be at least equal to the most
optically thick cell the photon found itself in. Since the
emission originates from the most optically thick cells (see
Figure 8), τ0 will be at least 103.
If we assume for ex-
ample a temperature T = 105K, the above relation yields
τ0 ≃ 109 which is roughly the optical depth from the center
of the object to its virial radius along the direction of ob-
servation. Thus, the maximum of the spectrum is roughly
where it is expected to be if one assumes the scaling from
the Neufeld solution (≃ 2550 km/s).
The emerging spectrum looks pretty similar to the spec-
trum that would emerge from a static configuration, namely
it has two quite similar peaks, one to the red and one to
the blue of the Ly-α resonance. Note however that the
peaks are not really symmetric, since the flux decreases
more rapidly near the resonance. The width of the blue
peak at a flux level of 10−22 ergs s−1 cm−2 A−1 is roughly
180 A or ≃ 5000 km/s. We obtain quite a similar spectrum
if we set the bulk velocity field to zero in the code, that is
kinematics do not seem to play a crucial role in this case.
In the case of the specific Ly-α emitter and for the specific
direction of observation, analyzing the bulk velocity field
(i.e., the peculiar velocity field since the Hubble expan-
sion is negligible at the distances we are working) we find
that there is some net infalling motion, but with signifi-
cant transverse velocity components as well. Hence, the
obtained static -- like spectrum does not come as a surprise.
Furthermore, the peak asymmetry due to the existence of
bulk fields depends on the relative magnitudes of the bulk
and thermal velocities (e.g., if the bulk velocity is close to
the thermal we do not expect a significant asymmetry since
one scattering can give, e.g., a red photon moving in a con-
tracting medium a large enough shift to erase the effect of
the contraction) which varies from cell to cell, and it also
depends on the optical thickness. Since thermal velocities
are typically small compared to bulk velocities in simula-
tions, the optical thickness is a more crucial factor. For
such extremely optically thick media where the spectrum
is expected to have in the context of the Neufeld solution
a typical frequency of ≃ 2550 km/s, bulk velocities of at
most some hundreds km/s will not really favor blue versus
red photons (even if the bulk motion was purely inwards)
that much, since both red and blue photons see a very
optically thick medium.
It would be interesting to have a sense of what is the
number of scatterings each photon undergoes before exit-
ing. With the acceleration methods that we have to use
though it is difficult to keep track of this quantity. A
simple way to obtain an order-of-magnitude idea of the
number of scatterings in such or, at least, similar configu-
rations can be obtained by one of the examples discussed
in §2.2.3. For the most optically thick case (τ0 ≃ 8.3×106)
and a point source emitting photons that propagate in a
stationary medium the number of scatterings in one run of
≃ 2000 photons varies from 2.5×103 up to 4×107, with an
average of 8.3× 106, and a median of 6.6× 106. Two thirds
of the photons are in the [4.6 × 106, 2.1 × 107] scatterings
Ly-α radiative transfer in cosmological simulations
19
Fig. 10. -- Left panel: Image of Ly-α direct emission (i.e., assuming the Ly-α photons escape directly to the observer after they are
produced). The approximately 1.4 × 1.4 arcsecs2 (≃ 6.5 × 6.5 physical kpc) field corresponds to roughly one third of the virial extent of the
dark matter halo where the emitter lives. The surface brightness (SB) is bolometric and in units of ergs s−1cm−2arcsecs−2. The SB color
scale is logarithmic. The object had undergone a recent merger, that is why there are two distinct luminous blobs that dominate the emission.
Right panel: Initial Ly-α injection spectrum. Shown are the total spectrum (solid line), namely the spectrum for the image shown in the left
panel, and the spectrum of the smallest of the two blobs in the image (dotted line). Note that the wavelength is in 104A (i.e., µm). The
dashed line shows the Ly-α resonance for z ≃ 8. See text for discussion of the structure of the line.
range. More generally, we find that similar to the Neufeld
problem, the average number of scatterings in this spheri-
cal configuration scales linearly with optical depth at such
thick media (see discussion in §2.2.1), with the proportion-
ality constant of order unity. From this linear scaling of
the average number of scatterings with optical depth, one
can obtain a rough idea of the average number of scat-
terings of photons in the simulation environments (for the
cell optical depth range in the simulations see, e.g., the
left panel of Figure 8). These numbers also make clear
why it is absolutely not feasible to perform Ly-α RT in
the much thicker and more complicated simulation envi-
ronments without some acceleration schemes.
Photons at very optically thick regions have to shift off
resonance significantly to escape, and hence are the ones
responsible for the significant line width of the spectrum
(along with the 1/x2 behavior of the wing optical depth,
as discussed previously). It is meaningful to ask whether
one should really care about these photons, or instead ig-
nore them because may be they are trapped indefinitely
(for any practical purpose) in the dense cells and do not
participate in the radiation propagation. To answer this
question we estimate the photon diffusion time and com-
pare it to the sound crossing and dynamical time scales
(other time scales, such as the Hubble time scale for ex-
ample which is ∼ 1 Gyr at z = 8 are clearly large enough
to be non-relevant). Same as with the number of scat-
terings, to find the exact diffusion times one should fol-
low the detailed RT. Given our acceleration methods this
is not done. Instead we use some useful scalings. Since
the average number of scatterings in very optically thick
media is roughly equal to τ0, then the diffusion time is
roughly td ≃ Nsclmf p/c with lmf p the mean free path be-
tween scatterings defined through hτi = R ∞
0 τ e−τ dτ = 1.
In other words, since τ0 = nσ(x = 0)L, τ = nσ(x)l, then
the mean free path in units of the total (half) width of the
slab is σ(x = 0)/σ(x)1/τ0, with σ(x = 0) the cross section
at the line center and σ(x) the cross section calculated at
an effective x so that the above definition for the mean
free path is valid. Substituting in the expression for td we
obtain td ∼ σ(x = 0)/σ(x)L/c. For a slab with τ0 = 106
we obtain a mean number of scatterings equal to 9.5× 105
and a median equal to 7.2 × 105, whereas 67% of the pho-
tons have between 5.1× 105 and 2.3× 106 scatterings. For
the mean free path we find a mean equal to 2.4 × 10−5,
a median equal to 1.9 × 10−6 and 67% of the scatterings
correspond to mean free paths between 1.4 × 10−6 and
7.7 × 10−6, all in units of the (half) width of the slab
L. For the total distance traveled by the photons before
escaping, we find an average distance of 40.2, a median
of 32.3 -- implying a σ(x = 0)/σ(x) ratio of order 10 --
whereas 67% of photons exit after traveling a distance be-
tween 16.7 and 96.3, with these numbers as before in units
of the width of the slab. Based on spatial random walk
arguments one would have Nsc ∼ τ 2
0 , hence the distance
before escape would be ∼ τ0 or 106 for the specific exam-
ple we use here. However, as discussed in §2.2.1 Nsc scales
linearly with τ0 and this makes a big difference. We find
that the sound crossing time is significantly higher than
the dynamical time for most simulation cells, hence the
latter is the relevant time against which the diffusion time
must be compared. We find that the dynamical time scale
20
TASITSIOMI
Fig. 11. -- Left panel: Emerging Ly-α emission spectrum before adding the Ly-α Gunn-Peterson absorption (GPA) (no GPA, solid line),
with the GPA but without the red damping wing (GP, no DW, short-dashed line), and with GPA and the damping wing (GPA+DW, dotted
line). Note that the wavelength is in 104A (i.e., µm). The long-dashed line shows the Ly-α resonance for z ≃ 8. Right panel: Image of the
Ly-α emitter after RT, GPA and DW. The 1.4 × 1.4 arcsecs2 (≃ 6.5 × 6.5 physical kpc) field corresponds to roughly one third of the virial
extent of the dark matter halo where the emitter lives. The surface brightness (SB) is bolometric and in units of ergs s−1cm−2arcsecs−2.
The SB color scale is logarithmic.
is at least three orders of magnitude or more larger than
L/c which is within a factor of order 102 -- for the various
physical conditions in the simulation cells -- representa-
tive of td. Note that this comparison also justifies the use
of 'static' simulation outputs where the RT is performed,
even though we plan on investigating the possibility of in-
corporating the RT scheme into the dynamical evolution
in the simulations. Furthermore, the effect of the simula-
tion resolution on these conclusions will be investigated in
a future study.
The scattering process diffuses the initial number of
emitted photons on a larger area and hence lowers the
number surface brightness (i.e., number per s cm2 arcsec2
rather than energy per s cm2 arcsec2). In general the sur-
face brightness itself can go either up or down, depend-
ing for example on the velocity structure of the medium.
To quantify this effect on a photon-by-photon basis we
choose to calculate the distance on the plane of the image
between the initial emission point and the point (pixel)
where the photon makes its maximum contribution to the
image (see §2.4 on how spectra and images are obtained).
We find that these distances vary from roughly 10−3 to
10 physical kpc, with a median of 0.27 kpc and a mean of
0.31 kpc. Given that the largest of the two emission re-
gions has a diameter of ∼ 2− 2.5 physical kpc (see, Figure
10), this means that the 'size' of the luminous part of the
object increases on average by more that 10% due to scat-
tering. If, instead, we focus on the region where a certain
fraction of photons originates from we obtain quite similar
results. For example, ignoring the effects of RT, 90% of
the emitted photons that would reach the observer would
originate within a radius of roughly 2.5 physical kpc. The
same percentage of photons after taking into account RT
would come from a radius of roughly 2.9 physical kpc.10
So far we have been ignoring the GP absorption. When
adding this absorption we consider two distinct cases. In
the first case we include the red damping wing of the GP
absorption, and in the second case we set it equal to zero.
The latter best-case scenario is what would happen if for
example the emitter was inside the HII region of a very
bright quasar. The spectrum obtained in the first case is
shown with the dotted line in the left panel of Figure 11,
whereas the spectrum in the second case is shown with
the short dashed line. An image of the emitter as would
appear on earth with the GP absorption and the damping
wing included is shown in the right panel of Figure 11.
Not surprisingly, when the damping wing is not taken
into account the spectrum is identical with that before the
GP absorption with the difference that all flux blueward of
the Ly-α resonance is missing. When including the damp-
ing wing the maximum flux is suppressed by roughly a
factor of 61.7 with respect to the maximum flux without
it. This line is still quite wide, with a width of approxi-
mately 1370 km/s at a flux level of 10−21 ergs s−1 cm−2
A−1, and a FWHM roughly 620 km/s.
Lastly, these results have converged with respect to both
the number of photons and the radius where the detailed
RT stops (and beyond which the GP absorption is added).
More specifically, we find that the number of photons re-
quired for the initial (no RT) emission to converge (3×105)
10 Note that the pixels in the right panel of Figure 11 that give the
impression of a diffusion of the photons due to scattering possibly
larger than our ∼ 10% estimate, correspond to pixels with practically
zero number of photons.
Ly-α radiative transfer in cosmological simulations
21
is enough for the with RT and GP absorption results. And,
the results also converge if a 10 physical kpc radius is used
for the detailed RT and beyond that the GP absorption is
added. Convergence has been checked also with respect to
the minimum cell initial luminosity considered. We find
that the results converge if the minimum luminosity dis-
cussed in §3.2 in the context of initial emission convergence
is used.
3.5. Some additional physics considerations
Here we discuss the importance of collisions while the
photons are propagating, as well as the possible role of
dust (currently not taken into account).
3.5.1. Collisions
While the photons are undergoing scattering, collisions
should be considered in the following three contexts: (i)
collisional redistribution within the n = 2 state; if for ex-
ample a collision makes the atom go from the 2P3/2 to
the 2S1/2, then the Ly-α photon is destroyed through a 2
photon decay of the 2S1/2 state. If instead the collision
takes it to the 2P1/2, the scattering phase function will be
different and hence it is relevant in either case to see how
probable the collisional redistribution is (ii) collisional de-
excitation of the n = 2, in which case the photon is lost
(iii) collisional broadening of the line, which could cause
non-coherence in the rest frame of the atom. The RT code
can take all these processes into account, but here we de-
velop some intuition as to their importance. In fact, since,
as will be shown, these processes are in practice negligi-
ble, the corresponding calculations in the RT code were
switched off when producing the results presented in this
study.
Referring to cases (i) and (ii), the largest collisional cross
sections are for momentum changing transitions (∆L =
±1; e.g., Osterbrock 1989). As discussed already, both
collisions with electrons and protons are relevant, but pro-
tons are more significant in case (i), whereas electrons are
more significant in case (ii). We have already calculated
the probability per scattering that the 2P → 2S transition
of case (i) happens (see equation (26)). The maximum
value of this probability for the conditions of the simu-
lation cells is roughly 10−10 (assuming T4 = 1, np = 102
cm−3, with the latter being of the order of the maximum
proton number density of cells in simulations. The temper-
ature dependence is so weak that it does not really matter
what temperature one assumes, for order of magnitude
estimates). So, unless a photon undergoes 1010 scatter-
ings, collisions of the type (i) should not matter. The cells
that are relevant for this are optically thick cells where the
photons scatter repeatedly. Since as we saw Nsc ≃ τ0 and
none of the simulation cells has τ0 larger than a few times
109, collisions should not have a significant impact. Note
that for most cells the number of scatterings for which col-
lisions may start to matter is orders of magnitude higher
than 1010 (i.e., what is described above is the worst case
scenario as far as the effect of collisions is concerned since
it assumes the maximum proton number density, present
in very few cells). If these collisions do not matter then
collisions of type (ii), which have smaller cross sections,
should not matter either.
In case (iii), if the atom suffers collisions with other
particles while it is emitting, the phase of the emitted
radiation can be altered suddenly.
If the phase changes
completely randomly at the collision times, then informa-
tion about the emitting radiation is lost and coherence is
destroyed. In this case, in the rest frame of the atom, the
line profile is Lorentzian but the total width is the natural
width plus the frequency of collisions the atom experiences
on average. Since the importance of this effect as well is
determined by a comparison of the radiative decay time
and the time between collisions (i.e., equation 26), from
the above discussion it becomes clear that it is also negli-
gible.
3.5.2. Dust
Dust absorbs Ly-α photons. Thus, one would assume
that dust in the presence of scattering that traps photons,
could have a significant effect, and that this may be true
even if it is present in small amounts, as is expected to be
the case for the z ≃ 8 emitter we discuss (with a metal-
licity roughly equal to 0.1 the solar metallicity). Indeed,
Charlot & Fall (1991) found that only a tiny fraction of Ly-
α photons escape from a static, neutral ISM even if there
is a tiny amount of dust present. To include the effect of
dust absorption in simulations we will have to implement
a recipe to estimate the amount of dust. Even though
one can come up with an observationally motivated recipe
(albeit with unknown applicability at redshifts as high as
8), we postpone such a treatment for a future study, since
the main focus of the current study is the Ly-α RT scheme
(which nevertheless includes the probability per scattering
that the photon will be absorbed, but this probability is
currently set to 0).
However, the Ly-α emitter results we present in this
study should not be taken as unrealistic, since it is not
obvious how these results will change if we include the ef-
fects of dust. More specifically, many starforming galaxies
are observed to have significant Ly-α luminosities (e.g.,
Kunth et al. 1998; Pettini et al. 2000), and this is usually
attributed to the presence of galactic winds in these sys-
tems that allow the Ly-α photons to escape after much
fewer scatterings than in the static medium case. These
data seem to support the idea that it is the kinematics of
the gas rather than the dust content that is the dominant
Ly-α escape regulator.
Furthermore, Neufeld (1991) found that under suitable
conditions the effects of dust absorption may actually in-
crease rather than diminish the observed Ly-α line strength
relative to radiation that suffers little or no scattering.
This would happen for example in a multiphase medium
consisting of dusty clumps of neutral hydrogen embedded
within a relatively 'transparent' medium. If most of the
dust lies in cold neutral clouds then Ly-α photons, not be-
ing able to penetrate those clumps, will not be affected as
much by the presence of dust (see also Hansen & Oh 2005).
Although there is no direct observational evidence to sup-
port this structure for the ISM (i.e., that dust lies preferen-
tially in cold, neutral hydrogen clumps, even though the
clumpiness in the distribution of neutral hydrogen itself
seems to be established observationally (see Hansen & Oh
2005, and references therein)), such a morphology of the
dust and atomic hydrogen distribution could help account
for the lack of strong correlation between dust content --
inferred from metallicity or submillimeter emission -- and
22
TASITSIOMI
Ly-α equivalent width. For example, some dust-rich galax-
ies have substantially higher Ly-α escape fraction than less
dusty emitters (Kunth et al. 1998, 2003). In addition, Gi-
avalisco et al. (1996) found that there is no correlation
between the Ly-α equivalent widths and the slope of the
UV continuum, which is a measure of the continuum ex-
tinction and hence of dust content.
Another reason why it is not obvious how the results
presented here will change if we take dust into account,
is that in the current version of the ART code molecular
hydrogen forms only through the catalytic action of elec-
trons. When molecular hydrogen formation on grains is
included in the code, some of what is currently taken to
be neutral atomic hydrogen will transform into molecular
hydrogen, hence this effect will decrease the optical thick-
ness of what currently are the thickest cells.
4. summary
We develop a Ly-α RT code applicable to gasdynam-
ics cosmological simulations. High resolution, along with
appropriately treated cooling can lead to very optically
thick environments. Solving the Ly-α RT even for one
very thick simulation cell takes a long time. Solving it
for the whole simulation box, or a significant fraction of
it, takes unrealistic time. Thus, we develop accelerating
schemes to speed up the RT. We treat the moderately thick
cells by skipping the numerous core scatterings which are
not associated with any significant spatial diffusion, and
go directly to the scattering that takes the photon outside
of the core. We use depth dependent core definitions, and
find that quite large core values can be used. For the very
optically thick cells we motivate our treatment from the
classical problem of resonant radiation transfer in a semi-
infinite slab. We find that with some modifications, since
the simulations have cubic cells rather than slabs, we can
use the analytical solution derived by Neufeld (1990) for
the problem of the semi-infinite slab. With these acceler-
ating methods, along with the parallelization of the code
we made the problem of Ly-α RT in the complex environ-
ments of cosmological simulations tractable and solvable.
Even though our approach assumes a cell structure for
the simulation outputs, as is inherently the case in AMR
codes, the Ly-α RT code we discuss is applicable to out-
puts from all kinds of cosmological simulation codes. This
is true since one can always create an effective mesh by in-
terpolating the values of the various physical parameters.
We perform a series of tests of the RT code, and then
we apply it to ART cosmological simulations. We focus
on the brightest emitter in those simulations at z ≃ 8.
A first interesting result for this emitter pertains to its
intrinsic emission region and mechanisms. The emission
region consists of two smaller regions, each corresponding
to one of the two main progenitors that merged to form
the emitter at z ≃ 8. Both regions are pretty small, with
the larger of the two having a diameter of 2 − 2.5 physical
kpc. Furthermore, recombination produced Ly-α photons
is the dominant intrinsic Ly-α emission mechanism, with
collisional excitation and SNR produced Ly-α photons be-
ing subdominant. The intrinsic luminosity of the emitter
is 4.8 × 1043 ergs/s, whereas the injection spectrum (i.e.,
initial emission spectrum) shows significant velocity struc-
ture.
After performing the Ly-α RT, but before adding the
GP absorption, the emitter spectrum obtained resembles
that of a very optically thick static configuration, despite
the slight trend for inward radial motions. More specif-
ically, we obtain the usual double horn spectrum. This
happens because (i) even though there is some net inward
radial motion, there are still significant tangential peculiar
velocity components, and (ii) the optical depth is so high
that velocities of order some hundreds km/s will not fa-
vor blue versus red photons (i.e., in order to escape, both
kinds of photons have to shift off resonance much more
than the shift because of peculiar velocities, thus none of
the two kinds of photons is favored in particular because
of the existence of bulk motions). Namely, the velocity
information is in fact lost because of the extremely high
optical depth. The width of the two horns is noticeably
high (∼ 5000 km/s), but in agreement with what is ex-
pected for the high simulation column densities. The size
of the emitter increases, since the scatterings disperse the
photons on a larger area. We find that on the plane of the
emitter image, a photon on average escapes at a distance
of about 10% of the initial (before RT) emitter size from
the point it was originally emitted.
We include the GP absorption in two different ways:
without and with the red damping wing. In the first case
the spectrum is identical to that when the GP is not in-
cluded, with the difference that now we get only the red
peak (rather than both the red and blue peaks). This case
would correspond to the situation where the Ly-α emit-
ter lies within the HII region of a very bright quasar. In
the second case, where the damping wing is taken into ac-
count, the red peak is also affected. Its maximum flux is
suppressed compared to when no damping wing is used by
roughly a factor of 61.7. The resulting line after including
the wing is still quite broad with a velocity width of about
1350 km/s at a flux level of 10−21ergs s−1 cm−2 A−1, and
a FWHM of about 620 km/s. The line is quite displaced
redward from the Ly-α resonance, and reach a maximum
monochromatic flux of 10−20.2 ergs s−1 cm−2 A−1.
Attempting a detailed comparison with existing obser-
vations, or discussing detection prospects for an object
such as the simulated emitter is beyond the scope of this
study. We have studied only one emitter, and this for only
one direction of observation since our main goal was to use
it as an application for the Ly-α RT code. Thus, we do not
have a large enough and representative simulation sam-
ple yet. Furthermore, currently the highest redshift where
a Ly-α line has been observed is ∼ 6.6 (Kodaira et al.
2003)11 and it is not known how different the properties
of higher redshift emitters are from that of lower redshift
ones. The most recent report at z = 9 is that of Willis &
Courbin (2005). This study finds no detections. The sky
area coverage is possibly a significant factor contributing
to this no detection result. Instead, we content ourselves
here with a simple order of magnitude comparison. The
intrinsic Ly-α luminosity of our emitter is consistent with
luminosities reported in literature. For example, the high-
est Ly-α luminosity of the z = 5.7 sample of Hu et al.
11 The detection of a z = 10 Ly-α emitting galaxy was recently
reported by Pell´o et al. (2004) following a color selected survey for
z > 7 galaxies located behind a well studied gravitational lens clus-
ter, but the exact nature of this source remains contentious (e.g.,
Weatherley et al. 2004)
Ly-α radiative transfer in cosmological simulations
23
(2004) is roughly 6× 1043 ergs/s. Higher luminosities than
those have been inferred for Ly-α blobs, rather than emit-
ters. For example, the most luminous blob in the sample
of Matsuda et al. (2004) has a Ly-α luminosity of 1.1×1044
ergs/s. Most observed Ly-α emitters are unresolved and
so is expected to be the simulated emitter. Reported sizes
for the observed objects are in the ∼ few kpc range (e.g.,
Hu et al. 2002). Ly-α blobs on the other hand are quite
more extended, with sizes ∼ 100 kpc (Matsuda et al. 2004).
The widths of the (lower z) observed lines are typically a
few hundred km/s, whereas the FWHM of the simulated
line is roughly 620 km/s. As discussed already, the ve-
locity width of the ART emitter could be affected by the
very high H column densities which will drop as soon as
molecular hydrogen formation on dust grains is taken into
account. In terms of the detectability, if one adopts the
present day limit of ground based detections of ∼ 10−18
ergs s−1 cm−2 A−1, clearly our simulated emitter would
be orders of magnitude fainter. If the emitter is embedded
within the HII region of a bright quasar, in which case
the red damping wing will be suppressed, the brightness is
marginally below the sensitivity of current ground based
instruments. Note, however, that the prospects of detec-
tion will be much better for JWST which is expected to be
able to detect ∼ 400 times fainter objects than currently
studied with ground based infrared telescopes.
I am grateful to N. Y. Gnedin and A.V. Kravtsov for
many useful discussions and guidance, for comments on
the manuscript, and for allowing me to use their simula-
tions.
I would like to thank P. Jonsson, D. Neufeld, J.
Rhoads, Y. Tanigucchi, and Z. Zheng for fast and compre-
hensive responses to my questions. This work benefited
greatly from my interaction with J. Carlstrom, A. Konigl,
and A.V. Olinto, and was supported by the National Sci-
ence Foundation (NSF) under grants ASTR 02-06216 and
ASTR 02-39759, by NASA through grants NAG5-13274
and NAG5-12326, and by the Kavli Institute for Cosmo-
logical Physics at the University of Chicago. The au-
thor also acknowledges support through an award from
the Onassis Foundation. The simulations discussed were
performed on Linux Clusters and IBM690 arrays at the
National Center for Supercomputer Applications and the
San Diego Supercomputer Center under the National Part-
nership for Advanced Computational Infrastructure grant
#MCA03S023. This work was presented as part of a
dissertation to the Department of Astronomy and Astro-
physics, The University of Chicago, in partial fulfillment
of the requirements for the Ph.D. degree.
REFERENCES
Adams, T. F. 1972, ApJ, 174, 439
Ahn, S., Lee, H., & Lee, H. M. 2001, ApJ, 554, 604
-- . 2002, ApJ, 567, 922
Auer, L. H. 1968, ApJ, 153, 783
Avery, L. W. & House, L. L. 1968, ApJ, 152, 493
Barton, E. J., Dav´e, R., Smith, J.-D. T., Papovich, C., Hernquist,
L., & Springel, V. 2004, ApJ, 604, L1
Bouwens, R. J., Thompson, R. I., Illingworth, G. D., Franx, M., van
Dokkum, P., Fan, X., Dickinson, M. E., Eisenstein, D. J., & Rieke,
M. J. 2004, ApJ, 616, L79
Bower, R. G., Morris, S. L., Bacon, R., Wilman, R. J., Sullivan,
M., Chapman, S., Davies, R. L., de Zeeuw, P. T., & Emsellem, E.
2004, MNRAS, 351, 63
Cantalupo, S., Porciani, C., Lilly, S. J., & Miniati, F. 2005, astro-
ph/0504015
Cashwell, E. & Everett, C. 1959, Monte Carlo Method for Random
Walk Problems (New York: Pergamon Press)
Chandrasekhar, S. 1960, Radiative Transfer (New York: Dover
Publications)
Charlot, S. & Fall, S. M. 1991, ApJ, 378, 471
-- . 1993, ApJ, 415, 580
Delliou, M. L., Lacey, C., Baugh, C. M., Guiderdoni, B., Bacon, R.,
Courtois, H., Sousbie, T., & Morris, S. L. 2005a, MNRAS, 357,
L11
Delliou, M. L., Lacey, C., Baugh, C. M., & Morris, S. L. 2005b,
astro-ph/0508186
Dickinson, M. et al. 2004, ApJ, 600, L99
Dijkstra, M., Haiman, Z., & Spaans, M. 2005a, astro-ph/0510407
-- . 2005b, astro-ph/0510409
Fardal, M. A., Katz, N., Gardner, J. P., Hernquist, L., Weinberg,
D. H., & Dav´e, R. 2001, ApJ, 562, 605
Field, G. B. 1959, ApJ, 129, 551
Furlanetto, S. R., Schaye, J., Springel, V., & Hernquist, L. 2003,
ApJ, 599, L1
-- . 2005, ApJ, 622, 7
Giavalisco, M., Koratkar, A., & Calzetti, D. 1996, ApJ, 466, 831
Gnedin, N. Y. & Abel, T. 2001, New Astronomy, 6, 437
Gnedin, N. Y. & Prada, F. 2004, ApJ, 608, L77
Gould, A. & Weinberg, D. H. 1996, ApJ, 468, 462
Haiman, Z. & Spaans, M. 1999, ApJ, 518, 138
Haiman, Z., Spaans, M., & Quataert, E. 2000, ApJ, 537, L5
Hamilton, D. R. 1940, Physical Review, 58, 122
Hansen, M. & Oh, S. P. 2005, astro-ph/0507586
Harrington, J. P. 1973, MNRAS, 162, 43
Hu, E. M., Cowie, L. L., Capak, P., McMahon, R. G., Hayashino,
T., & Komiyama, Y. 2004, AJ, 127, 563
Hu, E. M., Cowie, L. L., McMahon, R. G., Capak, P., Iwamuro, F.,
Kneib, J.-P., Maihara, T., & Motohara, K. 2002, ApJ, 568, L75
Hui, L. & Gnedin, N. Y. 1997, MNRAS, 292, 27
Hui, L., Gnedin, N. Y., & Zhang, Y. 1997, ApJ, 486, 599
Hummer, D. G. 1962, MNRAS, 125, 21
Jefferies, J. T. & White, O. R. 1960, ApJ, 132, 767
Kodaira, K. et al. 2003, PASJ, 55, L17
Kravtsov, A. V. 2003, ApJ, 590, L1
Kravtsov, A. V. & Gnedin, O. Y. 2005, ApJ, 623, 650
Kunth, D., Leitherer, C., Mas-Hesse, J. M., Ostlin, G., & Petrosian,
A. 2003, ApJ, 597, 263
Kunth, D., Mas-Hesse, J. M., Terlevich, E., Terlevich, R., Lequeux,
J., & Fall, S. M. 1998, A&A, 334, 11
Loeb, A. & Rybicki, G. B. 1999, ApJ, 524, 527
Maier, C. 2002, Ph.D. Thesis, Naturwissenschaftlich-Mathematische
Gesamtfakultat der Universitat Heidelberg, Germany
Matsuda, Y. et al. 2004, AJ, 128, 569
Neufeld, D. A. 1990, ApJ, 350, 216
-- . 1991, ApJ, 370, L85
Osterbrock, D. E. 1962, ApJ, 135, 195
-- . 1989, Astrophysics of Gaseous Nebulae and Active Galactic
Nuclei (Mill Valey, CA: Univ. Sci.)
Partridge, R. B. & Peebles, P. J. E. 1967, ApJ, 147, 868
Pell´o, R., Schaerer, D., Richard, J., Le Borgne, J.-F., & Kneib, J.-P.
2004, A&A, 416, L35
Pettini, M., Steidel, C. C., Adelberger, K. L., Dickinson, M., &
Giavalisco, M. 2000, ApJ, 528, 96
Phillips, K. C. & Meszaros, P. 1986, ApJ, 310, 284
Rhoads, J. E., Dey, A., Malhotra, S., Stern, D., Spinrad, H., Jannuzi,
B. T., Dawson, S., Brown, M. J. I., & Landes, E. 2003, AJ, 125,
1006
Santos, M. R. 2004, MNRAS, 349, 1137
Santos, M. R., Ellis, R. S., Kneib, J., Richard, J., & Kuijken, K.
2003, ApJ, 606, 683
Shull, J. M. & Silk, J. 1979, ApJ, 234, 427
Spergel, D. N. et al. 2003, ApJS, 148, 175
Spitzer, L. 1978, Physical Processes in the Interstellar Medium (New
Yorl: Wiley)
Stanway, E. R., Bunker, A. J., McMahon, R. G., Ellis, R. S., Treu,
T., & McCarthy, P. J. 2004, ApJ, 607, 704
Steidel, C. C., Adelberger, K. L., Shapley, A. E., Pettini, M.,
Dickinson, M., & Giavalisco, M. 2000, ApJ, 532, 170
Taniguchi, Y. et al. 2005, PASJ, 57, 165
Tasitsiomi, A. 2006, astro-ph/0601562
Unno, W. 1952, PASJ, 4, 100
Weatherley, S. J., Warren, S. J., & Babbedge, T. S. R. 2004, A&A,
428, L29
Willis, J. P. & Courbin, F. 2005, MNRAS, 357, 1348
Wouthuysen, S. A. 1952, AJ, 57, 31
Yusef-Zadeh, F., Morris, M., & White, R. L. 1984, ApJ, 278, 186
Zanstra, H. 1949, Bull. Astron. Inst. Netherlands, 11, 1
Zheng, Z. & Miralda-Escud´e, J. 2002, ApJ, 578, 33
|
0804.3823 | 2 | 0804 | 2008-08-07T06:52:10 | A Global Stability Analysis of Clusters of Galaxies with Conduction and AGN Feedback Heating | [
"astro-ph"
] | We investigate a series of steady-state models of galaxy clusters, in which the hot intracluster gas is efficiently heated by active galactic nucleus (AGN) feedback and thermal conduction, and in which the mass accretion rates are highly reduced compared to those predicted by the standard cooling flow models. We perform a global Lagrangian stability analysis. We show for the first time that the global radial instability in cool core clusters can be suppressed by the AGN feedback mechanism, provided that the feedback efficiency exceeds a critical lower limit. Furthermore, our analysis naturally shows that the clusters can exist in two distinct forms. Globally stable clusters are expected to have either: 1) cool cores stabilized by both AGN feedback and conduction, or 2) non-cool cores stabilized primarily by conduction. Intermediate central temperatures typically lead to globally unstable solutions. This bimodality is consistent with the recently observed anticorrelation between the flatness of the temperature profiles and the AGN activity (Dunn & Fabian 2008) and the observation by Rafferty et al. (2008) that the shorter central cooling times tend to correspond to significantly younger AGN X-ray cavities. | astro-ph | astro-ph |
Draft version November 1, 2018
Preprint typeset using LATEX style emulateapj v. 03/07/07
A GLOBAL STABILITY ANALYSIS OF CLUSTERS OF GALAXIES WITH CONDUCTION AND AGN
FEEDBACK HEATING
Fulai Guo1, S. Peng Oh 1 and M. Ruszkowski2
Draft version November 1, 2018
ABSTRACT
We investigate a series of steady-state models of galaxy clusters, in which the hot intracluster gas
is efficiently heated by active galactic nucleus (AGN) feedback and thermal conduction, and in which
the mass accretion rates are highly reduced compared to those predicted by the standard cooling flow
models. We perform a global Lagrangian stability analysis. We show for the first time that the global
radial instability in cool core clusters can be suppressed by the AGN feedback mechanism, provided
that the feedback efficiency exceeds a critical lower limit. Furthermore, our analysis naturally shows
that the clusters can exist in two distinct forms. Globally stable clusters are expected to have either:
1) cool cores stabilized by both AGN feedback and conduction, or 2) non-cool cores stabilized pri-
marily by conduction. Intermediate central temperatures typically lead to globally unstable solutions.
This bimodality is consistent with the recently observed anticorrelation between the flatness of the
temperature profiles and the AGN activity (Dunn & Fabian 2008) and the observation by Rafferty
et al. (2008) that the shorter central cooling times tend to correspond to significantly younger AGN
X-ray cavities.
Subject headings: conduction -- cooling flows -- galaxies: clusters: general -- galaxies: active -- insta-
bilities -- X-rays: galaxies: clusters
1. INTRODUCTION
Clusters of galaxies are the largest gravitationally
bound systems in the universe. They are filled with
hot gas with T ∼ 2 − 10 keV, which loses thermal en-
ergy prolifically by emitting X-rays. The X-ray surface
brightness of many galaxy clusters shows a strong cen-
tral peak that was previously interpreted as the signa-
ture of a cooling flow with mass accretion rates of up
to several hundred M⊙ yr−1 (see Fabian 1994, for a re-
view). Although the gas temperature is observed to de-
cline toward cluster centers, recent high-resolution Chan-
dra and XMM-Newton observations show a remarkable
lack of emission lines from the gas at temperatures be-
low about ∼ 1/3 of the ambient cluster temperature
(e.g., Peterson et al. 2001, 2003; Tamura et al. 2001; for
a review see Peterson & Fabian 2006). In addition, the
spectroscopically determined mass deposition rates are
significantly smaller than the classic values estimated
from the X-ray luminosity within the cooling regions
(Voigt & Fabian 2004). The absence of a cool phase in
cores of galaxy clusters is suggestive of one or more heat-
ing mechanisms maintaining the hot gas at keV temper-
atures for a period at least comparable to the lifetime of
galaxy clusters.
Amongst the many candidate heating mechanisms put
forth recently, there are two leading contenders:
1. thermal
regions
ter
conduction
of
(e.g.,
from the
to
cluster
the
Bertschinger & Meiksin
hot
the
outer
cen-
1986;
1 Department of Physics, University of California, Santa Bar-
bara, CA 93106, USA;
E-mail:
(SPO)
[email protected] (FG); [email protected]
2 Department of Astronomy, The University of Michigan, 500
Church Street, Ann Arbor, MI 48109, USA;
E-mail: [email protected]
Zakamska & Narayan
Voigt & Fabian 2004);
2003,
hereafter ZN03;
2. heating of the intracluster medium (ICM) by out-
flows, bubbles, or cosmic rays generated by AGNs
at cluster centers (e.g., Bruggen & Kaiser 2002,
Ruszkowski & Begelman 2002, Ruszkowski et al.
2004, Chandran & Rasera 2007, Guo & Oh 2008,
Sijacki et al. 2008).
2003)
Recent
numerical work
and
2001; Cho et al.
(e.g.
theoretical
Narayan & Medvedev
has
shown that a turbulent magnetic field is not as efficient
in suppressing thermal conduction as previously thought.
In particular, Narayan & Medvedev (2001) showed that
the effective thermal conductivity κ in a turbulent MHD
medium is a substantial fraction (∼ 1/5) of the classical
Spitzer value κSp if magnetic turbulence extends over
at least two decades in scale. On the other hand,
recent work shows that bouyancy instabilities could
potentially strongly suppress conductivity in the cluster
core (Quataert 2008; Parrish & Quataert 2008), a point
we discuss in §4. Following this work, ZN03 shows that
the electron density and temperature profiles of half of
the clusters they investigated can be fitted by a pure
conduction model with the conductivity suppression
factor f ≡ κ/κSp ∼ 0.2 − 0.4. However, if only thermal
conduction operates to balance the cooling, extreme
fine-tuning of the conduction suppression factor f is
required (Guo & Oh 2008; also see Bregman & David
1988):
if f is too low, then a strong cooling flow
develops, while if f is too high, the temperature profile
becomes nearly isothermal, in contrast to observations
of cool core clusters where the temperature invariably
declines toward the cluster center. Furthermore, al-
though thermal conduction is well known to stabilize
short-wavelength perturbations against
in-
stability (e.g., Field 1965; Malagoli et al. 1987), the
thermal
2
Guo et al.
pure conduction models of the cool core clusters are
thermally unstable against global perturbations (Soker
2003; Kim & Narayan 2003, hereafter KN03). Using a
Lagrangian perturbation analysis, KN03 showed that
the pure conduction model has one globally unstable
radial mode with the instability growth (e-folding) time
of ∼ 2 − 5 Gyr. Furthermore, if strong perturbations
are applied (as would be the case in, for instance, a
cluster merger), the growth times can be even shorter.
Guo & Oh (2008) showed that if one started from
arbitrary initial conditions (rather than an equilibrium
solution), a catastrophic cooling flow quickly develops
in a conduction-only model with a moderate level of
conductivity.
Fortunately, other sources of heating exist. A par-
ticularly promising candidate is heating by the central
AGN, for which observational evidence has been grow-
ing in recent years (see McNamara & Nulsen 2007 for a
recent review). A majority (∼ 71%) of cool core clus-
ters harbor radio sources at their cluster centers (Burns
1990). Following the launch of Chandra and XMM-
Newton, recent high-resolution X-ray observations also
indicate that these radio sources are interacting with
their surroundings and often displace the ICM, produc-
ing X-ray cavities (e.g., Fabian et al. 2000, Bırzan et al.
2004, Forman et al. 2007).
By contrast with conduction, heating by the dis-
sipation of mechanical energy released by central
AGNs provides a self-regulating feedback mechanism
(e.g., Ciotti & Ostriker 2001, Ruszkowski & Begelman
2002, Brighenti & Mathews 2003, Kaiser & Binney 2003,
Guo & Oh 2008).
If AGN activity is triggered by
cooling-induced gas accretion toward cluster centers,
the AGN heating increases until it halts further accre-
tion. Thus, the gas accretion rate is self-regulated as
brief bursts of AGN activity alternate with cooling (e.g.,
Voit & Donahue 2005). Due to this AGN feedback heat-
ing, the accretion flow may automatically adjust itself to
a low value of the accretion rate, which depends mainly
on the feedback efficiency ǫ (see equation 12), and, in
a time-averaged sense, the ICM may reach a quasi-
equilibrium state. This has been clearly demonstrated by
Ruszkowski & Begelman (2002, hereafter RB02) in hy-
drodynamic simulations, who showed that a model clus-
ter heated by a combination of thermal conduction and
AGN feedback does not suffer from the cooling catas-
trophe, but instead relaxes to a stable quasi-equilibrium
state. More recently, Guo & Oh (2008) proposed a new
model of AGN feedback heating, where the ICM is effi-
ciently heated by both thermal conduction and the cos-
mic rays produced by accretion-triggered AGN activity.
In their model, the ICM also relaxes to a stable steady
state with the mass accretion rate highly reduced, and,
more importantly, their results do not require fine tuning
of the various adjustable parameters, including thermal
conductivity and the AGN heating efficiency. Moreover,
unlike the conduction-only case, the simulation relaxes to
a stable state independent of the initial conditions. Al-
though the detailed dynamics of how the released AGN
energy is transferred into thermal energy of the ICM may
be much more complicated than these models and is still
poorly understood at the present time, these simulations
strongly indicate that AGN feedback heating may poten-
tially solve the fine-tuning problem associated with the
pure conduction model.
These simulations also suggest that AGN feedback
heating plays a key role in suppressing global thermal
instability in the ICM (see Rosner & Tucker 1989 for a
local analysis). While local thermal instability may only
produce small-scale structures (e.g., local mass dropout,
emission-line filaments) in galaxy clusters, global ther-
mal instability may result in a cooling catastrophe and
a strong cooling flow. Thus, a successful model for the
ICM must be globally stable, or at least only have in-
stabilities which grow on extremely long timescales. In
the present paper, we will use the Lagrangian pertur-
bation method to formally investigate thermal
insta-
bility in quasi-equilibrium galaxy clusters with thermal
conduction and AGN feedback heating. Global stabil-
ity analysis is a method complementary to numerical
simulations as it allows for the quick identification of
global trends, quick systematic parameter search and
helps to build physical intuition. For the spatial dis-
tribution of AGN feedback heating, we adopt the ana-
lytically tractable model proposed by Begelman (2001).
This model has been compared with observations by
Piffaretti & Kaastra (2006), who show that the model
usually provides a satisfactory explanation of the ob-
served structure of cool core clusters, although in a fair
fraction of their sample the model provide relatively poor
fits. However, our emphasis is not on the background so-
lutions of this particular model but their global stability.
With slight modification, our methods can be applied to
any model of AGN feedback heating.
We will show that the feedback mechanism can indeed
effectively suppress global radial thermal instability in
cool core (CC) clusters, provided that the AGN feed-
back efficiency is larger than a lower limit (see § 3.3 and
§ 3.4 for details). We will also study the dependence
of the cluster stability on the background ICM profiles
(§ 3.5): for non-cool core (NCC) clusters, which have rel-
atively flat temperature profiles and which are less stud-
ied in the literature, the stabilizing effect of the feedback
mechanism becomes small, but thermal conduction may
completely suppress global thermal instability. Thus, we
propose that thermal stability of the ICM favors two dis-
tinct categories of cluster steady state profiles: CC clus-
ters stabilized mainly by AGN feedback and NCC clus-
ters stabilized by thermal conduction. Interestingly, X-
ray observations suggest that clusters can be subdivided
into two distinct categories according to the presence or
absence of a cool core (e.g., Peres et al. 1998, Bauer et al.
2005, Sanderson et al. 2006, Chen et al. 2007). Our sta-
bility analysis thus naturally explains these two distinct
cluster categories. Our model is also consistent with the
observation by Rafferty et al. (2008) who show that the
short central cooling time corresonds to younger AGN
(i.e., shorter X-ray cavity ages) and the anticorrelation
between the flatness of the temperature profiles and the
AGN activity recently reported by Dunn & Fabian 2008.
The issue of the formation and evolution of NCC and
CC clusters has recently been addressed by Burns et al.
(2008) who performed large scale cosmological simula-
tions. They found out that the CC and NCC clusters
follow different evolutionary tracks, with CC clusters ac-
creting more slowly over time and growing enhanced cool
cores via hierarchical mergers. In contrast, they argued
that NCC suffered early mergers that disrupted embry-
A Global Stability Analysis of Galaxy Clusters
3
onic cool cores. However, this pioneering work does not
include the effects of AGN to stop catastrophic cooling in
the centers of CCs and the numerical resolution in their
simulations is still too low (15.6h−1 kpc) to accurately
study the stability and structure of the cores once they
are formed. McCarthy et al.
(2008) invoked different
preheating histories to explain the difference between CC
and NCC clusters. Like us, they find that AGN heating
is required to stabilize CC clusters. Our present calcu-
lations make the new suggestion that the bifircation be-
tween CC and NCC clusters emerges naturally, from the
fact that clusters with intermediate central temperatures
are globally unstable.
The rest of the paper is organized as follows. In § 2,
we describe the time-dependent equations of the ther-
mal intracluster gas and construct a series of steady-
state cluster models, in which the mass accretion rate
is highly suppressed compared to that predicted by stan-
dard cooling flow models. We then carry out a detailed
formal linear stability analysis of local and global modes
of thermal instability in steady-state galaxy clusters in
§ 3, where we also study the dependence of the clus-
ter stability on the AGN feedback efficiency and on the
background cluster profiles. We summarize our main re-
sults in § 4 with a discussion of the implications. The
cosmological parameters used throughout this paper are:
Ωm = 0.3, ΩΛ = 0.7, h = 0.7. We have rescaled ob-
servational results if the original paper used a different
cosmology.
where κ is the effective isotropic conductivity. Depending
on the details of plasma magnetization and MHD turbu-
lence, heat transport in the ICM may be very complex,
and both electron conduction and turbulent mixing may
contribute to heat transport (e.g., Lazarian 2006); ad-
ditionally various instabilities could alter the nature of
conductivity within the cooling region (Quataert 2008;
Parrish & Quataert 2008). Since at present there is no
consensus on the nature of conductivity in a turbulent
magnetized plasma, we adopt the same assumption of
Spitzer conductivity (with a factor f due to magnetic
field suppression) that most authors do (e.g., ZN03,
KN03),
κ = f κSp,
(5)
where κSp is the classical Spitzer conductivity (Spitzer
1962),
κSp =
1.84 × 10−5
ln λ
T 5/2ergs s−1K−7/2cm−1,
(6)
with the usual Coulomb logarithm lnλ ∼ 37.
In this
paper, we assume that f (0 ≤ f ≤ 1) is constant in both
space and time.
According to the ideal gas law, the gas pressure is re-
lated to the gas temperature T and the electron number
density ne via
2. STEADY-STATE MODELS
2.1. Time-dependent equations
P =
ρkBT
µmµ
=
µe
µ
nekBT,
(7)
In our model, the intracluster medium is subject to ra-
diative cooling, AGN feedback heating and thermal con-
duction. The governing hydrodynamic equations are
dρ
dt
ρ
dv
dt
+ ρ∇ · v = 0,
= −∇P − ρ∇Φ,
(1)
(2)
1
γ − 1
dP
dt
−
γ
γ − 1
P
ρ
dρ
dt
= H − ∇ · F − ρL,
(3)
where d/dt ≡ ∂/∂t + v · ∇ is the Lagrangian time
derivative, ρ is the gas density, P is the gas pres-
sure, v is the gas velocity, Φ is the gravitational po-
tential, γ = 5/3 is the adiabatic index of thermal gas,
ρL = n2
eT 1/2 ergs cm−3 s−1 is
the volume cooling rate due to thermal bremsstrahlung3
(Rybicki & Lightman 1979; KN03), and F is the conduc-
tive heat flux
eΛ(T ) = 2.1 × 10−27n2
F = −κ∇T,
(4)
3 We have simplified the form of the cooling function here, ig-
noring the contribution of metal line cooling, which is important
at lower temperatures. Fits to the full cooling function do ex-
ist (Sutherland & Dopita 1993), but they complicate the analytic
derivation of the global stability analysis. We have experimented
with the full cooling function and didn't find qualitative changes
to our results (see Fig. 11).
where kB is Boltzmann's constant, mµ is the atomic mass
unit, and µ and µe are the mean molecular weight per
thermal particle and per electron, respectively. As in
ZN03, we use µ = 0.62 and µe = 1.18, corresponding to
a fully ionized gas with hydrogen fraction X = 0.7 and
helium fraction Y = 0.28.
In equation (2), we neglect the self-gravity of the gas
and any dynamical effects of magnetic fields. The gravi-
tational potential Φ is determined by the dark matter dis-
tribution ρDM, which we assume has a modified Navarro-
Frenk-White (NFW) form (Navarro et al. 1997) with a
softened core (ZN03):
ρDM(r) =
M0/2π
(r + rc)(r + rs)2 ,
(8)
where rc is a softening radius which introduces a core in
the dark matter distribution in the very inner regions, rs
is the standard scale radius of the NFW profile and M0
is a characteristic mass. The corresponding gravitational
potential is:
Φ = −2GM0
rc
(rs − rc)2 (cid:20)ln
1 + r/rc
1 + r/rs
+
ln(1 + r/rc)
r/rc
−2GM0
rs(rs − 2rc)
rc(rs − rc)2
ln(1 + r/rs)
r/rc
,
(cid:21)
(9)
where G is the gravitational constant. The parameters
M0 and rs for a given cluster are obtained from the ob-
served temperature in the outer regions of the cluster,
4
Guo et al.
as described in ZN03. We adopt their calculated values
of these two parameters and their best-fit values of rc
directly.
The term H in equation (3) is the volume heating
rate due to AGN feedback. We adopt the "effervescent
heating" mechanism proposed by Begelman (2001) to de-
scribe the energy deposition into the ICM by the rising
bubbles, which are produced by the central AGN. Be-
cause of the non-negligible gas pressure gradient in the
ICM, the bubbles will expand as they rise, doing pdV
work and converting the internal bubble energy to kinetic
energy of the ICM. The resulting disorganized motion of
the ICM is quickly converted to heat. Assuming that this
heating mechanism reaches a quasi-steady state, the de-
tails of the bubble filling factor, rise rate and geometry
should cancel.
If the cavity expands adiabatically (see
Guo & Oh (2008) for an alternative scenario), the lumi-
nosity passing through the surface of a sphere at radius
r is:
E ∝ pb(r)(γb−1)/γbr,
(10)
where pb(r) is the pressure of buoyant gas inside bubbles,
γb ≈ 4/3 is the adiabatic index of buoyant gas (assuming
it is primarily composed of relativisitic plasma), and r is
the unit vector along the radial direction. Assuming that
the bubble rises subsonically so that pressure equilibrium
is maintained, pb(r) = P (r), where P (r) is the thermal
pressure of the ICM, we may rewrite equation (10) as:
E ∼ Lagn(cid:18) P
P0(cid:19)β
r,
(11)
where P0 is the gas pressure at the cluster center, Lagn
is the AGN mechanical luminosity, and β = (γb − 1)/γb.
AGN activity is likely to be intermittent on a timescale
of order the Salpeter time tS ∼ 107 yr, and possibly
as short as ti ∼ 104 − 105 yr (Reynolds & Begelman
1997). Note that the bubble rise time is typically compa-
rable to (at most several times) the sound crossing time
tsc ∼ 108r100c−1
s,1000yr for a radius r ∼ 100r100 kpc and
sound speed cs ∼ 1000cs,1000 km s−1 (e.g., see table 3 in
Bırzan et al. (2004)), and is usually shorter than the gas
cooling time. Thus, it is justifiable to treat AGN heating
in a time-averaged sense and assume that the mechanical
energy of central AGN is injected into the whole ICM in-
stantaneously (e.g., RB02; Brighenti & Mathews 2003).
We further assume the AGN mechanical luminosity to
be
Lagn = −ǫ Minc2,
(12)
where ǫ is the kinetic efficiency of AGN feedback, and
Min = 4πr2
inρ0v0 is the mass accretion rate at the inner
radius rin, where ρ0 and v0 are the density and radial
velocity of thermal gas at rin, respectively. Therefore,
the volume AGN heating rate H may be written as
H ∼ −∇ ·
E
4πr2
∼
ǫβ Minc2
4πr3 (cid:16)1 − e−r/r0(cid:17)(cid:18) P
P0(cid:19)β ∂ ln P
∂ ln r
,
(13)
where r0 is the inner heating cutoff radius, which is de-
termined by the finite size of the central radio source
(Ruszkowski & Begelman 2002; also see a discussion of
r0 in Roychowdhury et al. 2005). In the rest of this pa-
per, r0 is taken to be 20 kpc, unless otherwise stated.
2.2. Steady-state models
In this subsection, we will construct steady-state clus-
ter models, which will be used as initial unperturbed
states in stability analysis presented in the next section.
We assume that the cluster is spherically symmetric and
time independent. Equations (1), (2) and (3) are thus
simplified to
M ≡ 4πr2ρv = constant,
ρv
dv
dr
= −
dP
dr
− ρ
dΦ
dr
,
(14)
(15)
v
γ − 1
dP
dr
−
γ
γ − 1
P v
ρ
dρ
dr
= H −
1
r2
d
dr
(r2F ) − ρL,(16)
where v is the radial gas speed and F is the radial heat
flux.
Equations (4), (15) and (16) are three ordinary differ-
ential equations for the three variables P (r), T (r) and
r2F (r). We solve these equations as a boundary value
problem between r = rin and rout, where we impose the
boundary conditions:
ne(rin) = n0,
T (rin) = Tin,
T (rout) = Tout,
r2
inF (rin) = 0.
(17)
The first three conditions are taken from either the best-
fit models of ZN03 or the observational data (see Table
1), while the last condition ensures that there are no
sources or sinks of heat at the cluster center. Equations
(4), (15), (16) and boundary conditions (17) form an
M as the eigen-
eigenvalue problem with either f , ǫ or
value, provided that the other two are given as free model
parameters. The results of the steady-state cluster mod-
els presented in this subsection are not sensitive to the
specific choices for the value of rin and rout; here we
choose rin = 1 kpc and rout = 1000 kpc, unless otherwise
stated. We note that the cluster stability does depend on
the value of rin; this is degenerate with the choice of Tin,
which is explored in § 3.5. We have experimented with
different values of rout for our models (e.g., rout is taken
to be 300 kpc in models of A2597; see Table 1), and find
that the results of stability analysis are not sensitive to
this choice. This is consistent with the extremely long
gas cooling time in the cluster outer regions.
To illustrate our results clearly, we need to define
several physical quantities for each steady-state cluster
A Global Stability Analysis of Galaxy Clusters
5
Parameters and results of the steady-state models for typical cool core clusters
TABLE 1
Name
A1795
a
Tin
(keV)
2
a
Tout
(keV)
7.5
a
n0
(cm−3) Model
0.053 d
A2199
1.6
5
0.074
A2052
1.3
3.5
0.035 e
A2597 f
1
4
0.07 g
M
(M⊙/yr)
−0.15
−0.15
−0.05
−0.015
−0.015
−0.00375
−0.006
−0.006
−0.0015
−0.32
−0.32
−0.16
−0.17
−0.14
ǫ
0
0.1
0.3
0
0.05
0.2
0
0.05
0.2
0
0.05
0.1
0.1
0.1
f
0.27
0.12
0.12
0.43
0.36
0.36
0.31
0.20
0.20
1.30
0.40
0.40
0.30
0.50
b
nout
(cm−3)
1.42 × 10−4
1.26 × 10−4
1.26 × 10−4
4.22 × 10−5
4.00 × 10−5
4.00 × 10−5
1.78 × 10−5
1.62 × 10−5
1.62 × 10−5
1.39 × 10−3
1.31 × 10−3
1.31 × 10−3
1.30 × 10−3
1.33 × 10−3
hagn/LX
c
0
0.52
0.52
0
0.12
0.12
0
0.29
0.29
0
0.68
0.68
0.76
0.60
A1
A2
A3
B1
B2
B3
C1
C2
C3
D1
D2
D3
D4
D5
a These boundary values are adopted from the best-fit models of ZN03, unless otherwise stated.
bnout is the corresponding model electron number density at the outer boundary rout.
chagn/LX is the ratio of the overall AGN heating rate to X-ray luminosity of the cluster.
d Adopted from the Chandra observation (Ettori et al. 2002).
e Adopted from the Chandra observation (Blanton et al. 2001).
fTo provide a better fit to observations, we take the outer boundary of the cluster A2597 to be rout = 300 kpc and the
AGN heating cutoff radius to be r0 = 40 kpc.
g Adopted from the Chandra observation (McNamara et al. 2001).
model. We first define the volume-integrated AGN heat-
ing rate as
hagn = Z rout
rin
4πr2Hdr,
(18)
and the X-ray luminosity as
LX = Z rout
rin
4πr2ρLdr.
(19)
A cluster without any heating source will lose its ther-
mal energy by emitting X-rays. We can thus define the
isobaric cooling time from equation (3) as:
tcool ≡
γ
γ − 1 (cid:18) P
ρL(cid:19) .
(20)
Table 1 lists the physical parameters and results of the
steady-state models for four typical cool core clusters.
Note that the mass accretion rates in our steady-state
models are usually much less than the Eddington rate
Med ≈ 26(Mbh/109M⊙)(η/0.1)−1 M⊙/yr, where Mbh is
the mass of the supermassive black hole at the cluster
center and η is the radiative efficiency of AGN accre-
tion. We take the cluster Abell 1795 (Ettori et al. 2002)
as our fiducial cluster. In the first two models (A1 and
A2), the values of ǫ and M are given and f is obtained
as the eigenvalue. Model A1 is a pure conduction model
(ǫ = 0), while model A2 is a typical hybrid model with
both thermal conduction and AGN feedback heating. In
model A3, the AGN feedback efficiency ǫ is chosen to
be much larger than that in model A2. We further as-
sume that, in addition to the boundary conditions (17),
the electron density at the outer boundary is fixed (we
choose the same as that in model A2). Thus, with the
extra boundary condition, both M and f can be solved
as eigenvalues. The radial steady-state profiles of Abell
1795 are presented in Figure 1. As can clearly be seen,
the electron density and temperature profiles of these
three models fit the observational data quite well. The
profiles of models A2 and A3 are virtually the same, since
the gas density and temperature at both the inner and
outer boundaries are exactly the same for these two mod-
els. Using these cluster profiles and equation (20), we can
then calculate the radial profiles of the isobaric cooling
time, which are shown in the lower left panel of Figure
1. Obviously, tcool is much less than the age of the Uni-
verse within ∼ 100 kpc from the cluster center, suggest-
ing that the radiative cooling is dynamically important
in the central regions of the cluster. We show the relative
importance of AGN heating and thermal conduction in
the lower right panel of Figure 1. For our hybrid models,
while thermal conduction is significant in the outer parts
of the cluster cool core, AGN heating clearly dominates
at the center. Thus the gas temperature profile in the in-
nermost regions (. 10 kpc) of the cluster is flatter than
that in the pure conduction model, as clearly seen in the
upper right panel of Figure 1.
Since we will find later that the dependence of the clus-
ter stability on ǫ varies somewhat with cluster proper-
ties, particularly the density and temperature profiles,
we choose the cluster Abell 2199 (Johnstone et al. 2002)
as our second fiducial cluster and plot its radial steady-
state profiles in Figure 2. The gas cooling time in its
central cool core (r . 100 kpc) is clearly less than the
Hubble time.
for
ZN03 found that,
some cool core clusters,
conduction-only models may require implausibly high
values of thermal conductivity. We apply our model to
one of these clusters, Abell 2597. As shown in Table
1, the conduction-only model (D1) for A2597 requires
6
Guo et al.
Fig. 1. -- Electron number density (upper left), temperature (upper right), isobaric cooling time (lower left), and relative importance
of AGN heating and conduction (lower right) in three typical steady-state models of the cluster Abell 1795. The dot-dashed line in the
lower left panel shows the age of the universe (13.5 Gyr for the cosmology used in this paper). Crosses in the upper panels correspond to
Chandra data (Ettori et al. 2002). See text and Table 1 for additional information.
a thermal conductivity of f = 1.3. By including AGN
feedback heating, the required thermal conductivity may
be much smaller (e.g., f = 0.4 in model D3). Figure 3
shows radial profiles of electron temperature and num-
ber density in our models for A2597 (model D1 - D5),
which all fit the data reasonably well. We note that
some cool core clusters may not be well explained by
models with conduction and AGN "effervescent" heating
(see Piffaretti & Kaastra 2006). However, they may be
explained by models with a more realistic AGN heating
function, and our method of global stability analysis is
generally applicable to any steady-state AGN feedback
models with only slight modifications due to the new
AGN heating function.
For each cluster, the first model in Table 1 is the pure
conduction model, where the cooling is entirely balanced
by thermal conduction. This model determines the max-
imum value of f for each cluster: as the level of AGN
heating increases, the required thermal conductivity de-
creases.
In the current paper, we only present our re-
sults from some typical models with specific values of
f (e.g., f = 0.12 in model A3 for the cluster A1795),
but the conclusions drawn are general to models with
different levels of thermal conduction (also see Table 3
of Piffaretti & Kaastra 2006). As an example, we addi-
tionally present models with various values of f for the
cluster A2597 (models D4 and D5) in Table 1 and Figure
3 (also see Figure 7).
We note that a minimum level of thermal conduction is
usually required in our steady-state models, since H/ρL
is not uniform with radius, as readily seen in the lower
right panels of Figure 1 and 2. In other words, the "ef-
fervescent" AGN heating mechanism (equation 13) alone
could not offset cooling at all radii throughout the clus-
ter. Nonetheless, as we have shown, a combination of a
moderate level (f & fmin, where fmin ∼ 0.1 for A1795
and ∼ 0.2 for A2597) of thermal conduction and the "ef-
fervescent" AGN heating can balance cooling throughout
the entire cluster, and the resulting steady-state profiles
of electron density and temperature fit the observational
data quite well. As f decreases, the heating from thermal
conduction decreases, and thus the required AGN heat-
ing increases. When f . fmin, the required AGN heating
in steady-state models with the boundary conditions (17)
usually overheats the cluster central regions, while the
gas temperature and entropy start to drop rapidly with
increasing radius, which is not consistent with X-ray ob-
servations. Note that the requirement on conductivity
(f & fmin) may be alleviated if other forms of AGN
heating (e.g., viscous dissipation of sound waves, see
Ruszkowski et al. 2004; shock heating, see Bruggen et al.
2007; cosmic ray feedback, see Guo & Oh 2008) are taken
A Global Stability Analysis of Galaxy Clusters
7
Fig. 2. -- Electron number density (upper left), temperature (upper right), isobaric cooling time (lower left), and relative importance
of AGN heating and conduction (lower right) in three typical steady-state models of the cluster Abell 2199. Crosses in the upper panels
correspond to Chandra data (Johnstone et al. 2002). See text and Table 1 for additional information.
into account, since a substantial fraction of sound waves,
shocks or cosmic rays produced at the cluster center may
be dissipated in the outer regions of the ICM, with per-
haps different spatial heat deposition from what we have
assumed. Further work on this is clearly needed.
3. RADIAL STABILITY ANALYSIS
3.1. Perturbation equations
We linearize equations (1) − (4) by using the La-
grangian perturbation method. The background ICM
is assumed to be in steady state, as described in § 2.2. A
Lagrangian perturbation, denoted by an operator ∆, is
related to an Eulerian perturbation δ in the usual way,
∆ = δ + ξ · ∇,
(21)
Here we only consider radial perturbations and
∇ · ξ =
1
r2
∂
∂r
(r2ξ),
(24)
where ξ = ∆r denotes the radial component of ξ. To
derive the perturbation equations, we use the following
properties of ∆ (Shapiro & Teukolsky 1983):
∆
d
dt
=
d
dt
∆,
∆
∂
∂r
=
∂
∂r
∆ −
∂ξ
∂r
∂
∂r
.
(25)
(26)
where ξ is the displacement vector of a fluid element
(see, e.g., Shapiro & Teukolsky 1983, pp. 130-147). By
perturbing equations (1) and (7), we find
Applying ∆ to equations (2) − (4) and using equations
(3) and (22) - (26), we obtain
∆ρ = −ρ∇ · ξ,
∆P = P
∆T
T
− P ∇ · ξ.
(22)
(23)
d2ξ
dt2 =
P
ρ
∂
∂r
(∇ · ξ) −
1
ρ
∂
∂r (cid:18)P
∆T
T (cid:19) +
1
ρ
dP
dr
∂ξ
∂r
− ξ
d2Φ
(27)
dr2 ,
κT
∂
∂r (cid:18) ∆T
T (cid:19) = F (cid:18) 7
∆T
T
−
∂ξ
∂r
+
2ξ
r (cid:19) +
∆Lr
4πr2 ,(28)
2
8
Guo et al.
resulting heating of the ICM, which is justifiable since
both the AGN duty cycle and the bubble rising time are
usually much shorter than the gas cooling time (see the
discussion above equation 13).
Taking ξ, ∆T , ∆Lr as independent variables, we seek
solutions of equations (27) − (29) that behave as ∼ eσt
with time. The term ∆Hfeed in equation (30) then sim-
plifies to
∆Hfeed =
Hσ
v0
ξ(rin),
(32)
and equations (27) − (29) may be rewritten as
(cid:18) P
ρ
− v2(cid:19) d
dr
d2Φ
(∇ · ξ) = (cid:18)rσ2 + r
−2v2 d
dr2 (cid:19) ξ
r(cid:19) +(cid:18)2σv + v
dr (cid:18) ξ
r
+
dv
dr
1
ρ
d
dr (cid:18)P
∆T
T (cid:19)
dr (cid:19) dξ
,
(33)
dr
dP
−
1
ρ
κT
d
dr (cid:18) ∆T
T (cid:19) = F (cid:20) 7
2
∆T
T
− r
d
dr (cid:18) ξ
r(cid:19) +
ξ
r(cid:21) +
∆Lr
4πr2 ,(34)
Fig.
3. -- Electron temperature (upper ) and number den-
sity (lower ) in five typical steady-state models of the cluster
Abell 2597. The data points are from the Chandra Observation
(McNamara et al. 2001). See Table 1 for additional information.
+(cid:18) P σ
γ − 1
1
4πr2
d
dr
+ ρT LT +
v
∆Lr = (P σ − ρ2Lρ − H)(∇ · ξ) − ∆H
dr(cid:19) ∆T
dr (cid:18) ∆T
γ − 1
d
dr
γ − 1
P v
γ − 1
(∇ · ξ) +
dP
dr
P
ρ
d
+P v
γv
dρ
−
T
T (cid:19) .(35)
1
4πr2
+(cid:18) P
γ − 1
∂
∂r
d
dt
∆Lr = (cid:18)P
d
dt
+ ρT LT +
− ρ2Lρ − H(cid:19) (∇ · ξ) − ∆H
dt(cid:19) ∆T
dP
dt
γ − 1
P
ρ
dρ
−
γ
,(29)
T
1
γ − 1
where LT ≡ ∂L/∂T ρ, Lρ ≡ ∂L/∂ρT , Lr = −4πr2F is
the radial heat luminosity, and ∆H is the perturbation
of AGN heating rate (equation 13)
∆H =(cid:18)−2 +
r
r0
e−r/r0
1 − e−r/r0(cid:19) ξ
H +
H
∂P/∂r
∂
∂r
∆P − H
∂ξ
∂r
+ H(cid:20)(β − 1)
∆P
P
− β
P0 (cid:21) + ∆Hfeed,
r
∆P0
where ∆P0 is the value of ∆P at the inner boundary rin,
and ∆Hfeed ≡ H∆ M (rin)/ Min is the perturbation of the
AGN heating rate due to the feedback mechanism, where
∆ M (rin) is the perturbation of the mass accretion rate
at r = rin,
∆ M (rin) =
Min
v0
∂ξ
∂t
(rin),
(31)
which can be easily derived from perturbing the defini-
tion of the mass accretion rate ( M ≡ 4πr2ρv) and using
equations (22) and (24). Note that we have neglected
any time delay between central AGN activity and the
(30)
3.2. Local stability analysis of radial modes
In equations (33) − (35) and hereinafter, we omit eσt
from all perturbation variables.
Equations (33) − (35) form an eigenvalue problem,
which can be solved numerically with appropriate bound-
ary conditions to find global eigenmodes and the eigen-
value σ. Before considering the global modes, in the fol-
lowing subsection, we first study small-scale local modes,
which may not be captured in a global stability analysis
(see the discussion in Balbus & Soker 1989). The na-
ture of local thermal stability in a stratified system can
be subtle, and linked to the convective instability of the
system (Balbus 1988).
For local stability analysis, we consider local WKB
perturbations of the form ∼ eikrr+σt. Here we neglect
the feedback mechanism of AGN heating, i.e. ∆Hfeed
in equation (30) is taken to be zero. This term simply
affects the overall normalization of heating without spa-
tial dependence, and is only important in a global anal-
ysis. The subsonic background flow is ignored as well,
so that the unperturbed steady-state ICM is nearly in
hydrostatic equilibrium. In the local approximation, we
assume that krr ≫ 1 (plane-parallel approximation) and
that the wavelengths of perturbations are much shorter
than any spatial scale on which the background quan-
tities vary (e.g., kr ≫ 1/λp, where λp ≡ (d ln P/dr)−1
is the gas pressure scale height in the ICM). To elimi-
nate high-frequency sound waves from consideration, we
also assume that σ ≪ cskr, where cs = pP/ρ is the
A Global Stability Analysis of Galaxy Clusters
9
isothermal sound speed. Therefore, the perturbed dy-
namical equation of motion (equation 33) simplifies to
(KN03):
ikrξ =
∆T
T
.
(36)
(note that in general, ∆T /T is complex).
Similarly, the perturbed energy equation (35) may be
rewritten as
(P σ − ρ2Lρ − ρL)ikrξ +(cid:18) P σ
= ∆(cid:18) 1
γ − 1
4πr2
+ ρT LT(cid:19) ∆T
T
d
dr
Lr(cid:19) + ∆H. (37)
Taking the leading order terms from the perturbations
of thermal conduction and AGN heating, we obtain
∆(cid:18) 1
4πr2
d
dr
Lr(cid:19) = −κT k2
r
∆T
T
,
(38)
∆H =
H
∂P/∂r
∆
∂P
∂r
= −Hikrξ.
(39)
In the second equality of equation (39), the perturbation
of the radial pressure gradient, ∆(∂P/∂r), is evaluated
by perturbing the Euler equation (equation (2)) and as-
suming that σ2 ≪ krc2
s/λp, which corresponds to slowly
evolving perturbations4.
Equations (36) − (39) may be combined to give
σ = σ∞ −
γ − 1
γ
H
P
−
γ − 1
γ
κT
P
k2
r ,
where
σ∞ = −
γ − 1
γ
ρT 2
P (cid:18) ∂L/T
∂T (cid:19)P
(40)
(41)
is the growth rate of local isobaric thermal instability in
the ICM without any heating (e.g., Field 1965; KN03).
In the absence of any heating source, equation (40) re-
duces to σ = σ∞, and we thus immediately recover the
generalized Field criterion for isobaric thermal instability
(Balbus 1986)
(cid:18) ∂L/T
∂T (cid:19)P
< 0.
(42)
For the ICM with L ∝ ρT 1/2, the instability criterion
given by equation (42) is easily met, suggesting that local
radial perturbations grow exponentially with the growth
time
t∞ ≡ σ−1
∞
= 0.64 Gyr(cid:16)
ne
0.05cm−3(cid:17)−1(cid:18) kBT
2keV(cid:19)1/2
.
(43)
Equation (40) confirms the well known result that
thermal conduction stabilizes short-wavelength pertur-
bations (e.g., Field 1965; Malagoli et al. 1987; KN03).
More specifically, for the ICM with L ∝ ρT 1/2, thermal
conduction alone stabilizes perturbations whose wave-
lengths (λ ≡ 2π/kr) are smaller than the critical wave-
length
λField = 2π(cid:18) 2κT
3ρL(cid:19)1/2
= 25.6 kpc(cid:18) f
0.2(cid:19)1/2
(cid:16)
ne
0.05cm−3(cid:17)−1(cid:18) kBT
2keV(cid:19)3/2
(44)
.
Equation (40) also clearly shows that an AGN heating
term as implemented in our model (H ∝ ∂P/∂r) reduces
the growth rate of local thermal instability, although it
alone could not suppress local thermal instability com-
pletely (note that (γ − 1)H/(γP σ∞) = 2H/3ρL ≤ 2/3,
since H ≤ ρL). We note that local thermal instability of
the ICM may depend on the actual mechanism by means
of which the AGN mechanical energy is transferred to the
thermal ICM; we have assumed complete local dissipa-
tion of the pdV work done by the expanding bubbles.
Alternative dissipation mechanisms (e.g., sounds waves
which damp far away, or the heating effect of dispersed
cosmic rays) could yield different results. Furthermore,
in our 1D model we have assumed isotropic heating by
bubbles; in reality the angular variation of bubble cre-
ation will affect local (and global) thermal stability.
The local stability analysis is not valid for long-
wavelength perturbations, and a successful model for the
ICM must be globally stable. KN03 studied the global
stability of the pure conduction model, and found that
it is globally unstable with the typical instability growth
time of ∼ 2 − 5 Gyr; the growth time can be significantly
shorter if one applies non-linear perturbations.
In the
next subsection, we will perform a global stability analy-
sis for our steady-state cluster models presented in § 2.2,
and show that the feedback mechanism is essential to
stabilize global thermal instability in cool core clusters.
3.3. Global unstable modes
To analyze global stability of the steady-state models
of a given cluster, we numerically solve equations (33) −
(35), which are equivalent to four first-order differential
equations for the four variables ξ/r, ∆T /T , ∆Lr and
∇ · ξ = 3ξ/r + rd(ξ/r)/dr. We solve these equations
as an eigenvalue problem, where the eigenvalue is the
growth rate σ. Since we have four variables and one
eigenvalue, we need to specify five boundary conditions.
Here we choose the same boundary conditions as KN03.
The three inner boundary conditions are
4 Provided that σ ≪ cskr, σ2 ≪ krc2
s/λp is guaranteed as long
as σ < cs/λp (i.e., the growth time of the perturbation is longer
than the sound crossing time over a pressure scale height).
ξ
r
= 1,
d
dr (cid:18) ξ
r(cid:19) = 0, ∆Lr = 0,
at r = rin. (45)
10
Guo et al.
Timescales for the clusters shown in Table 1
TABLE 2
Name
A1795
a
tcool,0
(Gyr)
0.9
b
t∞,0
(Gyr)
0.6
d Model
ǫmin
0.28
A2199
0.6
0.4
0.17
A2052
1.1
0.7
0.14
A2597
0.5
0.3
0.07
c
tgrow
(Gyr)
3.8
3.3, 43.3 e
stable
2.8
4.4
16.9
6.2
5.9
20.0
2.0
3.3
38.1
27.5
20.3
A1
A2
A3
B1
B2
B3
C1
C2
C3
D1
D2
D3
D4
D5
atcool,0 is the isobaric cooling time at the cluster center.
bt∞,0 is the growth time of the local isobaric thermal insta-
bility at the cluster center in absence of any heating source
(see equation 43).
ctgrow is the growth time of the unstable global radial mode.
dǫmin is the lower limit of the AGN feedback efficiency, above
which the ICM is effectively (tgrow > tH ) or completely stable.
eThere are two unstable global radial modes in this model.
Fig. 5. -- Eigenfunctions of the radial unstable modes for the
steady-state models of Abell 2199 presented in Fig. 2, plotted as
a function of radius. Each model has one unstable radial mode.
The first condition is a normalization condition, while the
second condition guarantees that the solutions are regu-
lar (due to the presence of a (1/r)d/dr(ξ/r) term in the
simplified form of equation 33). Since rin is not exactly
zero, the second condition need not hold strictly. We
have experimented with different values for the second
condition (for instance, d/dr(ξ/r) = ξ/r2), and find that
the results are not sensitive to it. This is consistent with
the fact that the contribution of AGN feedback (equation
32) in the perturbed energy equation (35) is independent
of this condition, although the exact value of the instabil-
ity growth time for each model will be slightly affected
due to its contribution to the perturbed cooling term
(P σ∇ · ξ) in equation (35). The last condition demands
that the perturbed radial heat luminosity is zero at the
cluster center, which is equivalent to a zero tempera-
ture gradient there. The remaining two outer boundary
conditions are set by the requirement that perturbations
vanish at the outer boundary, which has cooling times
much longer than the cluster lifetime:
ξ = 0, ∆T = 0,
at r = rout.
(46)
We use the steady-state models constructed in § 2.2
as the background states to calculate the eigenmodes of
global perturbations. Similar to KN03, we first fix σ
and set ∆T /T to an arbitrary value at r = rin. We can
then integrate eqautions (33) − (35) from r = rin to
r = rout using a Runge-Kutta method. We use the bi-
section method to update the inner value of ∆T /T and
continue iterating until the first outer boundary condi-
Fig. 4. -- Eigenfunctions of the radial unstable modes for the
steady-state models of Abell 1795 presented in Fig. 1, plotted
as a function of radius. The solid lines stand for the lone unstable
mode for the pure conduction model (model A1 in Table 1). Model
A2 has two unstable modes: tgrow = 3.3 Gyr (dotted lines) and
tgrow = 43.3 Gyr (dashed lines). Model A3 has no unstable radial
modes.
A Global Stability Analysis of Galaxy Clusters
11
tion in equation (46) is satisfied. Finally, we scan σ in
the range (104Gyr)−1 < σ < (10−4Gyr)−1, and use the
second condition in equation (46) as a discriminant for
solutions.
We first study eigenmodes with a real positive σ, which
correspond to globally unstable modes with an instability
growth time tgrow = 1/σ. The results for the steady-state
models listed in Table 1 are shown in Table 2, where, for
comparison, we also list the central values of the isobaric
cooling time (equation 20) and the growth time t∞ (equa-
tion 43) of local isobaric thermal instability in absence of
any heating source. For our fiducial cluster Abell 1795,
the pure conduction model A1 (f = 0.27) has one un-
stable mode with growth time tgrow = 3.8 Gyr, which is
consistent with KN03, who found that the equilibrium
model of A1795 (f = 0.2) has one unstable mode with
tgrow = 4.1 Gyr. For model A2, where the AGN feed-
back efficiency is ǫ = 0.1, we found two unstable modes
with tgrow = 3.3 Gyr and 43.3 Gyr respectively. Model
A3 has virtually the same background ICM profiles as
model A2 (see § 2.2), but a higher AGN feedback effi-
ciency (ǫ = 0.3). Our calculation shows that model A3
has no unstable modes. This is a remarkable result, since
we, for the first time, show from a linear analysis that
AGN feedback completely eliminates (radial) global ther-
mal instability. We note that, without introducing the
feedback mechanism for the AGN heating (i.e., if ∆Hfeed
in equation 30 is taken to be zero), model A3 is still
unstable and has two unstable modes with tgrow = 2.7
Gyr and 52.8 Gyr respectively. We repeated our calcula-
tions for many different values of model parameters, and
found that the models without the feedback mechanism
are always globally unstable, even when the same models
with the feedback mechanism included are globally sta-
ble. Thus, the feedback mechanism of AGN heating is
the key ingredient for stabilizing global thermal instabil-
ity in cool core clusters.
For the cluster Abell 2199, the pure conduction model
(B1) is globally unstable with tgrow = 2.8 Gyr, which is
consistent with KN03. With both conduction and AGN
feedback heating included, models B2 and B3 are still
globally unstable. However, the growth time of the un-
stable mode in model B3 (ǫ = 0.2) is tgrow = 16.9Gyr >
tH, which suggests that thermal instability is dynami-
cally unimportant and thus is "effectively" suppressed.
Note that, without introducing the feedback mechanism
for the AGN heating, the growth time of the unstable
mode in model B3 is much shorter (tgrow = 2.2Gyr).
In Figure 4 and 5, we plot the eigenfunctions of the
unstable modes in galaxy clusters A1795 and A2199, re-
spectively. Clearly, the perturbations have largest am-
plitude in the cluster central regions (r . 10 kpc) and
decay rapidly with increasing radius, which suggests that
perturbations reach nonlinear amplitudes much faster in
central regions than in outer regions. Such global modes
have also been found by KN03 for equilibrium ICM mod-
els with thermal conduction only.
As shown in Table 2, we obtain similar results for the
other two cool core clusters listed in Table 1. It is worth
mentioning that the pure conduction model for the clus-
ter Abell 2597 requires an implausibly high value of the
thermal conductivity (f = 1.3) and is globally unstable.
By including AGN feedback heating, model D3 requires
a smaller conductivity (f = 0.4) and is effectively stable
Fig. 6. -- Effect of the AGN feedback efficiency on thermal
stability in typical cool core clusters. For different models of each
cluster, the values of f and ǫ M are roughly the same (equal to those
in the second model listed in Table 1; see the text for details). Top
M with ǫ. Bottom panel: the dependence of the
panel: scaling of
growth time of unstable modes on ǫ for each cluster. Note that
the clusters are stable or effectively stable when ǫ is greater than
a lower limit ǫmin. The line for A1795 is double-valued since it
has two unstable modes; both these modes vanish when ǫ > 0.28.
Here, we assume ǫ is a constant; if, as suggested by observations,
ǫ ∝ M ν (with ν ∼ 0.3 − 0.6), then ǫmin will be smaller (see text).
(tgrow = 38.1Gyr, for ǫ = 0.1).
3.4. Dependence on the AGN feedback efficiency
From the results of the previous subsection, one might
conjecture that the ICM is stable (or effectively stable)
if the AGN feedback efficiency ǫ is greater than a lower
limit. To fully explore the dependence of the cluster
stability on the AGN feedback efficiency, we consider
the steady-state cluster models subject to the following
boundary conditions
ne(rin) = n0,
T (rin) = Tin,
r2
inF (rin) = 0,
T (rout) = Tout, ne(rout) = nout,
(47)
where the value of nout for each cluster is chosen to be
that of the second model for that cluster, as listed in Ta-
ble 1. As described in § 2.2, such boundary conditions
12
Guo et al.
The lower panel of Figure 6 clearly shows that the clus-
ter is stable or effectively stable (tgrow > tH ) when ǫ is
greater than a lower limit ǫmin. We note that this result
generalizes as well to models with different values of f , as
seen in Figure 7, which shows the effect of AGN feedback
efficiency on global stability for models of A2597 with
f = 0.3, 0.4 and 0.5. Assuming that the real intracluster
gas is in a stable quasi-steady state, our global stability
analysis thus suggests a constraint on the kinetic effi-
ciency of AGN feedback. As listed in Table 2, the values
of ǫmin for these four typical cool core clusters are ǫmin ∼
0.07 − 0.28, which is roughly consistent with the recent
estimate of ǫ ∼ 0.3 for radio-loud AGNs by Heinz et al.
(2007) and is marginally consistent with observational
estimates of ǫ ∼ 0.01 − 0.1 by Allen et al. (2006) and
Merloni & Heinz (2007).
In our AGN feedback model,
we assume that ǫ is a constant. However, recent X-ray
observations seem to suggest that ǫ ∝ M ν, where ν ∼ 0.3
(Allen et al. 2006) or 0.6 (Merloni & Heinz 2007). In this
case, the AGN feedback will be even stronger (i.e., in
equation 30, ∆Hfeed = H∆ M (rin)/ Min + H∆ǫ/ǫ), and
M appropriate for
thus the value of ǫmin (evaluated at
the steady-state case) will be reduced. For the cluster
A1795, ǫmin is reduced from 0.28 to 0.22 (ν = 0.3) or
0.18 (ν = 0.6). For the cluster A2199, ǫmin is reduced
from 0.17 to 0.13 (ν = 0.3) or 0.10 (ν = 0.6). The ac-
tual value of ǫmin is also sensitive to the exact form of
the AGN heating law adopted, particularly its spatial de-
pendence. Here we have only considered the pdV work
due by rising bubbles, and ignored, for instance, cos-
mic ray heating (Guo & Oh 2008), viscous dissipation of
sound waves (Ruszkowski et al. 2004) or shock heating
(Bruggen et al. 2007). However, while we do not place
great store in the absolute value we obtain for ǫmin, the
fact that there is a minimal heating efficiency ǫmin for a
given temperature profile should be fairly robust, since
it only requires that Lagn ∝ ǫ M .
3.5. Dependence on the background profiles
In this subsection, we will study the dependence of
the cluster stability on the background steady-state ICM
profiles. We adopt the clusters Abell 1795 and Abell 2199
as our fiducial clusters. Since the gas cooling time at the
outer boundary is much longer than the Hubble time,
we consider steady-state cluster models with fixed val-
ues of Tout and nout, which are chosen to be the same as
those in the second model listed in Table 1 for each clus-
ter. Of the three inner boundary conditions in equation
(17), we choose two (including r2
inF (rin) = 0 and a given
value of Tin for a specific model). We have four bound-
ary conditions for three first-order ordinary differential
M can be solved
equations (§ 2.2). Thus, the value of
as the eigenvalue for a specific steady-state model with
a given value of f and ǫ (which are deemed to be fixed
by the physics of thermal conduction and black hole ac-
cretion respectively). The corresponding central electron
number density for each steady-state model (represented
by the varying value of Tin) is shown in the upper panels
of Figure 8 and 10. In addition, it is possible to vary the
inner boundary rin. Varying rin at fixed Tin is basically
denegerate with varying Tin at fixed rin, and thus we do
not explore this additional dependence.
We first consider the models of the cluster A1795 with
Fig. 7. -- Effect of the AGN feedback efficiency on thermal
stability in A2597. Each curve is plotted as a function of ǫ with
a fixed conductivity suppression factor f . Top panel: scaling of
M with ǫ. Bottom panel: the dependence of the growth time of
unstable modes on ǫ. For different levels of fixed conductivity, the
cluster always becomes more stable as the AGN feedback efficiency
increases.
ensure that the steady-state ICM profiles are almost ex-
actly the same for models with different values of ǫ (see
the lines for models A2 and A3 in Figure 1); it is only
meaningful to study the dependence of the cluster sta-
bility on ǫ when the background profiles are virtually
the same. Since the steady state cluster model only has
three variables P (r), T (r), r2F (r) (§ 2.2), the five bound-
ary conditions (47) can determine two eigenvalues. Al-
though our hybrid cluster models formally have three pa-
rameters (f , ǫ and M ), they are essentially determined
by two parameters: f , which determines the level of ther-
mal conduction, and ǫ M , which determines the level of
AGN heating, while the subsonic inflow itself ( M ) only
has a negligible effect. Thus, for each cluster, the values
of f and ǫ M are roughly fixed by the boundary condi-
tions (47), and M varies with the free parameter ǫ (see
Figure 6).
For each steady-state cluster model, we then repeat
our stability calculations and search for unstable modes.
The results are shown in Figure 6, where the steady-
state mass accretion rate and the growth time of unstable
modes are plotted as a function of the AGN feedback ef-
ficiency. The value of f for each cluster roughly equals to
that in the second model of that cluster listed in Table 1.
A Global Stability Analysis of Galaxy Clusters
13
Fig. 8. -- Dependence of the cluster stability on the background
profile for the cluster A1795. For a fixed value of f and ǫ, the cor-
responding central electron number density (upper panel ), steady-
state mass accretion rate (middle panel ), and the growth time of
unstable modes (lower panel ) are plotted as a function of the cen-
tral gas temperature Tin. The dot short-dashed line in the lower
panel shows the growth time of the unstable mode in steady-state
models with f = 0.12, ǫ = 0.1 and no feedback mechanism for
AGN heating (i.e., ∆Hfeed in equation 30 is taken to be zero),
while the dot long-dashed line stands for the central gas cooling
time in models with f = 0.12, ǫ = 0.1.
f = 0.12 and ǫ = 0.1 (i.e., variations of model A2). The
dependence of the cluster stability on the background
profile (represented by Tin) is plotted in Figure 8 (solid
line), which clearly shows that the model with either a
relatively flat (Tin > 4.5 keV) or steep (Tin < 1.7 keV)
temperature profile is (effectively) stable, while ther-
mal instability can develop at intermediate temperatures.
Figure 9 shows radial profiles of electron number density,
temperature and the eigenfunction (ξ/r) of the radial un-
stable mode for three typical steady state models with
Tin = 6 keV, 3 keV, and 1 keV respectively.
3.5.1. Cool core versus non-cool core clusters
X-ray observations also suggest that clusters can be
subdivided into two distinct categories according to the
presence or absence of a cool core (e.g., Peres et al. 1998,
Bauer et al. 2005, Sanderson et al. 2006, Chen et al.
2007).
In this section we discuss how our model may
account for this effect.
The dot-dashed line in the lower panel of Figure 8
shows the growth time of the unstable mode in steady-
state models with f = 0.12, ǫ = 0.1 and no feedback
Fig. 9. -- Radial profiles of electron number density (upper
panel ), temperature (middle panel ) and the eigenfunction of the
radial unstable mode (lower panel ) in three typical steady-state
models of the cluster Abell 1795 with f = 0.12 and ǫ = 0.1.
mechanism for AGN heating (i.e., ∆Hfeed in equation 30
is taken to be zero). Without the feedback mechanism,
the instability growth time in cool core clusters with rela-
tively steep temperature profiles is very short (∼ 2 Gyr),
suggesting that the feedback mechanism plays a key role
in stabilizing thermal instability in these clusters. As
the central gas temperature increases and the central
gas density decreases, the stabilizing effect of the feed-
back mechanism becomes smaller, which is reasonable
since the perturbation of the central mass accretion rate
(note that Lagn ∝ Min) scales as the central gas density
(equation 31) and the importance of the feedback mech-
anism in the energy perturbation equation (35) may be
expressed as ∆Hfeed/(P σ∇ · ξ), which (using equations
(13, 14, 32) and dP/dr ∼ ρg) scales as T −2 at any spe-
cific radius. Colder, denser (i.e.
low entropy) gas has a
steeper pressure gradient in a fixed potential well, which
increases the volumetric rate at which cavities perform
pdV work as they rise.
On the other hand, although the stabilizing effect of
the feedback mechanism becomes negligible for non-cool
core clusters with relatively flat temperature profiles, our
stability analysis surprisingly shows that these cluster
models are also (effectively) stable. As shown in §3.2,
the diffusive AGN heating (H ∝ ∂P/∂r) only increases
14
Guo et al.
Fig. 11. -- The growth time of unstable modes for models of the
cluster A2597 with f = 0.4 and ǫ = 0.05, plotted as a function
of the central gas temperature. Both the stable CC and NCC
branches are seen in the calculations with either our simplified
cooling function or a full cooling function (see text).
duction f < 1, or else the CC cluster is globally unstable
on fairly short timescales. We did the same calculations
for the cluster A2052 and A2597, and found similar re-
sults.
We stress that, except for models with very high val-
ues of f and ǫ (which may always be stable, as seen
from the trend of different lines shown in Figure 8 and
10), the clusters are usually stable either when the cen-
tral temperature is high (non-CC) or when the central
temperature is sufficiently low. Figure 8 (and the anal-
ogous Figure 10 for A2199) suggest that the interme-
diate values correspond to globally unstable solutions.
This is consistent with 1D hydrodynamic simulations of
RB02 and Guo & Oh (2008):
if conductivity is large,
the cluster relaxes to stable NCC states (see Figure 1
and 2 of Guo & Oh 2008); if f is small (i.e., conduction
couldn't offset the cooling), the cluster first cools grad-
ually through NCC states, and then quickly relaxes to
stable CC states (see Fig. 1 in RB02 or Fig. 4 and 5
of Guo & Oh 2008). We note that the radiative cooling
times in non-CC clusters are generaly longer than in the
CC ones and in some of the non-CC clusters no heating
is required to prevent thermal instability on a timescale
shorter than the Hubble time (note however that in the
cases shown in Figures 8 and 10 (dot long-dashed lines),
the central cooling timescale is always shorter then the
instability growth time). Irrespectively of whether this
is the case or not, our model naturally explains the di-
chotomy between CC clusters where AGN feedback sig-
natures are observed (and where the AGN play the key
role in establishing global stability) and the non-CC clus-
ters, where conduction may play a role but where the
AGN feedback is not readily observed. This trend for
the non-CC (flat temperature) clusters to have no AGN
as opposed to CC clusters with clear AGN observational
signatures is clearly seen in the recent data compiled by
Dunn & Fabian (2008). This is also consistent with the
observation by Rafferty et al. (2008) who show that the
short central cooling time corresponds to younger AGN
(i.e., shorter X-ray cavity ages).
Metal line cooling may become important when the
gas temperature is low. We checked our calculations
with a full cooling function (equation 35 of Guo & Oh
2008, which is based on Sutherland & Dopita 1993), and
did not find qualitative changes to our results. As an
example, Figure 11 shows the dependence of the cluster
stability on the central gas temperature for the cluster
Fig. 10. -- Dependence of the cluster stability on the back-
ground profile for the cluster A2199. For a fixed value of f and ǫ,
the corresponding central electron number density (upper panel ),
steady-state mass accretion rate (middle panel ), and the growth
time of unstable modes (lower panel ) are plotted as a function of
the central gas temperature Tin. The dotted line in the lower panel
stands for the Hubble time, while the dot short-dashed line shows
the growth time of the unstable mode in steady-state models with
f = 0.36, ǫ = 0.2 and no feedback mechanism for AGN heating,
while the dot long-dashed line stands for the central gas cooling
time in models with f = 0.36, ǫ = 0.2.
the growth time of local thermal instability, while ther-
mal conduction may completely suppress local perturba-
tions with wavelength less than λField, which increases
as the central gas temperature increases and which may
then be greater than the cooling radius. Thus, we ex-
pect that thermal conduction alone may completely sup-
press thermal instability in these non-CC clusters. This
is confirmed by our stability analysis: we constructed
NCC models with only conduction and found them to
be stable.
The short and long dashed lines in Figure 8 show the
models with a higher thermal conductivity and AGN
feedback efficiency respectively. Obviously, a higher level
of thermal conduction increases the stability of the clus-
ter. However, a higher level of AGN feedback efficiency
only increases the stability of CC cluster models, but
has a negligible stabilizing effect on non-CC cluster mod-
els. Thus, AGN feedback heating is not required in NCC
clusters. On the other hand, for CC clusters with lower
central temperatures, AGN heating is generally required:
otherwise one is either unable to build an equilibrium
profile with a physically plausible level of thermal con-
A Global Stability Analysis of Galaxy Clusters
15
A2597, which has the lowest Tin in our cluster sample.
With both free-free and metal line cooling included, the
gas cooling rate increases and the dependence of tgrow
on Tin becomes similar to that of the higher-temperature
cluster A1795 (see Fig. 8), but both the stable CC and
NCC branches still exist.
Through hydrodynamic numerical simulations, RB02
shows that the ICM heated by a combination of AGN
feedback and thermal conduction usually relaxes to a sta-
ble quasi-steady state. Surprisingly, our stability analysis
shows that a specific steady-state model may be globally
unstable if the AGN feedback efficiency is lower than a
critical value. This "inconsistency" may be explained if
the cluster with lower ǫ relaxes to a steady state with
lower central gas temperatures, which may then be effec-
tively stable, as clearly shown in this subsection. Recent
numerical simulations by Guo & Oh (2008) indeed con-
firms that the cluster central regions in their cosmic-ray
feedback models with lower ǫ cool to higher densities and
lower temperatures in the final steady state (see Figure
7 of Guo & Oh (2008)).
3.6. Global decaying modes
We have also searched for global decaying modes with
real and negative σ. For each steady-state model of the
ICM, similar to KN03 (see Figure 4b of KN03), we found
a series of decaying modes, within which there exists a
slowest decaying mode with the smallest decay rate. Fig-
ure 12 shows the eigenfunction (ξ/r) of the slowest decay-
ing mode in typical steady-state models for the cluster
Abell 1795 and Abell 2199. The decay rate of the most
slowly decaying mode could potentially serve as an indi-
cator of the "attractor" solution toward which a cluster
evolves after it has been reset by a merger, and might
be worthy of further study.
It may also be interesting
to explicitly study the effect of AGN feedback on global
non-radial modes and overstable modes, which we leave
to future work (see Malagoli et al. 1987, Balbus & Soker
1989 and KN03 for local analyses of these modes in the
ICM with thermal conduction).
4. SUMMARY AND DISCUSSION
Recent Chandra and XMM-Newton X-ray observations
suggest that the hot intracluster gas is maintained by one
or more heating sources at keV temperatures for a pe-
riod at least comparable to the lifetime of galaxy clusters
(§ 1). Since the emission lines expected from cooler gas
are notoriously absent (e.g., Peterson & Fabian (2006)),
the heating source (or sources) should also effectively
suppress the thermal instability of the ICM. Although
thermal conduction stabilizes thermal instability for per-
turbations with short wavelengths, the equilibrium ICM
heated by thermal conduction alone is globally unstable
and will evolve to a cooling catastrophe under global per-
turbations. On the other hand, using numerical hydro-
dynamic simulations, RB02 and Guo & Oh (2008) show
that the ICM, which is initially in a state far from equi-
librium and is heated by a combination of thermal con-
duction and AGN feedback, usually relaxes to a quasi-
equilibrium steady state. Although these simulations
adopt simplified 1D models for the elusive spatial distri-
bution of AGN heating, they strongly suggest that the
AGN feedback mechanism plays a key role in suppressing
global thermal instability.
Fig. 12. -- Eigenfunctions of the slowest decaying modes (σ < 0)
in typical steady-state models for the cluster: (a) Abell 1795 and
(b) Abell 2199. For each steady-state model, there exist decaying
modes with larger decay rates, which are not shown in this figure.
In this paper, we perform a detailed formal analysis of
thermal instability in the ICM with both AGN feedback
heating and thermal conduction. To build initial un-
perturbed states for stability analysis, we first construct
steady-state cluster models, where the gas density and
temperature profiles fit observations quite well and where
the mass accretion rates are highly suppressed compared
to those predicted by the standard cooling flow mod-
els (§ 2.2). Using the Lagrangian perturbation method,
we then derive a set of differential equations that form
an eigenvalue problem for global radial modes with the
mode growth rate σ as the eigenvalue. For pure con-
duction models of typical cool-core clusters, we find that
the ICM has one unstable radial mode with the typical
growth time ∼ 2 − 6 Gyr (Table 2), which is consistent
with the results of Kim & Narayan (2003). However, for
the hybrid models with both AGN heating and thermal
conduction, global thermal instability is effectively re-
duced or even completely suppressed if the feedback effi-
ciency ǫ is greater than a lower limit ǫmin. Interestingly,
if the AGN heating (equation 13) is independent of the
16
Guo et al.
central mass accretion rate (i.e., no feedback mechanism
for AGN heating), the ICM is still unstable (§ 3.3), which
suggests that the feedback mechanism is essential to sup-
press thermal instability.
Assuming that the real intracluster gas is in a sta-
ble quasi-steady state, our global stability analysis thus
suggests a minimum value ǫmin of the kinetic efficiency
of AGN feedback,
if it is to suppress a cooling flow.
The value of ǫmin required in typical cool-core clusters is
around ∼ 0.07 − 0.28 (see Table 2), which is roughly con-
sistent with the estimate of the jet production efficiency
(∼ 30%) for radio-loud AGNs by Heinz et al. (2007) and
which is marginally consistent with recent observational
estimations of ǫ ∼ 0.01 − 0.1 by Allen et al. (2006) and
Merloni & Heinz (2007). Note that the value of ǫmin will
be reduced if ǫ is an increasing function of the central
mass accretion rate as suggested by recent observations
(§ 3.4). Although the existence of ǫmin should be fairly
robust, its exact value will also depend on the assumed
form of AGN heating law.
A related important issue in AGN feedback models is
how and what fraction of the cooling gas at a distance of
order 1 kpc gets to the cluster center and finally fuels the
AGN. This could be a very complex process due to the
large range of distance scales involved. Our calculation
extended from ∼ 1Mpc in the cluster outer regions to
∼ 1kpc at our innermost integration point. However, the
black hole gravitational radius of influence rBH is much
further in:
GMBH
σ2 = 0.05 kpc(cid:18) MBH
109 M⊙(cid:19)(cid:16)
σ
300 km s−1(cid:17)−2
rBH =
(48)
It is conceivable that if the flow is steady all the way
down to the black hole,
it transitions from a cool-
ing flow to an adiabatic Bondi flow (see, for instance,
Quataert & Narayan 2000 and the discussion in § 5 of
Chandran & Rasera 2007), and there is tentative ob-
servational evidence that the Bondi formula may pro-
vide a reasonable approximation of the accretion process
in X-ray luminous galaxies (Allen et al. 2006). How-
ever, there are a host of potential complications: an-
gular momentum, which would cause a torus or accre-
tion disk to form instead, magnetic pressure and outflows
(Proga & Begelman 2003), and local thermal instability
resulting in the formation of stars and cold gas blobs
(Pizzolato & Soker 2005), resulting in "cold accretion" or
the possibility that the black hole is fed by stellar winds.
In this paper, in line with most numerical simulations
in the literature (e.g., Ruszkowski & Begelman 2002;
Brighenti & Mathews
2003; Hoeft & Bruggen 2004;
Vernaleo & Reynolds 2006) we have assumed that essen-
tially all of the inflowing gas in our innermost bound-
ary point eventually makes it to the black hole, over
timescales long compared to the AGN duty cycle, but
shorter than or comparable to the gas inflow timescale,
tflow ∼ 109yr. This behaviour is likely to be intermit-
tent rather than steady: gas accumulates in an accretion
disk/torus around the black hole, which eventually trig-
gers an outburst and leads to gas consumption, etc. We
assume that in steady state, the black hole will even-
tually consume all the gas which is supplied to it.
In
particular, in our model, while the black hole can exert
thermal feedback on the cooling gas (through its heat-
ing activity), it exerts negligible hydrodynamic feedback
(by consuming gas faster or more slowly than the supply
rate, thus affecting pressure forces throughout the cool-
ing flow region). If indeed such a scenario applies, the
results of our present paper should be recalculated in de-
tail, since in that case M cannot be freely varied, but is
further constrained by the accretion law. Given the large
uncertainties and difficulty of this calculation, we leave
this to future work.
In § 3.5, we study the dependence of the cluster stabil-
ity on the background steady-state profiles and find that
the unstable cluster models with ǫ < ǫmin may become
effectively stable if the central gas temperature drops to
much lower values. Numerical simulations by Guo & Oh
(2008) indeed confirm that the cluster central regions in
their cosmic-ray feedback models with lower ǫ usually
cool to higher densities and lower temperatures in the fi-
nal steady state. Thus, unlike the pure conduction mod-
els, where nonlinear evolution of global unstable modes
lead to the cooling catastrophe, the ICM in our hybrid
cluster models with AGN feedback included may always
evolve to a quasi-steady state, which is effectively stable.
On the other hand, we also show that thermal conduction
will completely suppress thermal instability in non-CC
clusters with relatively flat temperature profiles. Thus,
the stability of the ICM favors two distinct categories
of cluster steady state profiles: CC clusters stabilized
mainly by AGN feedback and non-CC clusters stabilized
by thermal conduction. Interestingly, recent X-ray ob-
servations also suggest that clusters can be subdivided
into two distinct categories according to the presence or
absence of a cool core (e.g., Peres et al. 1998, Bauer et
al. 2005, Sanderson et al. 2006, Chen et al. 2007).
It is perhaps even possible that these two categories
of clusters represent different stages of the same object.
The importance of thermal conduction on global scales
obviously depends on the large scale structure of the
cluster magnetic fields. Recent calculations suggest that
thermal conduction of heat into the cluster core can be
self-limiting:
in cases where the temperature decreases
in the direction of gravity, a buoyancy stability sets in
which re-orients a radial magnetic field to be largely
transverse, shutting off conduction to the cluster cen-
ter (Quataert 2008). Non-linear simulations indicate the
heat flux could be reduce to ∼ 1% of the Spitzer value
(Parrish & Quataert 2008). Thus, the following scenario
could arise: as conductivity falls, gas cooling and mass
inflow will increase, triggering AGN activity. The rising
buoyant bubbles or the gas convective overturn mediated
by cosmic rays (Chandran 2004) may re-orient the mag-
netic field to be largely radial again, increasing thermal
conduction and reducing mass inflow, shutting off the
AGN until the heat flux driven buoyancy instability sets
in once again. The cluster could therefore continuously
cycle between cool-core (AGN heating dominated) and
non cool-core (conduction dominated) states.
Global linear stability analysis can clearly serve as a
useful complement to simulations, both for the physi-
cal insight they can deliver and speed in exploring pa-
rameter space. Formally, a globally stable equilibrium
state is a necessary but insufficient condition for a suc-
cessful cluster model. A linear analysis fails for large
non-linear perturbations (as a cluster would undergo, for
A Global Stability Analysis of Galaxy Clusters
17
instance, during mergers). Nonetheless, the linear sta-
bility analysis qualitatively reproduces the same features
we have observed in 1D hydrodynamic simulations when
we start the simulation from arbitrary initial conditions:
conduction-only models suffer drastic cooling flows, while
conduction + AGN heating models relax to a stable state
with low mass inflow rates. For a given value of ǫ and
f , a wide range of stable profiles are accessible. From a
suite of linear stability analyses alone, it is not possible to
predict the detailed final temperature and density profile
the cluster relaxes to (which is somewhat sensitive to ini-
tial conditions), but it is possible to make general state-
ments about stability and whether the cluster will relax
to a cool core or non cool-core configuration. Finally,
we have restricted our attention to a 1D radial analy-
sis. While a 2D or 3D analysis would be more general, it
would be much more complicated and unlikely to prove
enlightening, at least for global modes. Non-radial modes
may prove important in the case of local thermal insta-
bility, but the fastest growing unstable global mode -- the
possible cooling flow we wish to stem -- is likely to respect
spherical symmetry, unless the heating sources (such as
the AGN jets) are severely anisotropic. Such models are
beyond the scope of the present paper.
We thank Omer Blaes, Ian Parrish, Mitch Begelman,
Marcus Bruggen, Paul Nulsen, Biman Nath for discus-
sions. We also thank Mark Voit and Noam Soker for
comments on the manuscript, and the anonymous referee
for a very detailed and helpful report. FG and SPO ac-
knowledge support by NASA grant NNG06GH95G. MR
acknowledges support by Chandra theory grant TM8-
9011X.
REFERENCES
Allen S. W., Dunn R. J. H., Fabian A. C., Taylor G. B., Reynolds
C. S., 2006, MNRAS, 372, 21
Balbus S. A., 1986, ApJ, 303, L79
Balbus S. A., Soker N., 1989, ApJ, 341, 611
Bauer, F. E., Fabian, A. C., Sanders, J. S., Allen, S. W., &
Johnstone, R. M. 2005, MNRAS, 359, 1481
Begelman M. C., 2001, in Hibbard J. E., Rupen M., van Gorkom
J. H., eds, ASP Conf. Ser. 240, Gas and Galaxy Evolution (San
Francisco: ASP), 363
Bertschinger E., Meiksin A., 1986, ApJ, 306, L1
Bırzan L., Rafferty D. A., McNamara B. R., Wise M. W., Nulsen
P. E. J., 2004, ApJ, 607, 800
Blanton E. L., Sarazin C. L., McNamara B. R., Wise M. W., 2001,
ApJ, 558, L15
Bregman, J. N., & David, L. P. 1988, ApJ, 326, 639
Brighenti F., Mathews W. G., 2003, ApJ, 587, 580
Bruggen M., Heinz S., Roediger E., Ruszkowski M., Simionescu A.,
2007, MNRAS, 380, L67
Bruggen M., Kaiser C. R., 2002, Nature, 418, 301
Burns J. O., 1990, AJ, 99, 14
Burns, J. O., Hallman, E. J., Gantner, B., Motl, P. M., & Norman,
M. L. 2008, ApJ, 675, 1125
Chandran, B. D. G. 2004, ApJ, 616, 169
Chandran, B. D. G., & Rasera, Y. 2007, ApJ, 671, 1413
Chen, Y., Reiprich, T. H., Bohringer, H., Ikebe, Y., & Zhang, Y.-Y.
2007, A&A, 466, 805
Cho J., Lazarian A., Honein A., Knaepen B., Kassinos S., Moin P.,
2003, ApJ, 589, L77
Churazov E., Sunyaev R., Forman W., Bohringer H., 2002,
MNRAS, 332, 729
Ciotti L., Ostriker J. P., 2001, ApJ, 551, 131
Dunn, R. J. H., & Fabian, A. C. 2008, MNRAS, 385, 757
Ettori S., Fabian A. C., Allen S. W., Johnstone R. M., 2002,
MNRAS, 331, 635
Fabian A. C., 1994, ARA&A, 32, 277
Fabian A. C., Sanders J. S., Ettori S., Taylor G. B., Allen S. W.,
Crawford C. S., Iwasawa K., Johnstone R. M., Ogle P. M., 2000,
MNRAS, 318, L65
Field G. B., 1965, ApJ, 142, 531
Forman W., Jones C., Churazov E., Markevitch M., Nulsen P.,
Vikhlinin A., Begelman M., Bohringer H., Eilek J., Heinz S.,
Kraft R., Owen F., Pahre M., 2007, ApJ, 665, 1057
Guo F., Oh S. P., 2008, MNRAS, 384, 251
Heinz S., Merloni A., Schwab J., 2007, ApJ, 658, L9
Hoeft, M., & Bruggen, M. 2004, ApJ, 617, 896
Johnstone R. M., Allen S. W., Fabian A. C., Sanders J. S., 2002,
MNRAS, 336, 299
Kaiser C. R., Binney J., 2003, MNRAS, 338, 837
Kim W.-T., Narayan R., 2003, ApJ, 596, 889
Lazarian A., 2006, ApJ, 645, L25
McCarthy, I. G., Babul, A., Bower, R. G., & Balogh, M. L. 2008,
MNRAS, 403
Malagoli A., Rosner R., Bodo G., 1987, ApJ, 319, 632
McNamara B. R., Nulsen P. E. J., 2007, ARA&A, 45, 117
McNamara B. R., Wise M. W., Nulsen P. E. J., David L. P., Carilli
C. L., Sarazin C. L., O'Dea C. P., Houck J., Donahue M., Baum
S., Voit M., O'Connell R. W., Koekemoer A., 2001, ApJ, 562,
L149
Merloni A., Heinz S., 2007, MNRAS, 381, 589
Narayan R., Medvedev M. V., 2001, ApJ, 562, L129
Navarro J. F., Frenk C. S., White S. D. M., 1997, ApJ, 490, 493
Parrish I. J., Quataert E., 2008, ApJ, 677, L9
Peres, C. B., Fabian, A. C., Edge, A. C., Allen, S. W., Johnstone,
R. M., & White, D. A. 1998, MNRAS, 298, 416
Peterson J. R., Fabian A. C., 2006, Phys. Rep., 427, 1
Peterson J. R., Kahn S. M., Paerels F. B. S., Kaastra J. S., Tamura
T., Bleeker J. A. M., Ferrigno C., Jernigan J. G., 2003, ApJ, 590,
207
Peterson J. R., Paerels F. B. S., Kaastra J. S., Arnaud M., Reiprich
T. H., Fabian A. C., Mushotzky R. F., Jernigan J. G., Sakelliou
I., 2001, A&A, 365, L104
Piffaretti, R., & Kaastra, J. S. 2006, A&A, 453, 423
Pizzolato, F., & Soker, N. 2005, ApJ, 632, 821
Proga, D., & Begelman, M. C. 2003, ApJ, 592, 767
Quataert, E., & Narayan, R. 2000, ApJ, 528, 236
Quataert E., 2008, ApJ, 673, 758
Rafferty, D., McNamara, B., & Nulsen, P. 2008, ArXiv e-prints,
802, arXiv:0802.1864
Reynolds C. S., Begelman M. C., 1997, ApJ, 487, L135
Rosner, R., & Tucker, W. H. 1989, ApJ, 338, 761
Roychowdhury S., Ruszkowski M., Nath B. B., 2005, ApJ, 634, 90
Ruszkowski M., Begelman M. C., 2002, ApJ, 581, 223
Ruszkowski M., Bruggen M., Begelman M. C., 2004, ApJ, 611, 158
Rybicki G. B., Lightman A. P., 1979, Radiative processes in
astrophysics. New York: Wiley
Sanderson A. J. R., Ponman T. J., O'Sullivan E., 2006, MNRAS,
372, 1496
Shapiro S. L., Teukolsky S. A., 1983, Black holes, white dwarfs,
and neutron stars. New York, Wiley-Interscience
Sijacki, D., Pfrommer, C., Springel, V., & Ensslin, T. A. 2008,
ArXiv e-prints, 801, arXiv:0801.3285
Soker, N. 2003, MNRAS, 342, 463
Spitzer L., 1962, Physics of Fully Ionized Gases. New York:
Interscience
Sutherland R. S., Dopita M. A., 1993, ApJS, 88, 253
Tamura T., Kaastra J. S., Peterson J. R., Paerels F. B. S., Mittaz
J. P. D., Trudolyubov S. P., Stewart G., Fabian A. C., Mushotzky
R. F., Lumb D. H., Ikebe Y., 2001, A&A, 365, L87
Vernaleo, J. C., & Reynolds, C. S. 2006, ApJ, 645, 83
Voigt L. M., Fabian A. C., 2004, MNRAS, 347, 1130
Voit G. M., Donahue M., 2005, ApJ, 634, 955
Zakamska N. L., Narayan R., 2003, ApJ, 582, 162
|
0710.0796 | 1 | 0710 | 2007-10-03T14:21:52 | Keck spectroscopy of the faint dwarf elliptical galaxy population in the Perseus Cluster core: mixed stellar populations and a flat luminosity function | [
"astro-ph"
] | We present the result of a photometric and Keck-LRIS spectroscopic study of dwarf galaxies in the core of the Perseus Cluster, down to a magnitude of M_B = -12.5. Spectra were obtained for twenty-three dwarf-galaxy candidates, from which we measure radial velocities and stellar population characteristics from absorption line indices. From radial velocities obtained using these spectra we confirm twelve systems as cluster members, with the remaining eleven as non-members. Using these newly confirmed cluster members, we are able to extend the confirmed colour-magnitude relation for the Perseus Cluster down to M_B = -12.5. We confirm an increase in the scatter about the colour magnitude relationship below M_B = -15.5, but reject the hypothesis that very red dwarfs are cluster members. We measure the faint-end slope of the luminosity function between M_B = -18 and M_B = -12.5, finding alpha = -1.26 \pm 0.06, which is similar to that of the field. This implies that an overabundance of dwarf galaxies does not exist in the core of the Perseus Cluster. By comparing metal and Balmer absorption line indices with alpha-enhanced single stellar population models, we derive ages and metallicities for these newly confirmed cluster members. We find two distinct dwarf elliptical populations: an old, metal poor population with ages ~ 8 Gyr and metallicities [Fe/H] < -0.33, and a young, metal rich population with ages < 5 Gyr and metallicities [Fe/H] > -0.33. Dwarf galaxies in the Perseus Cluster are therefore not a simple homogeneous population, but rather exhibit a range in age and metallicity. | astro-ph | astro-ph | Mon. Not. R. Astron. Soc. 000, 000 -- 000 (0000)
Printed 6 June 2018
(MN LATEX style file v2.2)
Keck spectroscopy of the faint dwarf elliptical galaxy population in
the Perseus Cluster core: mixed stellar populations and a flat
luminosity function
Samantha J. Penny1 and Christopher J. Conselice1
1School of Physics & Astronomy, University of Nottingham, Nottingham, NG7 2RD, United Kingdom
6 June 2018
ABSTRACT
We present the result of a photometric and Keck-LRIS spectroscopic study of dwarf galaxies
in the core of the Perseus Cluster, down to a magnitude of MB = −12.5. Spectra were ob-
tained for twenty-three dwarf-galaxy candidates, from which we measure radial velocities and
stellar population characteristics from absorption line indices. From radial velocities obtained
using these spectra we confirm twelve systems as cluster members, with the remaining eleven
as non-members. Using these newly confirmed cluster members, we are able to extend the
confirmed colour-magnitude relation for the Perseus Cluster down to MB = −12.5. We con-
firm an increase in the scatter about the colour magnitude relationship below MB = −15.5,
but reject the hypothesis that very red dwarfs are cluster members. We measure the faint-end
slope of the luminosity function between MB = −18 and MB = −12.5, finding α = −1.26 ±
0.06, which is similar to that of the field. This implies that an overabundance of dwarf galaxies
does not exist in the core of the Perseus Cluster. By comparing metal and Balmer absorption
line indices with α-enhanced single stellar population models, we derive ages and metallicities
for these newly confirmed cluster members. We find two distinct dwarf elliptical populations:
an old, metal poor population with ages ∼ 8 Gyr and metallicities [Fe/H] < −0.33, and a
young, metal rich population with ages < 5 Gyr and metallicities [Fe/H] > −0.33. Dwarf
galaxies in the Perseus Cluster are therefore not a simple homogeneous population, but rather
exhibit a range in age and metallicity.
Key words: galaxies: dwarf -- galaxies: clusters: general -- galaxies: clusters: individual:
Perseus Cluster -- galaxies: luminosity function, mass function
7
0
0
2
t
c
O
3
]
h
p
-
o
r
t
s
a
[
1
v
6
9
7
0
.
0
1
7
0
:
v
i
X
r
a
1 INTRODUCTION
Faint, low mass galaxies are the most numerous galaxy type in the
universe, and are thus fundamental in understanding galaxy forma-
tion. However, their low luminosities and low surface brightness
make detailed studies of them difficult. Because dwarfs are so com-
mon, any galaxy evolution/formation theory must clearly be able to
predict and describe the properties of these galaxies.
The luminosity functions of nearby galaxies in all environ-
ments reveal that dwarf galaxies are far more common than brighter
galaxies. Within galaxy clusters there furthermore appears to be an
overdensity of dwarf galaxies when compared to the field. The ori-
gin of these extra dwarfs, or if this excess is even real, remains un-
known. The cluster luminosity function itself is not universal, but
is strongly dependent on environment (Sabatini et al. 2003), with
the luminosity function often steeper in the more diffuse outer re-
gions of clusters than in the denser cluster cores. The luminosity
function also depends on the individual cluster, with each one hav-
ing a characteristic luminosity function. For example, the Fornax
Cluster is compact and rich in early-type galaxies, with a flat lumi-
nosity function faint-end slope of α = −1.1 (Mieske et al. 2007).
The Virgo Cluster has a high abundance of spiral galaxies and has
a steeper luminosity function with a faint end slope of α = −1.6
(Trentham & Hodgkin 2002; Sabatini et al. 2003).
We know that both Local Group (LG) dwarf elliptical and
dwarf spheroidal galaxies display varying star formation histories,
with metal poor populations as old as the classical halo globular
galaxies, but with evidence for star formation as little as 2-3 Gyr
ago. Some low-mass LG galaxies such as Sagittarius, contain sur-
prisingly high metal rich stellar populations considering their lu-
minosities, as is also seen in clusters of galaxies. Previous spectro-
scopic and ground based imaging has revealed that dwarfs in the
core of clusters are not a simple homogeneous population, with
cluster core galaxies fainter than MB = −15 containing a mix
of metallicities and ages (Poggianti et al. 2001; Rakos et al. 2001;
Conselice, Gallagher & Wyse 2003).
The formation scenarios for dwarf galaxies fall into two cat-
egories. The first of these is that dwarfs are old, primordial ob-
jects (hierarchical model) (e.g. White & Frenk 1991). The second
is that they have recently evolved or transformed from a progenitor
galaxy population (e.g. Moore, Lake & Katz 1998). Within hierar-
2
S. J. Penny and C. J. Conselice
chical models of galaxy formation, dwarf galaxies are formed from
small density fluctuations in the early universe. The lowest mass
systems (i.e. dwarf galaxies) form first, and massive galaxies are
built from these in mergers. If star formation occurred soon after the
gravitational collapse of initial density perturbations, dwarf galaxy
halos would be amongst the first objects to form (e.g. Dekel & Silk
1986). Theory also predicts that such systems would form first in
the densest environments, although cluster dEs would not neces-
sarily contain the first stars formed in the universe. These halos
could have formed their stars quite recently, if star formation was
suppressed in some way (Tully et al. 2002).
Dwarfs in clusters, on the other hand, could also be the rem-
nants of stripped discs or dwarf irregulars. Through galaxy ha-
rassment and interactions, these progenitors become morpholog-
ically transformed into dEs. Cluster galaxies become stripped of
their interstellar medium, and become dynamically heated by high-
speed interactions with other galaxies and the gravitational po-
tential of the cluster. To compensate, the galaxy loses stars, and
over time the spiral can morphologically transform into a dE
(Conselice, Gallagher & Wyse 2001a). This model is supported by
recent observations of embedded discs in dEs in the Virgo and
Fornax clusters (Barazza, Binggeli & Jergen 2002; De Rijcke et al.
2003). Distant clusters at z ≈ 0.4 are also filled with small spiral
galaxies, but this population is largely absent in nearby clusters,
where dwarf spheroidals make up the faint-end of the luminosity
function (Moore et al. 1998).
Downsizing (De Lucia at al. 2004) is the observed trend that
star formation occurs later, and over more extended timescales,
in smaller galaxies. In this scenario, dEs and lower mass galax-
ies formed or entered clusters after the giant galaxies. Low mass
galaxies on average end their star formation after the giant ellipti-
cal galaxies. For example, the faint end of the red sequence in clus-
ters is not formed until z < 0.8 (De Lucia at al. 2004), implying the
luminosity weighted stellar populations of lower mass galaxies are
younger than the stars in the giant elliptical galaxies. This seems
to contradict the hierarchical method of galaxy formation, where
dwarf systems form first. Also, dwarf elliptical galaxies are prefer-
entially found in dense cluster environments, with few examples of
isolated dEs, again contradicting the hierarchical model of galaxy
formation. However, some dEs in clusters contain old stellar pop-
ulations, so it is likely that multiple formation methods are needed
to explain the origin of cluster dwarfs.
The colour-magnitude relation (CMR) for cluster galaxies is
well defined at bright magnitudes and forms a tight sequence. How-
ever, it is unclear what the shape of the red sequence is at fainter
magnitudes (Andreon et al. 2006). Conselice, Gallagher & Wyse
(2002) find that galaxies at the brightest 4 magnitudes of Perseus
obey the CMR, with the fainter candidate members showing sig-
nificant scatter from the relation. In contrast, Andreon et al. (2006)
and others find no deviation from the CMR at fainter magnitudes
in other clusters. Instead, the faint red sequence is an extrapolation
of what is observed at brighter magnitudes. With spectroscopy, we
can test this directly by finding confirmed dwarf members of the
cluster to establish the true form of the colour magnitude relation.
Spectroscopy can also be used to measure the faint-end of the
luminosity function in galaxy clusters, which is fundamental in de-
scribing the galaxy population. The luminosity function contains
important information on the formation and evolution of galaxies,
and at low redshifts contains the combined influence of the galaxy
initial mass function, and the effects of any evolutionary processes
that have taken place in the cluster since its formation.
cluster membership for dwarfs at MB < −16, and present the
ages and metallicities of these galaxies based on the strengths of
Balmer and metal absorption lines in their spectra. We confirm that
twelve dwarfs are Perseus Cluster members, while the remaining
objects are background galaxies. Using these results, we examine
the colour-magnitude relation for the Perseus Cluster down to MB
= −12.5. We also measure the luminosity function faint-end slope
α based on our cluster membership results. Our main conclusion
is that some dwarf galaxies in Perseus have old, metal poor pop-
ulations, whilst others are younger, metal rich systems. This sug-
gests that dwarf galaxies in the core of the Perseus Cluster are not
a simple, homogeneous population, but require multiple scenarios
for their formation.
This paper is organised as follows: in § 2 we discuss the ob-
servations and the selection criteria for the dwarf galaxies, § 3.1
identifies the cluster members, § 3.2 presents the colour magnitude
relation, and the central luminosity function is presented in § 3.3.
In § 3.4 we derive luminosity weighted ages and metallicities for
the newly confirmed cluster members. A discussion of the main re-
sults is presented in § 4 and these results are summarised in § 5. We
assume the distance to the Perseus Cluster is 77 Mpc throughout
this paper.
2 OBSERVATIONS
The Perseus Cluster (Abell 426) is one of the richest nearby galaxy
clusters, with a redshift v = 5366 km s−1 (Struble & Rood 1999),
and at a distance D = 77 Mpc. It also has a high velocity dispersion
of σ = 1324 km s−1 (Struble & Rood 1999). Despite its proximity
it has not been studied in as much detail as clusters such as Virgo
and Fornax, most likely due to its low Galactic latitude (b = −13◦).
Candidate dwarf elliptical galaxies in Perseus for our Keck
spectroscopic observations were selected from those listed in
Conselice et al. (2003), and we follow the same numbering system
throughout this paper. All dwarfs in this list have colours (B−R)0
< 2, and are symmetric, round or elliptical in shape, without ev-
idence for star formation. All the candidates also have a central
surface brightness fainter than µB = 24.0 mag arcsec−2, and have
near-exponential surface brightness profiles. We restrict our study
to those dwarf galaxies with MB < −12.5 to be more certain of
cluster membership. For more detail of the selection process, see
Conselice et al. (2003) and Conselice et al. (2002).
We also observed an additional two objects not in this orig-
inal list, one of which is a dwarf galaxy, and the other an irreg-
ular galaxy (selected to fill the slits on one of the masks during
spectroscopy). We also carried out photometry on these additional
objects so they could be included on the colour-magnitude diagram
presented in § 3.2. All candidates lie within the central 173 arcmin2
of the cluster, and have MB = −16 to MB = −12.5.
The imaging data from which candidate dwarfs are acquired
was taken using the WIYN 3.5m f/6.2 telescope located at the Kitt
Peak National Observatory. Harris B and R images were acquired
on the nights of 1998 November 14 and 15 under photometric con-
ditions, with an average seeing of 0.7". Exposure times were 800s
in the R band and 1000s in B. Flat fields were taken prior to each
night of observing, and combined to create a single flat. Landolt
standard star fields were obtained throughout each night, allowing
us to find accurate zero points, air masses and colour terms. See
Conselice et al. (2002) and Conselice et al. (2003) for details of the
data reduction.
In this paper, we use deep Keck spectroscopy to determine
The new spectroscopic observations we present in this paper
The dwarf galaxy population in the Perseus Cluster
3
Table 1. Candidate dwarf galaxies in the Perseus Cluster core
Galaxy
(J2000.0)
(J2000.0)
α
δ
B
(mag)
(B−R)
(mag)
vradial
(km s−1)
Cluster Member
CGW 1*
CGW 2*
CGW 3
CGW 4*
CGW 5
CGW 6
CGW 7
CGW 8
CGW 9
CGW 10
CGW 11*
CGW 12*
CGW 13*
CGW 14
CGW 15*
CGW 16
CGW 17
CGW 18
CGW 19
CGW 20
CGW 21
CGW 22*
CGW 23*
CGW 24
CGW 25
CGW 26*
CGW 27
CGW 28
CGW 29
CGW 30
CGW 31
CGW 32*
CGW 33
CGW 34*
CGW 35
CGW 36*
CGW 37
CGW 38
CGW 39
CGW 40
CGW 41
CGW 42
CGW 43*
CGW 44
CGW 45
CGW 46
CGW 47
CGW 48
CGW 49
CGW 50*
CGW 51
CGW 52
CGW 53
54
Irregular
03 18 53.5
03 18 53.8
03 18 54.9
03 18 55.7
03 18 55.8
03 18 57.8
03 18 59.5
03 18 59.7
03 18 59.8
03 19 00.4
03 19 03.5
03 19 04.2
03 19 04.7
03 19 05.2
03 19 05.2
03 19 06.0
03 19 06.8
03 19 09.1
03 19 09.5
03 19 10.4
03 19 12.7
03 19 13.0
03 19 14.9
03 19 15.1
03 19 15.9
03 19 17.3
03 19 17.6
03 19 18.0
03 19 18.8
03 19 20.1
03 19 21.2
03 19 22.1
03 19 22.4
03 19 24.6
03 19 24.7
03 19 24.7
03 19 26.9
03 19 27.1
03 19 31.4
03 19 31.7
03 19 33.5
03 19 36.1
03 19 38.6
03 19 39.7
03 19 41.7
03 19 42.3
03 19 43.5
03 19 45.7
03 19 46.1
03 19 55.4
03 19 56.1
03 19 56.1
03 19 58.5
03 19 48.6
03 19 05.1
+41 31 50.1
+41 32 52.3
+41 25 55.3
+41 25 54.4
+41 23 41.3
+41 24 57.6
+41 31 18.9
+41 26 40.0
+41 30 37.1
+41 29 02.4
+41 35 13.6
+41 35 20.6
+41 32 24.6
+41 25 07.0
+41 34 48.1
+41 26 18.7
+41 26 40.8
+41 32 41.7
+41 31 30.9
+41 29 37.0
+41 30 37.7
+41 34 51.8
+41 30 27.1
+41 28 56.6
+41 30 20.3
+41 34 54.5
+41 29 57.7
+41 24 25.3
+41 26 32.4
+41 24 06.6
+41 28 42.5
+41 24 27.2
+41 24 04.2
+41 25 53.4
+41 24 36.3
+41 25 52.6
+41 33 05.1
+41 27 16.1
+41 26 28.7
+41 31 21.3
+41 24 39.4
+41 32 47.4
+41 33 46.1
+41 26 39.1
+41 29 17.0
+41 34 16.6
+41 28 52.9
+41 32 18.0
+41 24 51.2
+41 25 04.4
+41 29 09.4
+41 32 38.7
+41 31 02.0
+41 33 28.6
+41 28 12.4
20.54 ± 0.02
20.01 ± 0.01
21.61 ± 0.04
20.76 ± 0.03
21.46 ± 0.03
20.06 ± 0.01
21.39 ± 0.04
19.71 ± 0.01
21.34 ± 0.02
20.26 ± 0.02
20.21 ± 0.02
20.95 ± 0.02
21.50 ± 0.02
21.87 ± 0.04
20.37 ± 0.02
19.78 ± 0.02
20.86 ± 0.05
21.57 ± 0.02
20.59 ± 0.01
19.27 ± 0.01
21.61 ± 0.04
20.10 ± 0.01
21.77 ± 0.06
21.40 ± 0.05
21.89 ± 0.05
21.97 ± 0.04
21.40 ± 0.05
21.62 ± 0.03
20.94 ± 0.03
21.34 ± 0.04
20.77 ± 0.03
20.15 ± 0.01
21.75 ± 0.05
21.75 ± 0.05
21.90 ± 0.07
21.80 ± 0.03
20.74 ± 0.03
18.48 ± 0.01
18.72 ± 0.01
19.30 ± 0.01
21.76 ± 0.08
21.82 ± 0.03
21.92 ± 0.03
21.05 ± 0.02
18.72 ± 0.01
20.67 ± 0.02
20.87 ± 0.02
21.56 ± 0.05
21.43 ± 0.03
21.06 ± 0.02
20.71 ± 0.01
21.31 ± 0.02
21.86 ± 0.03
19.75 ± 0.03
15.19 ± 0.01
0.91 ± 0.05
1.14 ± 0.05
1.80 ± 0.05
1.44 ± 0.04
1.55 ± 0.05
1.22 ± 0.04
0.89 ± 0.07
1.32 ± 0.04
0.75 ± 0.05
1.29 ± 0.05
1.09 ± 0.05
1.15 ± 0.04
1.95 ± 0.05
1.58 ± 0.06
1.09 ± 0.05
1.26 ± 0.05
1.70 ± 0.06
1.01 ± 0.04
0.91 ± 0.04
1.27 ± 0.04
1.00 ± 0.07
1.72 ± 0.04
0.85 ± 0.11
1.12 ± 0.04
1.11 ± 0.08
1.63 ± 0.06
1.64 ± 0.05
1.25 ± 0.05
1.02 ± 0.06
1.57 ± 0.06
1.40 ± 0.06
1.16 ± 0.04
1.90 ± 0.07
1.29 ± 0.06
1.65 ± 0.08
1.48 ± 0.05
1.43 ± 0.06
1.43 ± 0.04
1.39 ± 0.04
1.24 ± 0.04
1.55 ± 0.10
1.01 ± 0.05
1.03 ± 0.05
1.21 ± 0.05
1.42 ± 0.04
1.04 ± 0.05
1.24 ± 0.05
0.44 ± 0.11
1.41 ± 0.05
1.87 ± 0.05
1.42 ± 0.04
1.87 ± 0.05
1.00 ± 0.06
1.30 ± 0.04
1.41 ± 0.05
. . .
. . .
. . .
. . .
80585 ± 182
. . .
. . .
. . .
. . .
5845 ± 47
. . .
. . .
. . .
. . .
. . .
6577 ± 45
. . .
46906 ± 210
67527 ± 114
7325 ± 48
8741 ± 31
. . .
. . .
64780 ± 146
70320 ± 59
. . .
125580 ± 51
. . .
5110 ± 77
. . .
8662 ± 14
. . .
. . .
. . .
. . .
. . .
70960 ± 23
4151 ± 59
6421 ± 50
. . .
. . .
51225 ± 148
. . .
. . .
3120 ± 27
. . .
2886 ± 48
. . .
8457 ± 38
. . .
69132 ± 82
40994 ± 105
64495 ± 79
3363 ± 38
4027 ± 39
. . .
. . .
. . .
. . .
N
. . .
. . .
. . .
. . .
Y
. . .
. . .
. . .
. . .
. . .
Y
. . .
N
N
Y
Y
. . .
. . .
N
N
. . .
N
. . .
Y
. . .
Y
. . .
. . .
. . .
. . .
. . .
N
Y
Y
. . .
. . .
N
. . .
. . .
Y
. . .
Y
. . .
Y
. . .
N
N
N
Y
Y
* Not targeted for spectroscopy
Units of right ascension are hours, minutes and seconds, and units of declination are degrees, arcminutes and arcseconds. Magni-
tudes and colours are taken from Conselice et al. (2003). 54 and the irregular galaxies are additional objects to the original list in
Conselice et al. (2003). Galaxies flagged with an * were not targeted in our spectroscopy, but are included in the colour-magnitude
relation and luminosity function for the Perseus Cluster.
4
S. J. Penny and C. J. Conselice
Table 2. Potential background cluster members
α
δ
Galaxy
(J2000.0)
(J2000.0)
CGW 19
CGW 24
CGW 25
CGW 37
CGW 51
CGW 53
03 19 09.5
03 19 15.1
03 19 15.9
03 19 26.9
03 19 56.1
03 19 58.5
+41 31 30.9
+41 28 56.6
+41 30 20.3
+41 33 05.1
+41 29 09.4
+41 31 02.0
z
0.22
0.23
0.23
0.24
0.23
0.22
were taken using the Keck Low Resolution Imaging Spectrometer
(LRIS) (Oke et al. 1995) on the Keck I Telescope on 2002 Novem-
ber 26 and 27. Using four LRIS masks, we targeted forty-one dwarf
galaxies, and one irregular, and potentially interacting, galaxy in
the Perseus Cluster.
Successful spectra were obtained for twenty-three dwarf
galaxy candidates from Conselice et al. (2003), and the potentially
interacting galaxy. The spectra were taken with an average seeing
of 1.1" under clear, but non-photometric, conditions. We obtained
spectroscopy with both the red and blue sides of LRIS. On the blue
side we used a 400 line mm−1 grism blazed at 3400 A, whilst
on the red side we used a 600 line mm−1 grating blazed at 5000
A, producing effective resolutions of 8 and 6 A, respectively. The
spectra were bias subtracted, flat-fielded and rectified using IRAF1
reduction techniques for multi-slit data. This combined the various
frames into a single one, from which the spectra for the individual
dwarf galaxies could be extracted separately.
The positions and photometric properties of all our science ob-
jects are listed in Table 1. The radial velocities in Table 1 are gener-
ally measured on the red spectra, where the wavelength calibration
was more reliable, and include heliocentric corrections. The errors
on the radial velocities are based on uncertainties in the wavelength
calibration, and determined by measuring the wavelengths of sky-
lines in the spectra and comparing these to expected values. The
difference in the actual and expected positions of the skylines were
then converted into a velocity error.
We also obtained Hubble Space Telescope (HST ) Advanced
Camera for Surveys (ACS) imaging in the F555W and F814W
bands (Penny et al. 2008, in preparation), from which we obtain
high resolution images of several of our cluster dwarfs and non-
members.
3 RESULTS
3.1 Cluster membership
Using our spectra we identify absorption features such as the Mg
b and Fe5270 lines, as well as Balmer lines such as Hβ. From
these lines we derive the radial velocities of each of the dwarfs
through comparing the measured absorption lines with those of
known early-type galaxy spectra by identifying well known lines
by eye.
We find that twelve of the dwarf candidates are at approxi-
mately the radial velocity of the Perseus Cluster, v ≈ 5366 km s−1
1 IRAF is distributed by the National Optical Astronomy Observatories
which are operated by the Association of Universities for Research in
Astronomy, Inc., under cooperative agreement with the National Science
Foundation
(Struble & Rood 1999), with the remaining eleven galaxies with ra-
dial velocities placing them outside the cluster. All cluster members
are within 2.5σ of the mean value of the radial velocity obtained by
Struble & Rood (1999). Several of the non-cluster members show
strong emission lines in their spectra. Interestingly, a number of the
background galaxies are at the same redshift (z ≈ 0.23) and could
be members of a background cluster. The positions and redshifts of
these galaxies are listed in Table 2.
The mean radial velocity for the dwarfs is 5750 ± 43 km
s−1, slightly higher than the value of ≈ 5366 km s−1 found by
Struble & Rood (1999). Struble & Rood (1999) found a velocity
dispersion σ = 1324 km s−1 for galaxies in the Perseus Cluster,
and we re-measure this value using just our newly confirmed dwarf
galaxy members. We find a velocity dispersion for the dwarfs of σ
= 1818 km s−1, again higher than that found by Struble & Rood
(1999) for bright galaxies in the Perseus Cluster. The Perseus Clus-
ter however has a complex X-ray morphology (Fabian et al. 1981),
and is also not a simple, relaxed system.
Conselice et al. (2001a) find a similar result for the dwarf el-
liptical population of the Virgo Cluster, where the velocity disper-
sion of the ellipticals is 462 km s−1, whilst the velocity dispersion
of the dwarf elliptical population is 726 km s−1. The fact that the
velocity dispersion of dwarf ellipticals is higher than the overall
cluster is potentially a sign that the dwarfs may originate from in-
falling galaxies. Dwarf ellipticals could originate from spiral galax-
ies that are stripped of mass by high-speed interactions. These
dwarfs would retain the velocity characteristics of the original in-
falling spirals, having velocity dispersions higher than the cluster
ellipticals, reflecting the mass distribution of the cluster when the
galaxies were accreted. Dwarfs that lie in the central parts of clus-
ters are also susceptible to cannibalism and destruction. Therefore
dwarfs would be found preferentially in the outer parts of clusters
at high velocity dispersion as these are the most likely dwarfs to
survive in the cluster environment. Also, we cannot rule out mass
segregation occurring (Conselice et al. 2001a).
3.1.1 HST ACS imaging
Deep HST ACS imaging reveals detail not seen in the ground
based WIYN imaging (Figures 1 and 2). From such images, galax-
ies such as background spirals are more easily identified. All the
confirmed cluster members are elliptical in shape, with no evidence
for internal sub-structure (Fig. 1). It is clear from Fig. 2 that three
of the non-members (CGW 19, 27 and 37) are spiral galaxies, al-
though this detail is not seen in the original WIYN imaging pre-
sented in Conselice et al. (2003), and as a result were incorrectly
classified as potential cluster members.
3.2 The colour-magnitude relation
There exists a well-defined colour-magnitude relationship amongst
the non-dwarf early-type galaxy population in galaxy clusters, such
that brighter galaxies are redder. This correlation is generally re-
garded as a relationship between a galaxy's mass (traced by magni-
tude) and metallicity (traced by colour). Using our newly confirmed
cluster dwarfs, we investigate whether this relationship extends to
the dwarf galaxy population of the Perseus Cluster. Our photome-
try is taken from Conselice et al. (2003), as well as including some
additional photometry on objects not included in Conselice et al.
(2003).
Our colour-magnitude diagram is shown in Figure 3, with the
The dwarf galaxy population in the Perseus Cluster
5
Figure 1. Montage of HST ACS images of the confirmed Perseus Cluster dwarfs (galaxies 10, 20, 21, 31, 45, and 47). The number at the upper right of each
image is MB, the number on the bottom right is the (B−R)0 colour, and the number on the left is the galaxy number from Conselice et al. (2003).
Figure 2. Montage of HST ACS images of the non-members (galaxies 18, 19, 24, 25, 27, 37, 42, and 51). The number at the upper right of each image is
MB (assuming the galaxy is at the distance of Perseus), the number on the bottom right is the (B−R)0 colour, and the number on the left is the galaxy number
from Conselice et al. (2003).
6
S. J. Penny and C. J. Conselice
solid line a least-squares fit found by Conselice et al. (2003) to the
bright cluster ellipticals, and the dashed line an extension of this fit
to fainter magnitudes. We carry out a least squares fit to this data
using only confirmed members to establish a relation between the
colour and magnitude for the confirmed population of Perseus as
a whole. This is shown as the dotted line on the colour-magnitude
diagram. This fit is slightly steeper than the relationship found by
Conselice et al. (2002) for the brighter cluster galaxies in Perseus.
The scatter in the colour magnitude relation for the cluster
members is small (σ = 0.05) down to MB = −15.5, but increases
to σ = 0.11 at magnitudes fainter than this. The degree of colour
scatter for the cluster member fainter than MB = −15.5 is real,
based on a comparison to the photometric errors, to a 2σ confi-
dence level. The confirmed faint cluster members that depart from
the colour-magnitude relation (CMR), tend to be a bluer than the
CMR. These bluer dwarfs are most likely metal poor, as spectro-
scopic results from previous studies have shown (Held & Mould
1994, Lotz et al. 2004). The redder colours of two of the dwarf
galaxies implies that they may have gone through a process of metal
enrichment (Conselice et al. 2001a). These dwarfs are also likely
part of a dwarf galaxy population that have a different origin to
the blue dwarfs. Conselice et al. (2003) suggest that the most likely
origin for these red dwarfs is from infalling spirals. The red clus-
ter dwarfs are most likely not primordial objects, and would have
intermediate or young stellar populations.
From the colour magnitude diagram, it appears that galaxies
with (B−R)0 > 1.6 (i.e. the reddest elliptical) are not cluster mem-
bers. We find no confirmed cluster members with (B−R)0 > 1.6,
although there are non-cluster members with (B−R)0 less than this,
so cluster membership cannot be determined by colour and mor-
phology alone for more distant clusters such as Perseus. Radial ve-
locities obtained via spectroscopy are required as non-cluster mem-
bers can have colours and structures similar to cluster members, as
can be seen on the colour-magnitude diagram in Fig. 3. However,
morphology can still be a useful tool in determining cluster mem-
bership in more distant clusters from high resolution imaging, as
our HST observations have shown.
Broadband (B−R)0 colours also reveal information about the
nature of the stellar populations in dwarf galaxies, and scatter from
the colour-magnitude relationship can be partly explained by the
amount of metal enrichment that has occurred. A broad metal-
licity distribution has been inferred for dwarfs in other galaxy
populations such as Fornax and Coma (e.g. Poggianti et al. 2001;
Rakos & Schombert 2004), where a breakdown in the colour-
magnitude relation is seen below M5550 = −17. Age effects are
also likely important in the colours of low-mass dwarf galaxies,
with chaotic star formation histories in such galaxies. We inves-
tigate this further in § 3.4 using absorption line indices to derive
luminosity weighted ages and metallicities for our dwarfs.
3.3 The central luminosity function
One of the main reasons for fitting luminosity functions is to de-
termine the relative number of faint systems compared to brighter
galaxies. This is typically done by measuring the value of the
faint-end slope α, with values typically in the range ≈ −1
and −2.3 in clusters (Conselice et al. 2002; Mieske et al. 2007;
Beijersbergen et al. 2002).
Previous to this study, values were found for the faint-end
slope α = −1.56 ± 0.07 and −1.44 ± 0.04 (De Propris & Pritchett
1998; Conselice et al. 2002) in the central regions of the Perseus
Cluster. We refit this luminosity function after removing all galax-
Figure 3. Colour magnitude diagram for dwarf galaxies in the Perseus Clus-
ter, taken from Conselice et al. (2003). We include all objects with MB <
−12.5. All the dwarfs listed in Table 1 are included in this plot. The solid
line represents the fit between colour and magnitude for the objects with
MB < −16, with the dashed line the extension of the fit to fainter magni-
tudes. The dotted line is the colour-magnitude relation for the cluster ellip-
ticals and dwarf galaxies.
ies now classified as non-cluster members, and those with (B−R)0
> 1.6 (the colour of the reddest elliptical) from the luminosity func-
tion in Conselice et al. (2002). Galaxies with (B−R)0 > 1.6 are not
included in the luminosity function as they are most likely not clus-
ter members.
Figure 4 shows our new luminosity function to a limiting mag-
nitude of MB − 12.5. We fit this luminosity function with a power
law, of the form dN/dL = Lα. Galaxies are binned according to
their magnitude, and a fit was made to the number counts using a
least-squares. The resulting fit is shown in Figure 4 as a solid line.
We find the faint end slope to be α = −1.26 ± 0.06. This value is
flatter than that found previously by Conselice et al. (2002).
Importantly, our value of α also matches that of the field, with
Trentham, Sampson & Banerji (2005) finding an average logarith-
mic fit of α = −1.26 ± 0.11 for the field luminosity function. This
result implies that there is no difference between the relative den-
sity of dwarf galaxies to giant galaxies in the field and in the centre
of the Perseus Cluster.
It is worth mentioning that the determination of the faint end
slope is strongly dependent on the technique used to derive the lu-
minosity function, in particular how the cluster members are de-
termined. The selection process for cluster members as outlined in
Conselice et al. (2002) along with the spectroscopic technique for
selecting cluster members in this study has led to us removing all
dwarfs with (B−R)0 > 1.6. Other methods used to determine clus-
ter membership include selecting dwarfs according to their mor-
phology and surface brightness profiles (e.g. Trentham & Hodgkin
2002). Another method is to perform a background subtraction,
with the luminosity function determined from a measurement of
the excess of galaxies relative to the field Phillips et al. (1998a).
The dwarf galaxy population in the Perseus Cluster
7
ibrated ones by <1%, and this does not contribute significantly to
our errors (Strader et al. 2003). The spectra were then smoothed us-
ing a wavelength-dependent Gaussian kernel to match the Lick/IDS
resolution of ∼8-10 A (Worthey & Ottaviani 1997).
In our analysis, we measured the the Balmer line indices HαA,
HαF , Hβ, and the metal line indices Mg b, Fe λ5270 and Fe
λ5335 using the passband definitions in Trager et al. (1998) and
Nelan et al. (2005). We tested our method on model spectra pro-
vided online by Guy Worthey, and our line measurements matched
the expected ones to < 5%. We were unable to obtain Lick stan-
dards during our observations, so we are not able to provide an off-
set between our observations and the Lick System. However, when
Brodie et al. (2005) measured the Lick indices on six standard stars
under an identical LRIS set-up, no significant deviations from the
Lick system were found. The offsets found are typically smaller
than the index errors that they measured, so were not applied to the
indices. Therefore, it was not necessary to apply an offset between
our observations and the Lick System.
The strengths of metal lines can be influenced by the ages of
the stellar populations, as well as by their metallicity. This com-
plicates the process of constraining the metallicities of the dwarf
galaxies. However, when a Balmer index is plotted against a metal
line, it yields a two-dimensional theoretical grid from which the
luminosity weighted ages and metallicities can be estimated. Sev-
eral models exist that predict the Lick indices for SSPs at various
metallicities and ages, with the models used throughout this pa-
per being those of Thomas, Maraston & Bender (2003) and Smith
(2005). The Thomas et al. (2003) models cover lines in the wave-
length range 4000 A . λ . 6500 A, which is not sufficient to cover
the Hα line. We use the Smith (2005) models to compare with this
index.
3.4.1 Ages and metallicities
Before deriving the ages and metallicities for our Perseus dwarfs,
we estimated the [α/Fe] values for our galaxies. The α elements
are created rapidly in Type II supernovae, while iron is produced
by Type Ia supernovae on longer timescales. By examining the lo-
cation of our dwarf galaxies on a plot of <Fe> versus Mg b, we
can derive likely values of [α/Fe] for our cluster dwarfs (Fig 5).
Mg b is an indicator of the α elements, and <Fe> = 1
2 (Fe5270 +
Fe5335).
We find that two of our dwarfs are consistent with super-
solar abundance ratios ([α/Fe] ≈ +0.3). Super-solar abundances
are consistent with older stellar populations, such as globular clus-
ters and elliptical galaxies. Super-solar [α/Fe] indicate that enrich-
ment from Type II SNe has taken place. The Thomas et al. (2003)
models converge at low metallicities, and our errors are not small
enough to reject lower [α/Fe] ratios. Therefore, we use a value of
[α/Fe] = 0.3 when determining the ages and metallicities for our
dwarfs. Other studies have found [α/Fe] ratios as high as this in a
dwarf population (e.g. Conselice 2006).
To measure the ages of our Perseus dwarfs, we first compare
our measured Hα indices to the Smith (2005) single stellar popu-
lation models of HαA and HαF , as a function of age and metal-
licity (Fig 6), with metallicities [Fe/H] = −0.38, 0.00, +0.32, and
+0.56, although we exclude the [Fe/H] = +0.32 model for simplic-
ity in Figure 6. These models assume solar-scaled chemical abun-
dance ratios, whereas we have found our dwarfs to have super-solar
abundance ratios. However, this model is thought to be more robust
against variation in [α/Fe] than the higher order Balmer lines. The
Hα line has a weak response to metallicity variations, so it should
Figure 4. The luminosity function for galaxies with MB > −18 and
(B−R)0 < 1.6 in the central region of Perseus. The solid line is the least-
squares fit to the data.
This method can lead to a very steep faint end slope, due to con-
tamination from background galaxies.
Moore et al. (1998) explain the dwarf-density relation as the
result of galaxy harassment. Harassment operates more effectively
in denser regions, such as cluster cores. Tidal effects become in-
creasingly important in the inner regions of clusters, with smaller
galaxies more easily destroyed in cluster centres. This could ex-
plain why we find a faint-end slope α equal to that of the field,
as the lowest mass galaxies are simply not able to survive in the
innermost regions of galaxy clusters.
3.4 Ages & metallicities from Lick indices
Measurements of the ages and metallicities of the stellar popula-
tions in our dwarfs are made using the spectra of the newly con-
firmed members. The strength of the Balmer absorption lines in
a stellar population reflect the luminosity-weighted effective tem-
perature of the system, which is dominated by the main-sequence
turnoff (e.g. Smith 2005). In a single stellar population (SSP)
model, the tun-off luminosity and temperature depend on the age
of the stellar population.
To measure the equivalent widths if our absorption lines we
follow the procedure outlined in Brodie & Huchra (1990) who
measured the indices of globular clusters with low-resolution
spectra. We measure Lick indices from our spectra to determine
ages and metallicities for our dwarfs (Worthey & Ottaviani 1997;
Trager et al. 1998; Nelan et al. 2005), although it was not possible
to measure every absorption feature for each galaxy. These indices
are calculated with respect to pseudo-continua, defined on either
side of the feature bandpass. We can be sure that there is no star
formation occurring in our dwarfs based on deep Hα narrow-band
imaging of these galaxies (Conselice, Gallagher & Wyse 2001b).
The line-strength indices are computed by first shifting the
galaxy spectra to their rest-frame wavelengths. We did not flux cal-
ibrate our data, but flux calibrated indices differ from non-flux cal-
8
S. J. Penny and C. J. Conselice
Figure 5. Relationship between the <Fe> index and the Mg b index for the
cluster dwarfs. Model predictions by Thomas et al. (2003) for the relation-
ship between <Fe> and Mg b are plotted for the abundance ratios [α/Fe]
= 0.0, +0.3 and +0.5. The dotted lines represent the model predictions for
the ages 1, 5, 8, 11 and 15 Gyr, with age increasing from left to right. The
dashed lines represent the model metallicities of -2.25, -1.35, -0.33, 0.0 and
+0.35, with metallicity increasing from bottom to top.
Figure 6. Comparison between the HαA indices for our dwarf galaxies and
the [Fe/H] = 0.00 model from Smith (2005) (dashed line) of how HαA
evolves as a function of time. The models for [Fe/H] = −0.38, +0.56 are
also plotted as solid lines. The points are the measured values of Hαa that
we obtained from our dwarf spectra plotted at the ages at which these values
are expected in the [Fe/H] = 0.00 model
be a good indicator of stellar ages in integrated spectra. Several
ages are seen within the dwarfs, with some having ages > 5 Gyr,
and the remaining a younger population.
By combining a metal index with a Balmer index, we can
place constraints on the luminosity-weighted ages and metallicities
of the stellar populations of the newly confirmed Perseus dwarfs.
We use the Mg b index for this purpose, as well as the <Fe> in-
dex.
Figure 7 shows that at least one of the Perseus dwarfs is lo-
cated in the low-metallicity, and old age part of the Hβ-Mg b di-
agram, with ages ∼ 8 ± 3 Gyr and metallicities [Fe/H] < -0.33
at a 3σ confidence level. There is also a younger, more metal rich
dwarf with age ∼ 3 ± 2 Gyr and [Fe/H] ∼ 0.35. Comparing the
Hβ indices to the <Fe> indices produces similar results, with one
dwarf galaxy having an age of 11 ± 3 Gyr, and with a metallicity
less than [Fe/H] = −0.33. Within the error bars, another dwarf is
potentially older than 5 Gyr and again has [Fe/H] < −1.35. One
dwarf is clearly much younger (< 5 ± 3 Gyr) and more metal rich
([Fe/H] > −0.33. The remaining two galaxies lie below the grid so
it is not possible to make estimates for their ages and metallicities,
but they are most likely older (> 8 Gyr) and metal poor systems.
The differences in age and metallicity measured throughout this
paper are real to 2-3σ.
We also compare the Hβ and the [α/Fe]-insensitive in-
dex [MgFe]' = [(Mgb)(0.72 × Fe5270 + 0.28 × Fe5335)]0.5
(Thomas et al. 2003). This index servers as the best tracer of the
overall metallicity, Hβ is also less sensitive than the other Balmer
line indices to metallicity. Again, we find the same results with two
old, metal poor dwarfs ([Fe/H] < −1.35, age > 8 Gyr), and one
metal rich, young dwarf ([Fe/H] > −0.33, age < 5 Gyr).
We find several Hβ indices that are too low to be fitted by the
Thomas et al. (2003) models. This has been previously observed in
the spectra of dwarf galaxies by Poggianti et al. (2001), who sug-
gest that current models overestimate the turnoff temperatures in
the models. Such galaxies are interpreted as old stellar systems,
with ages > 8 Gyr for the purposes of this paper.
3.5 Comparison of colour with age and metallicity
Dwarf ellipticals without star formation that are blue in colour are
expected to be metal poor, whereas those with redder colours are
expected to be metal rich. Unfortunately, we are unable to derive
ages and metallicities for the two redder dwarfs (galaxies CGW 31
and CGW 49), as not all the spectral features could be measured
reliably.
One galaxy, 54 (Table 1), is younger and more metal rich than
the other dwarfs for which we obtained a good set of indices. It
has an age < 5 Gyr, and a metallicity [Fe/H] > -0.33. The (B−R)0
colour of this particular galaxy is 1.22, and it lies on the fit between
colour and magnitude for the cluster ellipticals and dwarf galaxies.
The other galaxies for which ages and metallicities are determined
are CGW 16, and CGW 20. These galaxies are metal poor and have
(B−R)0 colours of 1.18, 1.19 and 1.16 respectively. These galaxies
again lie close to the colour-magnitude relation for the cluster ellip-
ticals and dwarf galaxies. Therefore it appears that both young and
old dwarf galaxies in Perseus lie on the same colour-magnitude re-
lation. This is confirmed with ages obtained using the HαA index,
when again both young and old dwarf galaxies in Perseus lie on the
same colour-magnitude relation.
We find no correlation between either metallicity or age, de-
termined using the HαA index, with (B−R)0 colour as shown in
Figures 8a and 8b. However, we do find a clear anti-correlation be-
tween age and metallicity (Fig. 8c). The older the dwarf elliptical,
Table 3. Line-strength indices
The dwarf galaxy population in the Perseus Cluster
9
Mg b
. . .
<Fe>
[MgFe]'
HαA
HαF
. . .
. . .
3.0 ± 0.5
1.9 ± 0.3
0.8 ± 0.2
1.5 ± 0.2
0.36 ± 0.1
0.70 ± 0.1
0.38 ± 0.2
0.96 ± 0.3
. . .
. . .
1.6 ± 0.2
1.3 ± 0.1
. . .
2.5 ± 0.7
0.5 ± 0.1
0.5 ± 0.1
2.4 ± 0.3
2.5 ± 0.3
. . .
. . .
. . .
. . .
. . .
. . .
0.71 ± 0.2
0.36 ± 0.1
1.7 ± 0.2
0.56 ± 0.2
0.69 ± 0.2
2.3 ± 0.4
. . .
. . .
. . .
. . .
. . .
. . .
. . .
1.0 ± 0.3
1.7 ± 0.3
1.7 ± 0.1
1.1 ± 0.1
2.1 ± 0.2
3.0 ± 0.4
1.9 ± 0.3
2.6 ± 0.5
. . .
1.2 ± 0.3
1.2 ± 0.3
1.8 ± 0.1
2.2 ± 0.2
1.8 ± 0.2
2.1 ± 0.3
1.6 ± 0.2
2.2 ± 0.4
Galaxy
CGW 10
CGW 16
CGW 20
CGW 21
CGW 29
CGW 31
CGW 38
CGW 39
CGW 45
CGW 47
CGW 49
54
Hβ
2.7 ± 0.5
3.8 ± 0.7
2.5 ± 0.4
. . .
. . .
. . .
1.7 ± 0.2
1.4 ± 0.2
1.1 ± 0.2
2.7 ± 0.5
. . .
2.35 ± 0.6
4.2 ± 1.0
1.9 ± 0.5
2.8 ± 1.2
1
Line-strength indices as determined by the definitions in Trager et al. (1998) and Nelan et al. (2005). The combined iron index is defined as <Fe> =
2 (Fe5270+Fe5335) and the [α/Fe]-insensitive index [MgFe]' is defined as [MgFe]' = [(Mgb)(0.72 × Fe5270 + 0.28 × Fe5335)]0.5 (Thomas et al.
2003). Not every index was measured for every galaxy.
Figure 7. Plot of Hβ versus Mg b, <Fe> and [MgFe]' (left to right respectively) for the Perseus dwarfs superimposed on Thomas et al. (2003) model
isochrones and isometallicity lines for [α/Fe] = 0.3. The different ages and metallicities are labelled.
(a) (b) (c)
Figure 8. (a) Plot of (B−R)0 colour against the Mg b index to show how metallicity varies with colour. (b) (B−R)0 colour plotted against ages determined
using the HαA index. (c) Plot of age against the Mg b index to show how age varies with metallicity.
10
S. J. Penny and C. J. Conselice
Figure 9. WIYN Harris R band image of the irregular galaxy at α = 03 19
05.1, δ = +41 28 12.4. The solid black line in this figure is 10" in length.
the more metal poor it is. This has been found in previous studies
such as that of Poggianti et al. (2001), who find an anti-correlation
between age and metallicity for galaxies in the Coma Cluster in any
luminosity bin.
3.6 Irregular galaxy
We also obtained a spectrum of an irregular galaxy at α = 03 19
05.1 and δ = +41 28 12.4. From the absorption lines, this galaxy
was found to be at a radial velocity of 4032 ± 39 km s−1, con-
firming it as a member of the Perseus Cluster. A second object was
present in the slit for this galaxy, which had Balmer lines in emis-
sion in its spectrum. Based on the positions of these emission lines,
the additional object was found to be at z ≈ 0.055, giving the object
a radial velocity of ≈ 16500 km s−1. This would make it either a
background galaxy, or an interacting galaxy with a velocity differ-
ence of ≈ 12450 km s−1 from the main galaxy.
From the image of this galaxy, a tidal tail is clearly visible to
the left hand side of the galaxy as seen in Figure 9. This irregu-
lar galaxy is therefore most likely a merged system. Assuming the
Perseus Cluster is at a distance of 77 Mpc, this tidal tail has a length
of ≈ 7.5 kpc.
4 INTERPRETATION AND DISCUSSION
We find that the radial velocity dispersion for dwarf galaxies, σ =
1818 km s−1, is higher than the value of 1324 km s−1 found by
Struble & Rood (1999) for the giant galaxies in Perseus. These re-
sults agree with results found by Conselice et al. (2001a) for dwarf
galaxies in the Virgo Cluster. An infall origin for dwarfs can explain
this difference in velocity dispersion between the brighter cluster
galaxies and the dwarfs. Infalling spirals can become morphologi-
cally transformed into dwarf ellipticals by high speed interactions
with other cluster galaxies, and by the gravitational potential of the
cluster itself (harassment). Another possibility is that infalling dark
matter halos enter the cluster at later times, with the gas in the halos
being converted into stars due to tidal forces during the cluster as-
sembly. The velocity characteristics of low-mass, infalling galaxies
should change little over several gigayears, thus the velocity dis-
persion would represent the mass profile of the cluster at the time
of accretion.
We also find in Perseus an increase in the scatter of the colour-
magnitude relation for magnitudes fainter that MB = −15.5, which
can be interpreted as a diversity in the ages and metallicities of the
low mass systems. A range in the ages and metallicities of low-
mass systems suggests that dwarf galaxies are not a homogeneous
population, as redder, more metal rich cluster dwarfs must have
been formed more recently than the blue population.
De Lucia at al. (2004) suggest that a large fraction of faint red
galaxies in current clusters moved onto the colour-magnitude rela-
tion relatively recently, with their star formation activity coming to
an end at around z ≈ 0.8. Clusters at redshift z ≈ 1 have a deficit
of red-sequence galaxies, with the dwarf cluster population becom-
ing progressively fewer in number as we move out to increasing
redshifts. The rate of infalling spiral galaxies is expected to peak
at z ≈ 0.8 in a CDM cosmology with ΩM = 1 and H0 = 50 km
s−1 (Kauffmann 1995), the same redshift at which De Lucia at al.
(2004) suggest a large fraction of faint red galaxies in current clus-
ters moved onto the colour-magnitude relation. If dEs are indeed
formed from infalling spirals, it seems likely that they originate
from this epoch.
We find that the faint-end luminosity function is less steep than
what was found in previous studies of the Perseus Cluster, and in
fact matches the value of α = −1.26 found by Trentham et al.
(2005) for the field. We remove all galaxies incorrectly classified
as dwarfs by Conselice et al. (2003) from the luminosity function,
to refit this faint end slope.
The shape of the luminosity function at faint magnitudes is
dependent on many processes, one of which is the efficiency at
which gas is converted into stars (e.g. Dekel & Silk (1986). Gas
loss from supernova driven winds would reduce the star formation
efficiency. Star formation and/or nuclear activity in dwarf galax-
ies could also photoionize the intergalactic medium, regulating the
galaxy formation process (Efstathiou 1992). Another effect on the
faint-end slope is that dark halos in denser, evolved clusters are
more efficient at collecting gas used to form stars than those in un-
evolved clusters (Trentham & Hodgkin 2002). Yet another process
changing the shape of the faint end slop is dynamical stripping of
higher mass galaxies that would increase the number of low-mass
systems in a cluster (Conselice 2002). Also, dwarfs in the inner re-
gions of dense clusters are less able to survive due to harassment
(Moore et al. 1998), causing the faint-end slope to flatten for such
environments.
Whilst our value of α = −1.26 ± 0.06 for dwarfs in the
Perseus Cluster core is comparable with that of the field, it is also
similar to the value of α = −1.23 ± 0.13 (Pracy et al. 2004) in
the core region of Abell 2218. In Abell 2218, the slope of the lu-
minosity function varies with cluster environment, going from α
= −1.23 ± 0.13 in the central core of the cluster to α = −1.49
± 0.06 outside of the core. Pracy et al. (2004) infer that the core
of Abell 2218 is "dwarf depleted", with the dwarf-to-giant ratio
decreasing monotonically with increasing cluster density. Galaxy
clusters with less prominent dwarf populations are in fact those
with the highest projected galaxy densities (Phillipps et al 1998b).
The reason for this depletion in the core regions of clusters can
perhaps be explained by galaxy harassment. Tidal effects become
more important in the centre of clusters, with smaller spheroidal
The dwarf galaxy population in the Perseus Cluster
11
galaxies being easily destroyed (Conselice 2002). The lowest sur-
face brightness objects will disintegrate in the inner regions, de-
creasing the dwarf-to-giant ratio, whilst in outer regions such galax-
ies would be able to survive.
More recently, Harsono & De Propris (2007) find no evidence
for a faint-end upturn in the luminosity function in clusters at z =
0.3. There will be less contamination from background galaxies in
clusters at these redshifts, and therefore such galaxies would not
artificially steepen the faint-end slope to the degree seen in local
clusters. Harsono & De Propris (2007) argue that the faint-end up-
turn in the LF in nearby clusters is of recent origin, although our
results could contradict this theory as we find no evidence for a
faint-end slope upturn.
A hierarchical model of galaxy formation predicts that within
a given environment, low-mass galaxies would be the first to form.
However, it appears that many low mass galaxies finish forming
after the giants, through "downsizing". For example, results from
De Lucia at al. (2004) suggest that a large fraction of dwarf galax-
ies in current clusters moved onto the red sequence relatively re-
cently, with a cessation in star formation activity at z ≪ 0.8. This is
also seen for field galaxies (Bundy et al. 2006), suggesting down-
sizing is a possible origin for dwarf galaxies in both environments.
We find both old, metal-poor dwarfs and young, metal rich dwarfs
in Perseus. This is consistent with results by Poggianti et al. (2001)
for faint galaxies in Coma, where dwarf galaxies occupy a large
region on the SSP model grid, covering a range of metallicity and
ages, with most lying in the region [Fe/H] < −0.25 and ages < 1
Gyr.
These results imply that downsizing is a real effect, with the
younger dwarfs ending star formation at some time after the giant
cluster members. Whilst some dwarfs are indeed primordial, not all
cluster dwarfs are old. The formation of these young dwarfs may
involve the removal of mass from infalling higher-mass spirals (e.g.
Conselice et al. 2003).
5 CONCLUSIONS
We have analysed Keck/LRIS spectra for twenty-three dwarf
galaxy candidates in the Perseus Cluster, from which we have con-
firmed cluster membership for twelve systems based on radial ve-
locities measured from absorption lines.
We extend the confirmed member colour-magnitude relation
for Perseus down to MB = −12.5, finding that the slope of the
colour-magnitude relation becomes bluer when the low-surface
brightness dwarf galaxies are included. The fainter dwarfs also
scatter more from the colour-magnitude relation, following the
trend observed by Rakos & Schombert (2004) for low-mass galax-
ies in Fornax and Coma. This scatter can be interpreted as a spread
in the metallicity distributions of dwarf galaxies, which has been
inferred for dwarf galaxies in other clusters by Rakos et al. (2001)
and Poggianti et al. (2001).
After removing non-members from the B-band luminosity
function for the Perseus Cluster we find a faint-end slope α =
−1.26 ± 0.06, similar to the field. Previous studies of other galaxy
clusters have found that the dwarf-to-giant ratio is a function of
local projected galaxy density. Extending this study to the outer
regions of Perseus would enable us to see how the luminosity func-
tion changes with the local galaxy density.
Colours, morphologies, and central surface brightnesses are
not sufficient criteria to confirm cluster membership, as this work
has shown. Cluster membership cannot be confirmed without
spectroscopy, so the faint-end luminosity functions calculated for
galaxy clusters where membership has not been confirmed spectro-
scopically are most likely artificially steepened by non-members
resulting in higher dwarf-to-giant ratios.
By comparing the observed dwarf spectral absorption indices
with population synthesis models of Thomas et al. (2003), we de-
rive luminosity-weighted ages and metallicities for the dwarf galax-
ies. A range of ages is observed, ranging from older than 8 Gyr to
younger than 5 Gyr. The metallicity distribution of the faint cluster
members is also not that of a simple homogeneous population, with
the younger galaxies typically having higher metallicities.
More observations of dwarf galaxies in rich, nearby cluster en-
vironments are required in order to help improve our understanding
of the formation scenarios for cluster dwarfs. Further spectroscopic
work with a larger sample would also allow better constraints on
the ages and metallicities of cluster dwarf galaxies and would help
with modelling the formation of such galaxies.
ACKNOWLEDGMENTS
This research has made use of the NASA/IPAC Extragalactic
Database (NED) which is operated by the Jet Propulsion Labora-
tory, California Institute of Technology, under contract with the Na-
tional Aeronautics and Space Administration. We thank the staff of
Keck, WIYN and NOAO for help in obtaining the spectroscopy and
imaging presented here. We also thank Jay Gallagher for help with
the WIYN optical imaging, and Dolf Michielsen for useful advice.
S. J. P. acknowledges the support of a PPARC studentship. We also
wish to acknowledge the highly significant cultural role and rever-
ence that the indigenous Hawaiian community hold for the summit
of Mauna Kea. It is a privilege to be able to conduct observations
from this mountain.
REFERENCES
Andreon, S., Cuillandre, J. C., Puddu, E., Mellier, Y., 2006, MN-
RAS, 372, 60
Barazza, F. D., Binggeli, B., Jergen, H., 2002, A&A, 391, 823
Beijersbergen, M., Hoekstra, H., van Dokkum, P. G., van der
Hulst, T., 2002, MNRAS, 329, 385
Brodie, J. P., Huchra, J. P., 1990, ApJ, 362, 503
Brodie, J. P., Strader, J.,Denicol´o, G., Beasley, M. A., Cenarro, A.
J., Larsen, S. S., Kuntscner, H., Forbes, D. A., 2005, AJ, 129,
2643
Bundy, K., et al, 2006, ApJ, 651, 120
Conselice, C. J., 2002, ApJ, 573, L5
Conselice, C. J., Gallagher, J. S., III, Wyse, R. F. G., 2001a, ApJ,
559, 791
Conselice, C. J., Gallagher, J. S., III, Wyse, R. F. G., 2001b, AJ,
122, 2281
Conselice, C. J., Gallagher, J. S., III, Wyse, R. F. G., 2002, AJ,
123, 2246
Conselice, C. J., Gallagher, J. S., III, Wyse, R. F. G., 2003, AJ,
125, 66
Conselice, C. J., 2006, ApJ, 639, 120
De Lucia, G., et al., 2004, ApJ, 610, L77
De Propris, R., Pritchett, C. J., 1998, AJ, 116, 1118
De Rijcke, S., Dejonghe, H., Zeilinger, W. W., Hau, G. K. T.,
2003, A&A, 400, 119
Dekel, A., Silk, J., 1986, ApJ, 303, 39
12
S. J. Penny and C. J. Conselice
Efstathiou, G., 1992, MNRAS, 256, 43p
Fabian, A. C., Hu, E. M., Cowie, L. L., Grindlay, J., 1981, ApJ,
248, 47
Harsono, D., De Propis, R., 2007, MNRAS, accepted (astro-ph
0706.2994)
Held, E. V., Mould, J. R., 1994, AJ, 107, 1307
Kauffmann, G., 1995, MNRAS, 274, 153
Kent, S. M., Sargent, W. L. W., 1983, AJ, 88, 6
Lotz, J. M., Miller, B. W., Ferguson, H. C., 2004, ApJ, 613, 262
Mieske, S., Hilker, M., Infante, L., Mendes de Oliveira, C., 2007,
A&A, 463, 503.
Mobasher, B., et al., 2003, ApJ, 587, 605
Moore, B., Lake, G., Katz, N., 1998, ApJ, 495, 139
Nelan, J. E., Smith, R. J., Hudson, M. J., Wegner G. A., Lucey, J.
R., Moore, S. A. W., Quinney, S. J., Suntzeff, N. B., 2005, ApJ,
632, 137
Oke, J. B., Cohen, J. G., Carr, M., Cromer, J., Dingizian, A., Har-
ris, F. H., 1995, PASP, 107, 375
Phillips, S., Parker, Q. A., Schwartzenberg, J. M., Jones, J. B.,
1998a, ApJ, 493, L59
Phillipps, S., Driver, S. P., Couch, W. J., Smith, R. M., 1998b,
ApJ, 498, L119
Poggianti, B. M., et al., 2001, ApJ, 562, 689
Pracy, M. B., De Propris, R., Driver, S. P., Couch, W. J., Nulsen,
P. E. J., 2004, MNRAS, 352, 1135
Rakos, K., Schombert, J., Maitzen, H. M., Prugovecki, S., Odell,
A., 2001, AJ, 121, 1974
Rakos, K., Schombert, J., 2004, AJ, 127, 1502
Sabatini, S., Davies, J., Scaramella, R., Smith, R., Baes, M., Lin-
der, S. M., Roberts, S., Testa, V., 2003, MNRAS, 341, 981
Smith, R. J., 2005, MNRAS, 359, 975
Strader, J., Brodie, J. P., Schweizer, F., Larsen, S. S., Seitzer, P.,
2003, AJ, 125, 626
Struble, M. F., Rood, H. J., 1999, ApJS, 125, 36
Thomas, D., Maraston, C., Bender, R., 2003, MNRAS, 339, 897
Trager, S. C., Worthey, G., Faber , S. M., Burnstein, D., Gonz´alez,
J. J., 1998, ApJS, 116, 1
Trager, S. C., Faber, S. M., Worthey, G., Gonz´alez, J. J., 2000, AJ,
119, 1645
Trentham, N., Hodgkin, S., 2002, MNRAS, 333, 423
Trentham, N., Sampson, L., Banerji, M., 2005, MNRAS, 357, 783
Tully, R. B., Somerville, R. S., Trentham, N., Verheijen, M. A. W.,
2002, ApJ, 569, 573
White, S. D. M., Frenk, C. S., 1991, ApJ, 379, 52
Worthey, G., Ottaviani, D. L., 1997, ApJS, 111, 377
|
astro-ph/0007295 | 1 | 0007 | 2000-07-19T23:17:27 | Hierarchical galaxy formation and substructure in the Galaxy's stellar halo | [
"astro-ph"
] | We develop an explicit model for the formation of the stellar halo from tidally disrupted, accreted dwarf satellites in the cold dark matter (CDM) framework, focusing on predictions testable with the Sloan Digital Sky Survey (SDSS) and other wide-field surveys. Subhalo accretion and orbital evolution are calculated using a semi-analytic approach within the Press-Schechter formalism. Motivated by our previous work, we assume that low-mass subhalos (v < 30 km/s) can form significant populations of stars only if they accreted a substantial fraction of their mass before the epoch of reionization. With this assumption, the model reproduces the observed velocity function of galactic satellites in the Local Group, solving the ``dwarf satellite problem'' without modifying the popular LCDM cosmology. The disrupted satellites yield a stellar distribution with a total mass and radial density profile consistent with those observed for the Milky Way stellar halo. Most significantly, the model predicts the presence of many large-scale, coherent substructures in the outer halo. These substructures are remnants of individual, tidally disrupted dwarf satellite galaxies. Substructure is more pronounced at large galactocentric radii because of the smaller number density of tidal streams and the longer orbital times. This model provides a natural explanation for the coherent structures in the outer stellar halo found in the SDSS commissioning data, and it predicts that many more such structures should be found as the survey covers more of the sky. The detection (or non-detection) and characterization of such structures could eventually test variants of the CDM scenario, especially those that aim to solve the dwarf satellite problem by enhancing satellite disruption. | astro-ph | astro-ph | Astrophysical Journal, submitted
Preprint typeset using LATEX style emulateapj v. 04/03/99
0
0
0
2
l
u
J
9
1
1
v
5
9
2
7
0
0
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
HIERARCHICAL GALAXY FORMATION AND SUBSTRUCTURE IN THE GALAXY'S STELLAR
HALO
James S. Bullock1, Andrey V. Kravtsov2 and David H. Weinberg
Department of Astronomy, The Ohio State University, 140 W. 18th Ave, Columbus, OH 43210-1173
Astrophysical Journal, submitted
ABSTRACT
We develop an explicit model for the formation of the stellar halo from tidally disrupted, accreted
dwarf satellites in the cold dark matter (CDM) framework, focusing on predictions testable with the
Sloan Digital Sky Survey (SDSS) and other wide-field surveys. Subhalo accretion and orbital evolution
are calculated using a semi-analytic approach based on the extended Press-Schechter formalism. Moti-
vated by our previous work, we assume that low-mass subhalos (vc < 30 km s−1) can form significant
populations of stars only if they accreted a substantial fraction of their mass before the epoch of reioniza-
tion. With this assumption, the model reproduces the observed velocity function of galactic satellites in
the Local Group, solving the "dwarf satellite problem" without modifying the basic tenets of the popular
Λ+CDM cosmological scenario. The tidally disrupted satellites in this model yield a stellar distribution
whose total mass and radial density profile are consistent with those observed for the Milky Way stellar
halo. Most significantly, the model predicts the presence of many large-scale, coherent substructures in
the outer halo. These substructures are remnants of individual, tidally disrupted dwarf satellite galaxies.
Substructure is more pronounced at large galactocentric radii because of the smaller number density of
tidal streams and the longer orbital times. This model provides a natural explanation for the coherent
structures in the outer stellar halo found in the SDSS commissioning data, and it predicts that many
more such structures should be found as the survey covers more of the sky. The detection (or non-
detection) and characterization of such structures could eventually test variants of the CDM scenario,
especially those that aim to solve the dwarf satellite problem by enhancing satellite disruption.
Subject headings: cosmology: theory – galaxies:formation
1.
INTRODUCTION
The origin of the Milky Way's stellar halo is a long-
standing astronomical problem. The poles of the debate
are defined by the monolithic collapse model of Eggen,
Lynden-Bell, & Sandage (1962) and the chaotic accre-
tion model of Searle (1977) and Searle & Zinn (1978).
The Searle & Zinn picture has gained currency in recent
years in part because of growing recognition that the halo
and bulge are distinct components that may have differ-
ent formation mechanisms (see, e.g., the reviews by Wyse
1999ab) and in part because of "smoking gun" evidence
that includes the tidally distorted Sagittarius dwarf galaxy
(Ibata et al. 2000a and references therein) and the pres-
ence of extra-tidal stars around many dwarf spheroidal
satellites (Gould et al. 1992; Irwin & Hatzidimitriou 1995;
Kuhn, Smith & Hawley 1996). The Searle & Zinn scenario
also bears a strong anecdotal resemblance to the hierarchi-
cal galaxy formation scenario characteristic of inflationary
cold dark matter (CDM) cosmological models. In this pa-
per, we make the connection between CDM cosmology and
hierarchical stellar halo formation much more explicit, by
presenting a simple but quantitative model for halo for-
mation in the CDM framework and obtaining predictions
for the degree of residual substructure in the outer halo.
Previous studies of stellar halo formation in the hierar-
chical framework have focused on the fossil evidence for
satellite disruption preserved in phase space substructure
1james,andrey,[email protected]
2Hubble Fellow
1
of the halo stars (e.g., Johnston, Spergel & Hernquist 1995;
Helmi & White 1999). These studies were aimed primarily
at exploiting surveys of the halo in the solar neighborhood
(e.g., Arnold & Gilmore 1992; Preston, Beers & Shect-
man 1994; Majewski, Munn & Hawley 1994, 1996). Wide-
angle, deep, multi-color surveys, such as the Sloan Digital
Sky Survey (SDSS; York et al. 2000), open up new avenues
for studying the structure of the stellar halo. RR Lyrae
stars, detected by their variability and color, can provide
a three-dimensional map of the distribution of halo stars.
The more general population of A-colored halo stars can be
used for the same purpose; relative to RR Lyrae they have
the advantages of greater numbers and detectability in a
single observation epoch but the disadvantage of being less
precise standard candles. Studies of RR Lyrae stars and
A-colored stars in SDSS commissioning data have already
revealed two large substructures in the outer halo (Ivezi´c
et al. 2000; Yanny et al. 2000). The photometric depth
of the SDSS and the intrinsic brightness of RR Lyrae and
A stars allows a probe of halo structure out to large dis-
tances, ∼ 75 kpc, and the restricted absolute-magnitude
range of RR Lyrae and A stars prevents 3-dimensional
substructure from being washed out by projection. Car-
bon star surveys (e.g., Ibata et al. 2000a) and surveys of
giant stars (Majewski et al. 2000) offer similar prospects.
The model presented in this paper is an extension of
our previous work aimed at explaining the observed abun-
2
BULLOCK, KRAVTSOV, & WEINBERG
dance of dwarf satellite galaxies in the Local Group within
the CDM framework (Bullock, Kravtsov & Weinberg 2000;
hereafter BKW). In BKW we showed that the observed
shape and amplitude of the velocity function of dwarf
satellite galaxies in the Local Group can be explained if gas
accretion and star formation are suppressed in low-mass
dark matter clumps after intergalactic gas is reheated dur-
ing the epoch of reionization. In this picture, the observed
dwarf satellites around our Galaxy are those that assem-
bled a large fraction of their mass before reionization and
survived the decay of their orbit as a result of dynamical
friction and avoided tidal disruption by the Milky Way
potential. As shown in Fig. 1 of BKW, the number of
tidally disrupted objects is similar to the number of sur-
viving dwarf satellites. Here we show that the disrupted
satellites produce a population of stars whose total mass
and radial profile are consistent with observations of the
Milky Way's stellar halo.
In the outer halo, where the
number of contributing satellites is relatively small and
the orbital times are long, the model predicts substantial
substructure, which should be detectable with the SDSS
and other deep, wide-angle surveys.
Our model is simple and is unlikely to be accurate in full
quantitative detail. However, the qualitative predictions
should be characteristic of conventional CDM cosmologies
combined with straightforward assumptions about the star
formation in low-mass dark matter potential wells. If the
observed stellar halo is found to be radically different from
these predictions, it will mean that either our star forma-
tion assumptions or the CDM predictions for hierarchical
small-scale structure in the dark matter distribution are
incorrect. In this sense, studies of the stellar distribution
in the outer halo can play a valuable role in testing more
general ideas about galaxy formation.
For convenience, we will focus on predictions for the RR
Lyrae distribution. RR Lyrae are especially useful probes
of the stellar halo because they are relatively easy to iden-
tify, they are luminous enough to be detected out to large
distances (r ∼ 100 kpc), and they are nearly standard can-
dles and therefore yield 3-dimensional maps. RR Lyrae
are believed to be good tracers of the more general halo
stellar distribution, and they have often been used in kine-
matic studies of the halo (Hawkins 1984). Finally, while
RR Lyrae are numerous enough to trace halo substructure,
they are rare enough that we can construct numerical real-
izations that contain every individual RR Lyrae star. The
stellar density fluctuations predicted by our model are far
in excess of Poisson fluctuations, so it is straightforward
to scale our predictions to other halo tracers like blue hor-
izontal branch stars or carbon stars, just by putting in
appropriate stellar population weights.
The remainder of the paper is organized as follows. In
§ 2 we will describe our model and assumptions. Specifi-
cally, we will describe our semi-analytic method for follow-
ing the accretion and orbital evolution of satellite galaxies
in § 2.1 and our modeling of stellar debris of disrupted
satellites in § 2.2. We present our results in § 3. We finish
with discussion and conclusions in § 4.
2. METHOD
2.1. Accretion and orbital evolution of satellites
We use a semi-analytic method to trace the accretion
history and orbital evolution of satellite galaxies within a
typical Milky Way-size dark halo.3 Detailed description
of the model is given in BKW. Here, we briefly review
its essential aspects. The model uses the extended Press-
Schechter formalism (Bond et al. 1991; Lacey & Cole
1993) to construct the accretion history for each galactic
halo. The mass of a halo in this formalism is accumulated
via accretion of individual subhalos of different masses, and
we keep track of all the accreted subhalos down to some
minimum mass. The second part of our model is a semi-
analytic prescription for orbital evolution of the accreted
subhalos. This prescription is used to determine whether
a subhalo survives to the present day, is tidally disrupted,
or is dragged into the central galaxy by dynamical fric-
tion. Only subhalos that accreted a significant fraction
( >
∼0.2 − 0.3) of their mass before intergalactic gas was re-
heated during the epoch of reionization are assumed to
host luminous galaxies. In the model presented in this pa-
per, the stellar halo of Milky Way-like galaxies is formed
from the debris of those subhalos that once hosted lumi-
nous galaxies but were tidally disrupted before the present
day. For our analysis, we adopt a flat CDM model with
a non-zero vacuum energy and the following parameters:
Ωm = 0.3, ΩΛ = 0.7, h = 0.7, σ8 = 1.0, where σ8 is the
rms fluctuation on the scale of 8h−1 Mpc, h is the Hubble
constant in units of 100 km s−1Mpc−1, and Ωm and ΩΛ
are the density contributions of matter and the vacuum
respectively in units of the critical density.
We assume that the density profile of each dark matter
halo is described by the NFW profile (Navarro, Frenk, &
White 1997): ρNFW(x) ∝ x−1(1 + x)−2, where x = r/rs,
and rs is a characteristic inner radius. Given a halo of
mass Mvir at redshift z, the model of Bullock et al. (2000a)
supplies the typical rs value and specifies the profile com-
pletely. The circular velocity curve, v2(r) ≡ GM (r)/r,
peaks at a value vm at a radius rm ≃ 2.16rs.
We use the merger tree method of Somerville & Kolatt
(1999) to construct mass growth and halo accretion his-
tories for an ensemble of galaxy-sized dark matter halos.
We start with halos of mass Mvir = 1.1 × 1012 h−1M⊙,
at z = 0, and trace subhalo accretion histories back to
z = 10. We record the mass growth for the primary halo,
Mvir(z), as well as the mass of each accreted subhalo, Ma,
and the time of its accretion, ta (or za). We assign the
subhalo vm according to the mass-velocity relation at the
epoch of accretion. For the results presented below, we
use 100 ensembles of formation histories for galactic host
halos. We obtain very similar results if this number is
increased.
Each subhalo is assigned an initial orbital circularity,
ǫ, defined as the ratio of the angular momentum of the
subhalo to that of a circular orbit with the same energy,
ǫ ≡ J/Jc. We choose ǫ randomly in the range 0.1 − 1.0
(Ghigna et al. 1998). To determine whether the accreted
halo's orbit will decay, we use Chandrasekhar's formula to
calculate the decay time, τ c
DF , of the orbit's circular radius
3In this paper, the term "halo" sometimes refers to a dark matter halo and sometimes to a stellar halo. Usually the meaning is clear from
context, but we will specify "dark" or "stellar" where necessary. The term "subhalo" always refers to a dark matter halo, one that is accreted
into a larger dark matter halo before redshift zero.
HIERARCHICAL GALAXY FORMATION AND SUBSTRUCTURE IN THE STELLAR HALO
3
rc - the radius of a circular orbit with the same energy
as the actual orbit. Each subhalo is assumed to start at a
randomly assigned radius ra
c = (0.4 − 0.75)Rvir(ta), where
Rvir(ta) is the virial radius of the host halo at the time of
accretion. We determined this distribution of circular radii
by measuring the range of binding energies of subhalos in
the ART simulations used by Klypin et al (1999a). Once
τ c
DF is known, the decay time for the given circularity is
τDF = τ c
DF ǫ0.4 (Colpi et al. 1999). If τDF is smaller than
the time left between accretion and z = 0, τDF ≤ t0 − ta,
then the subhalo will merge with the central object. In our
modeling of the stellar halo, we do not consider the contri-
bution due to galaxies that subsequently merge with the
central object. Due to the rapid decay of the orbits, any
debris associated with these objects will likely remained
confined within the radius where stripping first becomes
important, typically r ∼< 10 kpc. For this reason, we con-
sider the predictions for the stellar distribution only for
galactocentric radii r > 10 kpc.
If τDF is too long for the orbit to have decayed com-
pletely (τDF > t0 − ta), we check whether the subhalo
would have been tidally disrupted. We assume that the
halo is disrupted if the tidal radius becomes smaller than
rm. The tidal radius, rt, is determined at the pericenter of
the orbit at z = 0, where the tides are the strongest, fol-
lowing Klypin et al. (1999b). If rt ≤ rmax we declare the
subhalo to be tidally destroyed and record its mass and
orbital parameters so that we may model the evolution of
its tidal debris (§2.2).
The resulting average mass functions for all accreted ha-
los and the subset of the halos that were tidally disrupted
are shown in Figure 1 with the thin dashed and solid lines,
respectively. The error bars represent the run-to-run dis-
persion over 100 realizations. For comparison, the dotted
line corresponds to halos that were dragged to the halo
center as a result of dynamical friction. As expected, the
dynamical friction is more efficient for massive subhalos.
The mass function of surviving subhalos is similar to that
of the disrupted halos, but we omit it here to preserve
visual clarity.
As in BKW, we assume that of all subhalos with vm <
30 km s−1 only those that accreted a fraction f of their
gas before the redshift of reionization, zre, host luminous
galaxies. Although our results are consistent with the ob-
served abundance of dwarf satellites for a range of values of
f and zre, here we use the fiducial values of BKW, f = 0.3
and zre = 8. We expect that our results for disrupted satel-
lites would be similar if we chose other f , zre combinations
that also match the observed (surviving) dwarf satellite
population. For a given subhalo of mass Ma and accretion
redshift za, we use equation (2.26) of LC93 to probabilisti-
cally assign the redshift, zf , at which the main progenitor
of the subhalo reached mass Mf = f Ma for the first time.
The subhalo hosts a luminous galaxy only if zf ≥ zre. We
also assume that subhalos with vm < vl = 10 km s−1 do
not host galaxies, since any gas that was initially accreted
in these small systems would be unable to cool, and it
should quickly boil out of the halo after reionization (e.g.,
Barkana & Loeb 1999). Our results do not change if we
vary vl by 50%. The thick solid line in Figure 1 shows the
average mass function of disrupted halos that once hosted
luminous galaxies. The mass function of the surviving
galaxies is similar to that of the disrupted galaxies, but
Fig. 1.- Cumulative mass function of all accreted dark
matter subhalos (dashed line), the fraction that decayed due to
dynamical friction (dotted line), tidally disrupted halos (thin
solid line), and the fraction of disrupted halos that host galax-
ies (thick solid line). Not shown is the mass function of sur-
viving halos, which is similar to that of disrupted halos, and
the mass function of surviving galaxies, which is roughly half
that of the disrupted galaxies (reflecting the tendency for sur-
viving halos to have been accreted later). The mass function
represents the average over 100 merger histories for host halos
of mass Mvir(z = 0) = 1.1 × 1012h−1 M⊙. The errorbars show
the dispersion over the different merger histories.
lower in amplitude by about a factor of two. Surviving
halos are typically accreted later than tidally destroyed
halos, and they are less likely to form before reionization
and host a galaxy.
2.2. Modeling stellar tidal debris
Estimating the number of tidally stripped stars, and
RR Lyrae stars in particular, requires several uncertain
assumptions about gas cooling, star formation, and stel-
lar population morphology. The following approximations
are extremely simplified, and the estimated number of dis-
rupted stars for each object is uncertain at the factor of
2 − 3 level. This uncertainty is passed on to the overall
amplitude of calculated stellar density distribution in the
halo. Nevertheless, our statistical measures of substruc-
ture depend only on the ratio to the background density,
and they are thus largely insensitive to the precise values
we assume. Furthermore, we adopt the same assumptions
that we used in BKW to obtain consistency with the ob-
served dwarf satellite population, and this matching nor-
malizes out some of the uncertainties in the overall stellar
halo amplitude.
We estimate the luminosity of every disrupted luminous
galaxy assuming that it has a baryonic mass f Ma(Ωb/Ωm)
(the mass of baryons accreted at z > zre). We assume that
a fraction ǫ∗ of this baryonic mass is converted to a stel-
4
BULLOCK, KRAVTSOV, & WEINBERG
lar population with mass-to-light ratio M∗/LV . The total
mass to light ratio of the subhalo is thus
(cid:18) Mvir
LV (cid:19) = f −1(cid:18) Ωm
Ωb (cid:19)(cid:18) M∗
LV (cid:19) ǫ−1
∗ .
(1)
Adopting M∗/LV ≃ 0.7, typical for galactic disk stars
(e.g., Binney & Merrifield 1998), Ωm/Ωb ≃ 7 (based on
Ωm = 0.3, h = 0.7, and Ωbh2 ≃ 0.02 from Burles & Tytler
1998), and ǫ∗ = 0.5, we obtain (M/LV ) ≃ 10f −1 ≃ 33.
We estimate the number of horizontal branch stars in each
galaxy using NHB = LV /(540L⊙) (Preston, Shectman, &
Beers 1991). The fraction of the horizontal branch stars
that are RR Lyrae variables is strongly dependent on the
metallicities and ages of the populations, and it will vary
significantly from one object to another. For simplicity, we
assume that all disrupted objects have NRR = 0.3NHB.
This fraction is high compared to local halo stars, but
it is consistent with fractions observed for more distant
(r ∼> 10 kpc) globular clusters, likely reflecting the ten-
dency for the outer halo to be younger (Preston et al.
1991; Brocato et al. 1996; Layden 1998).
For each disrupted luminous subhalo, we randomly as-
sign a direction for its angular momentum vector. This
direction fixes the plane of the orbit since we assume that
the dark halo potential is spherical (i.e., no orbital preces-
sion). We follow the orbit from the time it was accreted
at ta to t0 using
dr
dt
dψ
dt
= ±v(rc)s 2
v2(rc)
[Φ(re) − Φ(r)] + 1 −
ǫ2r2
c
r2 ,
=
v(rc)rcǫ
r2
,
(2)
(3)
where r is the distance to the galactic center, ψ is the an-
gle in the orbital plane, Φ(r) = −4.6v2
m ln(1 + x)/x is the
potential of the host, and the ± sign signifies whether the
object is approaching its apocenter or pericenter, respec-
tively. We will work in the approximation that the dynam-
ical friction timescale is long compared to the time remain-
ing for the orbital evolution and disruption: τDF ≫ t0−ta.
This is a good approximation for ∼ 90% of the disrupted
halos - not surprisingly, since we have deliberately re-
stricted our analysis to halos whose orbits have not de-
cayed (i.e., long τDF ). In order to approximately account
for cases where this approximation breaks down, we start
integrating the satellite orbit at t = ta but set its starting
radius equal to the circular radius it will have decayed to
by t = t0: r(ta) = rc(t0) ≡ rc. The initial value for the
angle, ψ(ta), is chosen randomly, and we assume that the
satellite is initially infalling (approaching the pericenter).
We assume that the satellite is tidally disrupted after
the first passage of its orbit pericenter. At this time, the
tidal debris will obtain an energy distribution from the en-
counter. Our model for the evolution of the debris along
the tidal tail is motivated by numerical results of Johnston
(1998), who showed that the following approximations pro-
vide good description of the positions of stripped particles
in her simulations. The typical energy scale of the debris
is set by the change in the host halo potential energy over
the size of the tidally disrupted object,
ε = rt
dΦ
dr ≃ v2
m(cid:18) rt
rp(cid:19) ,
(4)
where rp is the pericenter radius of the orbit. The last ap-
proximation is exactly true for a logarithmic potential (a
singular isothermal density profile). We assume that the
satellite is completely disrupted after the first passage, and
that the energy of the debris is evenly distributed over the
energy range −ε > dE > ε (Evans & Kochanek 1989).
This assumption ignores the possibility of disruption over
several orbits, but we find that our results are robust to
the choice of distribution and do not change significantly
if the energy range is altered by 50%. Using intuition
gained from a circular orbit within a singular isothermal
density background, we may estimate how a change in en-
ergy from E to E + dE affects the orbit of a particle. In
this approximation, the azimuthal period, Tψ, and radial
period, Tr, depend only on the orbit energy, and they are
both increased or decreased depending on the sign of the
deposited energy: Tψ,r(E + dE) = τ Tψ,r(E), where
τ = exp(cid:18) dE
m(cid:19) .
v2
(5)
This result allows us to map the orbital trajectory of the
initial object with energy E (Eqs. 4 and 5) to that of
a debris particle with energy E + dE via [r(t),ψ(t)] →
[r(t/τ ),ψ(t/τ )].
For each RR Lyrae star in the disrupted galaxy, we as-
sign a change in energy dE and integrate the orbital equa-
tions to determine its position at z = 0. Since the dis-
rupted galaxy will have some finite spherical extent, we
add a random offset to this calculated central orbit posi-
tion. The magnitude of this offset is a Gaussian deviate
with dispersion 2 kpc, the typical optical radius for a dwarf
galaxy (Mateo 1998), and the direction is random.
3. RESULTS
Figures 2 and 3 show two realizations of the RR Lyrae
distribution from disrupted satellites in sky-projected
galactic coordinates. The panels in each figure corre-
spond to the indicated radial bins in galacto-centric ra-
dius. Qualitatively, it is clear that substructures become
more pronounced at larger radii. This radial trend re-
flects the smaller number of disrupted satellites with large
apocenters and the longer periods of their orbits, which
reduces the extent of debris spreading along the orbit. A
comparison of the maps in Figures 2 and 3 illustrates the
differences between different merger histories of the host
halo. One can see that the stochastic variations in merger
history at fixed host mass lead to substantial variations in
the appearance and abundance of substructures.
In light of the SDSS results referred to in § 1, the most
interesting predictions of the model are the radial number
density profile of halo stars and the clumpiness and spa-
tial extent of the stellar distribution. Figure 4 shows radial
number density profiles of the halo RR Lyrae stars. The
long dashed lines in each panel represent the power law
(n∗ ∝ r−3) RR Lyrae density profile derived by Wetterer
& McGraw (1996), based on their large compilation of RR
Lyrae. The solid points show the profile computed
HIERARCHICAL GALAXY FORMATION AND SUBSTRUCTURE IN THE STELLAR HALO
5
Fig. 2.- Distribution of stripped stars in various radial bins projected on the sky. Each point represents an RR Lyrae star, and
the number of stars in each radial bin, starting in the upper left panel, is 11331, 9052, 8237, 7182, 6173, and 5076. These views are
centered on the Galactic Center, but shifting to a solar origin makes no qualitative difference.
6
BULLOCK, KRAVTSOV, & WEINBERG
Fig. 3.- Same as Figure 2, but for a different merger history realization. Each point represents an RR Lyrae, and the number
of stars in each radial bin, starting in the the upper left panel, is 6511, 6402, 4962, 4363. 2783, and 2077.
HIERARCHICAL GALAXY FORMATION AND SUBSTRUCTURE IN THE STELLAR HALO
7
Fig. 4.- The average RR Lyrae density profile (thick solid
line) compared with the Ivezi´c et al. (2000) SDSS data (solid
points) and the power law determined by Wetterer & Mc-
Graw (1996) (dashed line). In the upper left panel, the error
bars reflect the dispersion in the average from realization to
realization. The thin solid lines in the other three panels are
results of random strips similar in solid angle and geometry to
the strips used to obtain the SDSS measurements (∼ 1 degree
wide, 100 square degree strip).
by Ivezi´c et al. (2000) for their sample of RR Lyrae can-
didates obtained from SDSS commissioning data, which
covers roughly a one-degree wide, 100 square degree strip
of sky. Note that the SDSS and Wetterer & McGraw pro-
∼35 kpc. At larger radii, however, the
files agree well at <
SDSS sample shows two significant deviations from the
smooth n∗ ∝ r−3 profile: a "bump" in number density
∼50 kpc. As noted
at r ≈ 40 kpc and a sharp drop at r >
by Ivezi´c et al., this structure in the radial profile likely
indicates significant clumpiness of the stellar halo at these
galactocentric radii, and the bump in particular is asso-
ciated with an identifiable coherent structure containing
∼ 70 RR Lyrae within the observed region.
The model predictions are shown in Figure 4 by thick
and thin solid lines. The thick solid lines represent the
computed RR Lyrae number density profile averaged over
all merging history realizations and the full sky. The error
bars in the upper left panel show the dispersion from real-
ization to realization around this average, demonstrating
that stochastic variations in merger histories lead to a fac-
tor of ∼ 2 rms variation in the overall normalization of the
predicted halo density profile. In the remaining three pan-
els, the thin solid lines show examples of density profiles
derived from a single host halo realization, viewed through
three randomly chosen strips similar in solid angle and ge-
ometry to the strips used to derive the SDSS sample.
Fig. 5.- The radial distributions of disrupted RR Lyrae
stars in 1 degree wide, great circle slices through a single model
halo realization. This halo formed from about 60 disrupted
satellites. Shown are four random cuts of great circle planes
through the halo center. Each point represents a single RR
Lyrae star. Concentric circles indicate galactocentric radii of
25, 50, and 75 kpc. Note that apparent "clumpiness" of the
stellar distribution increases with increasing radius.
The first remarkable feature of Figure 4 is the agreement
of the predicted average profile with the slope and ampli-
tude found by Wetterer & McGraw (1996). As discussed
in §2.2, the predicted amplitude is uncertain by a factor
of several, and even the statistical fluctuations from one
galaxy to another are significant, so the degree of agree-
ment must be somewhat fortuitous. However, it is worth
noting that we did not adjust any parameters to fit the ob-
served halo profile but chose "best guess" values based on
other considerations - in particular, the requirement of
matching the observed dwarf satellite population. Figure 4
suggests that disruption of accreted satellites can produce
not just substructure in the stellar halo but the entire stel-
lar halo itself. If we have overestimated the number of RR
Lyrae per unit dark matter mass, then there is room for
another physical mechanism that creates a smooth under-
lying halo, but it seems that no such additional mechanism
is necessary.
The second remarkable feature of Figure 4 is the jagged-
ness of the observed and predicted profiles of individ-
ual strips, which becomes especially pronounced at radii
r ∼> 40 kpc. These large fluctuations reflect the substruc-
ture visible in Figures 2 and 3. The steep drop in the
SDSS counts between 60 and 70 kpc suggests detection of
an "edge" of the stellar halo (Ivezic et al. 2000). However,
the second of our numerical realizations shows an equally
sharp edge, even though the average halo profile is smooth.
Our model predicts a gradual steepening of the halo profile
at r ∼> 60 kpc, but although surveys in small solid angles
should show large count fluctuations, the profile averaged
8
BULLOCK, KRAVTSOV, & WEINBERG
radii.
Fig. 6.- Same as Figure 5, but for a different merger his-
tory realization of the host halo. The host in this realization
accreted about 20 luminous satellites over its history.
over the full sky should not cut off sharply.
Figures 5 and 6 present a different view of the structure
associated with disrupted satellites, in a form more com-
parable to the plots of Yanny et al. (2000). Each figure
shows a one degree wide, randomly oriented, great circle
slice through a realization of the RR Lyrae distribution.
The two figures show stellar halos for different Monte Carlo
accretion histories, one with a total of about 60 tidally de-
stroyed galaxies (Figure 5) and one with a more quiescent
accretion history and about 20 tidally destroyed galax-
ies (Figure 6). This range roughly covers the scatter in
the number of disruption events expected from galaxy to
galaxy. The three concentric circles indicate galactocen-
tric radii of 25, 50, and 75 kpc. In both figures, structures
associated with the individual disrupted objects become
more easily identifiable at larger radii.
To make our predictions more quantitative, we present
two simple statistical measures of the halo clumpiness.
Figure 7 shows the probability distribution of model RR
Lyrae counts in solid angle cells of different sizes and
for different ranges of galactocentric radii. The cells are
roughly square on the sky; they are defined by dividing
the sky map for each realization into patches of a given
size. In order to take out the uncertainty in the overall
amplitude of the density distribution, and to factor out
the variation in the overall amplitude from one halo real-
ization to another, the counts are presented in units of the
average expected number, hNi, of RR Lyrae in patches
of the chosen size for each realization. Figure 7 shows
that the amplitude of patch-to-patch count fluctuations is
higher for smaller solid angles and for larger galactocentric
Fig. 7.- Probability distribution of simulated RR Lyrae
counts in solid angle cells for various radial ranges. Here,
N/hN i is the measured number of stars within the given ra-
dial bin and solid angle cell divided by the average number for
a cell of that size. The cells are roughly square in angle size
and were defined by dividing the sky of each realization into
patches of the indicated size, with the observer at the cen-
ter of the halo. The spikes at small N/hN i represent empty
cells. Note that higher amplitude fluctuations from patch to
patch are more likely for smaller patch areas and for larger
galactocentric radii.
Figure 8 shows the rms dispersion in RR Lyrae counts,
σ(N/hNi), as a function of galactocentric radius for cells
of different solid angle and geometry. The solid points/line
show counts in square cells, while open points represent the
counts in one degree wide strips. The error bars for the
solid points show the standard deviation of the dispersion
for different merger history realizations; the error bars are
similar for the open points. These error bars reinforce the
point made in our illustrative figures above, namely that
the degree of surviving substructure is quite variable from
one realization of the stellar halo to another.The dashed
lines show the expected amplitude of Poisson fluctuations
based on the average number of stars expected within the
given solid angle and radial bin. Note that the predicted
fluctuations are always larger than Poisson fluctuations,
especially at large radii, because they are dominated by
fluctuations in the number of debris streams rather than
√N fluctuations in the number of RR Lyrae. Again, the
fluctuations are larger for cells of smaller solid angle and
for larger radii. They are also larger for square patches
than for narrow strips of the same solid angle, reflecting
the fact that a narrow strip is less likely to enclose an
entire disrupted object and instead encloses fragments of
multiple debris streams.
HIERARCHICAL GALAXY FORMATION AND SUBSTRUCTURE IN THE STELLAR HALO
9
erage predicted profile falls off more steeply at r ∼> 50 kpc,
but it does not have a sharp outer boundary.
The amplitude of the mean profile depends on several
uncertain factors, such as the mean stellar luminosity to
dark mass ratio of the disrupted dwarfs and the mean num-
ber of RR Lyrae per solar luminosity. There are also sig-
nificant (factor of two) statistical fluctuations in the ampli-
tude from one realization to another. The amplitude of the
predicted RR Lyrae profile is therefore uncertain by a fac-
tor of a few. Nonetheless, with "best guess" parameter val-
ues chosen on the basis of other considerations, the mean
RR Lyrae profile agrees very well with that determined by
Wetterer & McGraw (1996), in slope and amplitude.
It
therefore appears that disruption of dwarf satellites is a
plausible mechanism for producing the entire stellar halo
within the CDM framework, though it could also co-exist
with some other mechanism. The uncertainty in our pre-
dicted normalization is reduced by the fact that we require
the model to self-consistently reproduce the velocity func-
tion of observed dwarf satellites, a point that we will return
to shortly.
The main qualitative prediction of the model is the pres-
∼30
ence of significant clumpiness in the outer regions ( >
kpc) of the Galaxy's stellar halo. This clumpiness is due
to the surviving tidal debris of dozens of satellite galax-
ies disrupted during evolution of their host. In the inner
∼30 kpc), the density distri-
regions of the stellar halo (r <
bution is relatively smooth. At larger radii, however, the
clumpiness of the stellar halo manifests itself when viewed
through fixed solid angle patches in the sky. For typical
modeled stellar halos, RR Lyrae profiles of the type ob-
served by Ivezi´c et al. (2000) in the SDSS commissioning
data (with a coherent structure at r ∼ 50 kpc) are not
uncommon.
We have quantified our predictions by presenting some
statistical measures of the "clumpiness" in our modeled
stellar halos. First, we measured the probability distribu-
tion (see Fig. 7) of RR Lyrae counts, N/hNi, in solid angle
cells of different sizes and for different galactocentric ra-
dial bins (here hNi is the average number of stars expected
in a cell). We also presented the rms width of this prob-
ability distribution as a function of galactocentric radius
for solid angle cells of different geometries (Fig. 8). These
statistics show that the variance in the stellar counts (i.e.,
clumpiness of the stellar halo) increases with increasing
galactocentric radius and decreasing solid angle. Although
current observational data sets are not sufficiently large
to derive similar statistics for comparison, future samples
of RR Lyrae and A-stars from the SDSS and 2dF surveys
should make such comparisons possible. For cell sizes of 36
deg2 or larger, predicted fluctuations in RR Lyrae counts
are much larger than Poisson fluctuations, at all galacto-
centric radii considered.
These quantitative predictions depend on some of the
model assumptions. For instance, one of the implicit as-
sumptions in our analysis is that the dark matter distri-
bution in the host halo is nearly spherically symmetric.
In particular, we do not include the possible precession
of satellite orbits. In the case of the significantly oblate
halo, such precession can erase signatures of the tidal tails
(at least in configuration space), so it would reduce the
Fig. 8.- Fractional rms fluctuation of stellar counts as a
function of radius for sky patches of different solid angle. The
solid points are for a roughly square angular patch geometry,
and the open points correspond to one degree wide strips. The
error bars represent the dispersion in the rms fluctuation from
realization to realization for an ensemble of merger histories.
The dashed lines show the expected Poisson uncertainty based
on the average number of RR Lyrae expected within the given
solid angle and radial bin.
4. DISCUSSION AND CONCLUSIONS
We have presented a model in which stellar halos of
Milky Way type galaxies are built via the accretion and
tidal disruption of satellite galaxies. The model is based
on the CDM structure formation scenario, with the crucial
assumption that only a fraction of dark matter halos with
∼30km s−1 are luminous and host a siz-
circular velocities <
able stellar system4. These luminous halos are those that
collapsed and accreted a substantial fraction of their mass
prior to the epoch of reionization, zre. Only the accretion
and tidal disruption of luminous dwarf galaxies contribute
to the build up of the stellar halo in our model.
The model predicts an average density profile for the
stellar halo of n∗ ∝ r−α, with α ∼ 3 in the range
r ≃ 10 − 50 kpc, in good agreement with observations.
The density profiles of individual realizations of the host
halo merger histories, however, exhibit a substantial scat-
ter around this average, with the power law varying from
α ∼ 2.5 − 3.5. The reason why the stellar halo density
profile is steeper than that of the dark matter (αDM ∼ 2,
which is indeed typical of that of the surviving subhalos in
our model), is that central satellites are more likely to be
destroyed than those at large radii, since satellites far from
the galactic center must have extremely eccentric orbits in
order to pass close enough to be tidally destroyed. The av-
4All DM halos with circular velocities > 30km s−1 are assumed to host a luminous galaxy.
10
BULLOCK, KRAVTSOV, & WEINBERG
predicted clumpiness of the outer halo (e.g., Ibata et al.
2000a). However, the observed narrowness of the tidal tail
of the Sagittarius dwarf galaxy implies a nearly spherical
∼60 kpc (Ibata et al. 2000a). A nearly
halo potential at r <
spherical mass distribution in the inner region of a galaxy-
mass halo is consistent with predictions of CDM models
(Bullock et al. 2000b). We have also neglected the effects
of the disk component on the background potential. In-
cluding this non-spherical central component would induce
precession in the tidal orbits (Helmi & White 1999) and
smear out residual structure at small galactic radii. But
since this effect would only be important in the central
regions (r <
∼20 kpc), the net effect would be to strengthen
the trend of increased variance in star counts with radius.
The assumption that affects our predictions the most,
however, is that only a small fraction of disrupted subhalos
host a stellar system and thereby contribute to the build
up of the stellar halo. In other words, the predicted prop-
erties of the stellar halo depend crucially on the way that
we have solved the dwarf satellite problem. This prob-
lem, namely that the predicted number of dark halos with
∼30 km s−1 is much larger than the ob-
circular velocities <
served number of dwarf satellites within the virialized dark
halos of the Milky Way and M31 (Kauffmann et al. 1993;
Klypin et al. 1999a; Moore et al. 1999), is one of the
few outstanding problems of the conventional inflationary
CDM model, which, with the inclusion of a cosmological
constant, accounts well for a wide variety of other obser-
vational data.
Other proposed solutions to the dwarf satellite problem
include modifying the inflationary fluctuation spectrum
(Kamionkowski & Liddle 2000) or modifying the proper-
ties of dark matter by making it warm (WDM; e.g., Hogan
& Dalcanton 2000) or self-interacting (SIDM; e.g., Spergel
& Steinhardt 2000). In the SIDM model, halos collapse
and accrete mass in a similar manner to conventional CDM
halos. However, the number of surviving subhalos within
a Milky Way mass halo is reduced because lower concen-
tration of the SIDM halos makes them easier to disrupt
and because interactions lead to "ram pressure" stripping
of the dark halos. Relative to our reionization solution,
the SIDM solution seems to predict a more massive and
more extended stellar halo built from a larger number of
tidal streams, since the abundance of dwarf satellites is re-
duced by a higher efficiency of satellite disruption rather
than suppression of star formation in low-mass systems.
Indeed, the SIDM model seems at some risk of overpro-
ducing the stellar halo. However, if the predicted stellar
halo were normalized to the mean stellar density of the ob-
served one, the SIDM model would probably predict less
substructure than the model presented here because of the
higher density of independent tidal streams.
The predictions of WDM and broken scale-invariance
models are less clear, since they can reduce dwarf satel-
lite numbers both by suppressing their formation in the
first place (White & Croft 2000) and by making them
less concentrated and therefore easier to disrupt (Col´ın
et al. 2000). If the second effect dominates, the predic-
tions might be closer to those of SIDM; if the first effect
dominates, they might be closer to those of the reioniza-
tion model. At present, we lack quantitative predictions
on this matter from SIDM and WDM, and we lack detailed
observational constraints, so we cannot draw conclusions
about which model fares best. However, it appears that
studies of the mass, radial profile, and substructure of the
stellar halo might provide useful constraints on these mod-
els in the near future.
It is interesting that all of the substructures detected in
the nearly all-sky Carbon star survey and the SDSS com-
missioning data could be produced by a single tidal stream
of the Sagittarius dwarf galaxy (Ibata et al. 2000b). In
particular, the observed excess in the RR Lyrae number
density profile at r ≈ 45 kpc is caused by a clump of stars
near the apocenter of the Sagittarius orbit. If most of the
RR Lyrae in the surveyed strip belong to the Sagittarius
stream, then no stars are expected beyond the apocenter
of the Sagittarius orbit, and this would naturally explain
∼60 kpc de-
the drop in the RR Lyrae number density at r >
tected in the SDSS data. However, the presence of only a
single stream in a large volume of sky would be both puz-
zling and intriguing, since the CDM models considered in
this paper predict at least ∼ 20 (and typically more) tidal
streams in the halo of a Milky Way-size galaxy. The SIDM
and WDM stellar halos should be built from even larger
number of tidal streams. Carbon stars are relatively young
and rare, and therefore older tidal streams may not be re-
vealed by their distribution. We predict, however, that
many more substructures not associated with the Sagit-
tarius tidal stream should be detected in the future as the
SDSS covers a larger area of the sky. Absence of such de-
tections would spell serious trouble for CDM models, and
possibly even more so for the variants of CDM that we
have discussed.
Several past theoretical studies of tidal stripping and
disruption of galactic satellites have shown that tidal
streams can be used as powerful probes of both the present
potential of the Milky Way (Johnston et al. 1999; Ibata et
al. 2000a) and its accretion history (e.g., Helmi & White
1999; Helmi et al. 1999; Helmi & de Zeeuw 2000). For
example, Johnston et al.
(1999) show that parameters
characterizing the mass distribution in the Milky Way halo
can be determined with an accuracy of a few percent using
a single tidal stream if accurate phase-space information
is available for as few as 100 stream stars. Helmi et al.
(1999) show how phase-space information can be used to
recover fossil remnants of the satellites accreted and dis-
rupted early in the Milky Way evolution. These techniques
can be used to construct a detailed formation history of
the Galaxy.
In this study, we have focused on the spatial distribution
of stars in the Milky Way halo. The radial density behav-
ior of stars may not prove to be as sensitive a probe of the
galactic potential as the accurate phase-space mapping of
tidal streams (Johnston et al. 1999), but it should be very
useful in recovering details of the Milky Way accretion
history. In some respects, the number of surviving tidal
streams or the degree of clumpiness of the stellar distribu-
tion can provide better constraints on the variants of the
CDM scenario than the abundance of galactic satellites.
An advantage here is that the spatial distribution of halo
stars should be possible to map in the very near future,
when large samples of halo stars from the SDSS and other
surveys become available. Accurate, large-scale, phase-
space mapping of tidal streams, on the other hand, will
become possible only after launch of the next generation
astrometric space missions (i.e., at the end of the decade).
HIERARCHICAL GALAXY FORMATION AND SUBSTRUCTURE IN THE STELLAR HALO
11
On the theoretical side, our model and predictions can be
improved upon by combining mass accretion histories typ-
ical for the galaxy-size halos formed in CDM models with
more sophisticated numerical models of orbital evolution
of satellites and their tidal debris.
We thank Amina Helmi and Kathryn Johnston for use-
ful discussion and suggestions. This work was supported
in part by NASA LTSA grant NAG5-3525 and NSF grant
AST-9802568. Support for A.V.K. was provided by NASA
through Hubble Fellowship grant HF-01121.01-99A from
the Space Telescope Science Institute, which is operated
by the Association of Universities for Research in Astron-
omy, Inc., under NASA contract NAS5-26555.
12
BULLOCK, KRAVTSOV, & WEINBERG
REFERENCES
Arnold, R., & Gilmore, G., 1992, MNRAS, 257, 225
Barkana, R., & Loeb, A. 1999, ApJ, 523, 54
Binney, J., & Merrifield, M., 1998, Galactic Astronomy (Princeton:
Princeton University Press)
Johnston, K.V., Spergel, D.N. & Hernquist, L. 1995, ApJ, 451, 598
Johnston, K.V., Zhao, H., Spergel, D.N., Hernquist, L. 1999, ApJ,
512, L109
Kamionkowski, M., & Liddle, A.R. 1999, preprint
(astro-
Bond, J. R., Cole, S., Efstathiou, G., Kaiser, N., 1991, ApJ, 379,
ph/9911103)
440
Brocato, E., Buonanno, R., Malakhova, Y., & Piersimoni, A. M.,
Kauffmann, G., White, S.D.M., & Guiderdoni, B. 1993, MNRAS,
264, 201
1996, A&A, 311, 778
Klypin, A.A., Kravtsov, A.V., Valenzuela, O., & Prada, F. 1999a,
Bullock, J.S., Flores, R., Kolatt, T.S., Kravtsov, A.V., Klypin, A.,
ApJ, 522, 82
& Primack, J.R.. 2000b, in preparation.
Bullock, J.S., Kolatt, T.S., Sigad, Y., Somerville, R.S., Kravtsov,
A.V., Klypin, A., Primack, J.R., & Dekel, A. 2000a, MNRAS,
submitted (astro-ph/9908159)
Bullock, J.S., Kravtsov, A.V., & Weinberg, D.H., ApJ, accepted,
astro-ph/0002214, BKW
Burles, S., & Tytler, D. 1998, ApJ, 507, 732
Col´ın, P. Avila-Reese, V., Valenzuela, O. 2000, ApJ submitted
(astro-ph/0004115)
Colpi, M., Mayer, L., & Governato, F. 1999, ApJ, 525, 720.
Dav´e, R., Spergel, D.N., Steinhardt, P.J., Wandelt, B.D. 2000, ApJ
submitted (astro-ph/0006218)
Eggen, O.J., Lynden-Bell, D., & Sandage, A.R., 1962, ApJ, 136, 748
Evans, C. R.,& Kochanek, C. S., 1989, ApJ, 346. L13
Ghigna, S., Moore, B., Governato, F., Lake, G., Quinn, T., & Stadel,
J. 1998, MNRAS, 300, 146
Gould, A., Guhathakurta, P., Richstone, D., & Flynn, C., 1992,
Klypin, A.A., Gottlber, S., Kravtsov, A.V., & Khokhlov, A.M.
1999b, ApJ, 516, 530
Kuhn, J. R., Smith, H. A., & Hawley, S. L., 1996, ApJ, 469, L93
Lacey, C., & Cole, S., 1993, MNRAS, 262, 627 (LC93)
Layden, A.C., 1998, in ASP Conf. Ser. 136, Galactic Halos, ed. D.
Zaritsky, (San Fransico:ASP), 14
Majewski, S.R., Munn, J.A., & Hawley, S.L, 1994, ApJ, 427, L37
Majewski, S.R., Munn, J.A., & Hawley, S.L, 1996, ApJ, 459, L73
Majewski, S.R., Ostheimer, J.C., Patterson, R.J., Kunkel, W.E.,
Johnston, K.V., Geisler, D. 2000, AJ 119, 760
Mateo, M.L. 1998, ARA&A, 36, 435
Moore, B., Ghigna, F., Governato, F., Lake, G., Stadel, J., & Tozzi,
P. 1999, ApJ, 524, L19
Narayanan, V.K., Spergel, D.N., Dav´e, R., Ma, C.-P. 2000, ApJ,
submitted (astro-ph/0005095)
Navarro, J.F., Frenk, C.S., White, S.D.M. 1997, ApJ, 490, 493
Press, W.H., & Schechter, P. 1974, ApJ, 187, 425
Preston, G.W., Schectman, S.A., & Beers, T.C., 1991, ApJ, 375,
Preston, G.W., Beers, T.C., & Schectman, S.A., 1994, AJ, 108, 538
Searle, L., 1977,
in The Evolution of Galaxies and Stellar
Populations, ed. B.M. Tinsley & R.B. Larson (New Haven: Yale
University Press),219
Searle, L. & Zinn, R., 1978, APJ, 225, 357
Somerville, R.S., & Kolatt, T.S. 1999, MNRAS, 305, 1 (SK99)
Spergel, D.N., & Steinhardt, P.J. 1999, preprint (astro-ph/9909386)
Wetterer, C.J., & McGraw, J.T. 1996, AJ, 112, 1046
White, M., & Croft, R.A., 2000, ApJ, submitted, astro-ph/0001247
Wyse, R. F. G. 1999a, in ASP Conf. Ser. 165: The Third Stromlo
Symposium: The Galactic Halo, 1
Wyse, R. F. G. 1999b, in The formation of galactic bulges, ed. C.M.
Carollo, H.C. Ferguson, R.F.G. Wyse, (Cambridge: Cambridge)
Yanny, B., et al., 2000, ApJ, accepted, astro-ph/0004128
York, D.G., et al., 2000, AJ, accepted, astro-ph/0006396
Hawkins, M.R.S. 1984, MNRAS 206, 433
Helmi, A., & White, S.D.M. 1999, MNRAS 307, 495
Helmi, A., White, S.D.M., de Zeeuw, P.T., & Zhao, H.-S. 1999,
121.
ApJ, 388, 345
Nature, 402, 53
ph/0007166
Helmi, A., & de Zeeuw, P.T. 2000, MNRAS submitted, astro-
Hogan, C.J., & Dalcanton, J.J. 2000, Phys.Rev. D,
in press
(astro-ph/0002330)
Ibata, R., Lewis, G.F., Irwin, M., Totten, E., Quinn, T. 2000a,
astro-ph/0004011
Ibata, R., Irwin, M., Lewis, G.F., Stolte, A. 2000b, ApJL submitted
(astro-ph/0004255)
Ibata, R.A., & Razoumov A.O. 1998, A&A, 336, 130
Irwin, M., & Hatzidimitriou, D., 1995, MNRAS, 277, 1354
Ivezi´c, Z, et al., 2000, AJ, submitted, astro-ph/0004130
Johnston, K.V., 1998, ApJ, 495, 297
|
astro-ph/9408104 | 1 | 9408 | 1994-08-31T23:26:39 | Matter Distribution for Power Spectra with Broken Scale Invariance | [
"astro-ph"
] | To test the primordial power spectra predicted by a double inflationary model with a break of amplitude $\Delta=3$ at a scale of $2\pi/k\approx 10 \hm$ and CDM as dominant matter content, we perform PM simulations with 128$^3$ particles on a 256$^3$ grid. The broken scale invariance of the power spectra explains the extra power observed in the large-scale matter distribution. COBE-normalized spectra and a linear biasing with $b\approx 2$ are shown to reproduce the reconstructed power spectra from the CfA catalog. Identifying galactic halos with overdensity of approximately two times the cell variance, we can fit the angular correlation function using both the Limber equation and creating a APM-like angular projection with the observed luminosity function. Finally, the higher order moments of the galaxy distribution are shown to fit reasonably well the observed values. | astro-ph | astro-ph | Matter distribution for power spectra
with broken scale invariance
Luca Amendola
, Stefan Gottl(cid:127)ober
, Jan Peter M(cid:127)ucket
, Volker M(cid:127)uller
;
Osservatorio Astronomico di Roma
Viale del Parco Mel lini,
I- Rome, Italy
NASA/Fermilab Astrophysics Center
PO Box
Batavia IL , USA
Astrophysical Institute Potsdam
An der Sternwarte
D- Potsdam, Germany
ABSTRACT
To test the primordial power spectra predicted by a double in(cid:13)ationary model
with a break of amplitude (cid:1) = at a scale of (cid:25)=k (cid:25) h
Mpc and CDM as
(cid:0)
dominant matter content, we perform PM simulations with
particles on a
grid. The broken scale invariance of the power spectra explains the extra
power observed in the large-scale matter distribution. COBE-normalized spec-
tra and a linear biasing with b (cid:25) are shown to reproduce the reconstructed
power spectra from the CfA catalog. Identifying galactic halos with overdensity
of approximately two times the cell variance, we can (cid:12)t the angular correlation
function using both the Limber equation and creating a APM-like angular pro-
jection with the observed luminosity function. Finally, the higher order moments
of the galaxy distribution are shown to (cid:12)t reasonably well the observed values.
Introduction
In(cid:13)ationary cosmological models predict that the observable part of the universe is quasi-
(cid:13)at, i.e. the total density parameter is (cid:10)
(cid:25) . Combined with the density of baryons
tot
(cid:10)
(cid:25) : h
(h denotes the Hubble constant in units of km s
Mpc
) following
bar
(cid:0)
(cid:0)
(cid:0)
from the theory of primordial nucleosynthesis this requires most of matter in the Universe
to be nonbaryonic. The most elaborate model of structure formation assumes an universe
dominated by cold dark matter (CDM) with a Harrison{Zeldovich spectrum of initial pertur-
bations. With a biasing b
(cid:25) it has successfully explained the observed hierarchy of cosmic
g
structures on scales smaller than approximately h
Mpc. However, recent observations
(cid:0)
seem to indicate that the predictions of the standard CDM model on very large scales and
on small scales are incompatible (Efstathiou et al. , Maddox et al. , Fisher et al.
, Saunders et al. , Vogeley et al. ).
The standard model is based on two assumptions, namely, () that the primordial per-
turbation spectrum generated during in(cid:13)ation is of Harrison-Zeldovich type, and () that
the dark matter, which gives the main contribution to the density of the Einstein-de Sit-
ter universe, is cold. In order to change the theoretical predictions of this model, one can
modify either of these assumptions. In the (cid:12)rst case, improved in(cid:13)ationary scenarios lead
to other than Harrison-Zeldovich spectra. In the second case, widely discussed candidates
for closing the universe are mixed dark matter or a cosmological term. In a recent paper, a
change of the (cid:12)rst type has been investigated (Gottl(cid:127)ober, M(cid:127)ucket, Starobinsky ). The
underlying in(cid:13)ationary model shows two consecutive stages of exponential expansion with a
short intermediate stage of power law expansion (Gottl(cid:127)ober, M(cid:127)uller, Starobinsky ). The
two driving mechanisms are vacuum polarisation e(cid:11)ects and a massive scalar (cid:12)eld. In this
case the resulting power spectra with broken scale invariance (BSI) are of Harrison-Zeldovich
type only in the limit of very small and very large scales. In the intermediate range they are
steeper. The spectra are characterized by the ratio (cid:1) of the power of a Harrison-Zeldovich
spectrum to the power of the BSI spectrum on small scales assuming both are normalized on
large scales (COBE normalization). The scale k
denotes the onset of the break in the per-
br
turbation spectrum at small scales, i. e. for k > k
the spectrum is of Harrison-Zeldovich
br
type. The quantity (cid:1) mainly depends on the parameters characterising the in(cid:13)ationary
stages (the mass of the scalar particle and the coupling constant of the higher-order terms),
whereas k
depends on the energy density of the scalar (cid:12)eld at the onset of in(cid:13)ation. The
br
best (cid:12)t to observations is reached with (cid:1) between and and k
in the range between
br
(cid:0)
(cid:0)
(cid:0)
h
Mpc and h
Mpc (Gottl(cid:127)ober, M(cid:127)ucket, Starobinsky ). In the following we adopt
(cid:1) = and k
= :h
Mpc. Starting from these spectra, we have performed N-body sim-
br
(cid:0)
(cid:0)
ulations using the particle-mesh code of Kates et al. ( ), extended to three dimensions.
In this paper we report on the linear and non-linear clustering properties that characterize
the matter in these simulations. In Section we begin by discussing the spatial correlation
function and the power spectrum in our simulations.
One of the most convincing evidence for rejecting the standard CDM model comes prob-
ably from the angular correlation function (ACF). Even if the information on redshift is lost,
the angular surveys provide such an enormous amount of data, millions of galaxy positions,
that they turn out to tightly constrain theoretical models. As it is well known, the ACF
reported by Maddox et al. ( ) is not matched by the standard CDM model; on angular
scales larger than two or three degrees, the observed ACF is found indeed to be signi(cid:12)ca-
tively larger than predicted by CDM, when CDM is normalized to small scales. One can say
that since the publishing of the APM results the minimal requirement for any new model of
galaxy formation is that the correct ACF be reproduced. In Section we report the test of
our model against the observed ACF.
As we already stated, to match the correlation functions is only a minimal requirement,
though a very signi(cid:12)cant one, for a model to be acceptable. In recent years many higher
order statistical measures of the observed distribution of galaxies have been discussed and
calculated. A direct extension of the correlation function are the three- and four-point cor-
relation functions, thoroughly discussed in literature. Since their estimate is very noisy and
time-consuming, often some integral version of the n-point correlation functions is computed
from the data. Particularly simple and interesting are, for instance, the skewness and the
kurtosis of the count-in-cells (e.g. Coles & Frenk , Saunders et al. , Bouchet et
al. , Gazta~naga , ). In Section we derive the higher order moments of our
simulations, and compare them with the data. Finally, in Section we draw our conclusions.
Power spectrum and correlation function
We have performed N -body simulations with
particles in cubes with a
grid. The
simulations were made with four di(cid:11)erent box sizes, L = h
Mpc; h
Mpc; h
Mpc
(cid:0)
(cid:0)
(cid:0)
and h
Mpc in order to get a resolution high enough to identify the places where galactic
(cid:0)
halos form and, for the same spectrum and with the same method, to model the large scale
matter distribution (cp. Kates et al. ).
We consider a CDM model ((cid:10) = ; H = km/s/Mpc) with the transfer function of Bond
and Efstathiou ( ) and the primordial perturbation spectra with broken scale invariance
calculated by Gottl(cid:127)ober, M(cid:127)uller, and Starobinsky ( ). We have normalized the spectra
with the one year COBE result (cid:27)
(
) = ( (cid:6) :)(cid:22)K=:K (Smoot et al. ). The in-
T
(cid:14)
(cid:13)ationary model we are considering produces a negligible contribution of gravitational waves
to the microwave anisotropy. The rms multipole values C
for our perturbation spectrum are
l
substantially smaller than for the standard CDM model for l > (Gottl(cid:127)ober and M(cid:127)ucket,
). We did not include the new analysis presented in Wright et al. ( ) which leads to
a almost % increase ( % following G(cid:19)orski et al. ) of the large-scale normalization.
Taking this into account, the biasing factor calculated from the spectra would decrease by
the same amount.
The linear analysis of the power spectrum (Gottl(cid:127)ober, M(cid:127)ucket, Starobinsky ) has
shown that we need a bias b = : for transforming structures of dark matter particles
into luminous matter. The power spectrum one actually computes from a (cid:12)nite size, (cid:12)nite
resolution simulation is an acceptable approximation of the theoretical one only in a range
of wavelengths k for which (cid:25)=L (cid:28) k (cid:28) (cid:25)=l, where l is the cell size and L is the box size.
Then, to extend the range of validity of the power spectrum, we estimate the true P (k) by
joining four power spectra for di(cid:11)erent box sizes (with L = l). We take the highest P (k)
for any value of k , since the e(cid:11)ect of both (cid:12)nite size and (cid:12)nite resolution is generally to
reduce the power amplitude. The reconstructed power spectrum coincides with the linear
one at very large scales. At scales smaller than k
(cid:25) :h Mpc
the BSI model shows more
(cid:0)
nl
n
power than the linear spectrum and has a slope of k
with n (cid:25) (cid:0):. Consequently, also the
variance at h
Mpc increases. Therefore, the biasing factor de(cid:12)ned as b = (cid:27)
decreases.
(cid:0)
(cid:0)
From the simulations we found an optimal value of b (cid:25) : (instead of the linear .).
To compare the reconstructed spectrum with the spectrum of the CfA catalogue we
transform it into redshift space using the Kaiser ( ) and Peacock ( ) corrections. The
resulting spectrum is shown in Fig. , where we have assumed a biasing factor b = :.
For comparison the spectrum of a simulation with the standard CDM spectrum is also
included in this (cid:12)gure (b = : ). Note that in our simulation the highest scale mode (k
)
min
is overestimated by about % due to the chosen representation of the spectrum in k-space.
The maximum value of the calculated spectrum would really be at k (cid:25) : h Mpc
, and
(cid:0)
the spectrum will be bent down at small k 's as it is indicated by the data (though with very
large error bars).
We identify galaxies in our simulations by means of the peak-background split formalism.
Let (cid:27)
be the density (cid:13)uctuation variance on the grid cell of a given simulation, and (cid:23) (cid:27)
the selection threshold: only particles residing in regions with density contrast (cid:14) > (cid:23) (cid:27)
are identi(cid:12)ed with galaxies. Also, let (cid:24) be the correlation function of all particles, and
(cid:24)
the correlation function relative to the galaxies. The relation between the linear biasing
(cid:23)
=
b (cid:17) ((cid:24)
=(cid:24) )
and the threshold (cid:23) is quite complicated, because it involves the full probability
(cid:23)
distribution of the underlying (cid:12)elds, which is in general unknown. The simple formula given
Figure : Comparison of simulated power spectra using initial conditions with broken scale
invariance and a bias b = : (solid line), standard CDM model with b = : (dashed line)
and data from the CfA catalog (Vogeley et al. ).
in the classical paper by Kaiser ( ) cannot be applied here because it holds only if one
identi(cid:12)es each region above threshold with a single galaxy, while we put the density (cid:12)eld
in the regions above threshold equal to the underlying density (cid:12)eld. Moreover, the Kaiser
formula holds only for Gaussian (cid:12)eld, while the matter clustering is certainly non-Gaussian
by the present. A biasing scheme closer to ours has been investigated by Catelan et al.
( ), who give the relation b = b((cid:23); (cid:27)) for a lognormal (cid:12)eld, which is known to approximate
the real density probability distribution. The general trend, at least in the limit of small
correlation, is that b increases with increasing (cid:23) and decreasing (cid:27) (i.e., increasing the box
size). As a consequence, in a larger box one has to use a smaller threshold (cid:23) to get the same
ampli(cid:12)cation b. The lognormal relation better approximates our results: to have b (cid:25) : in
a (cid:12)eld with (cid:27) (cid:25) : (relative to the h
Mpc box) one needs a threshold (cid:23) (cid:25) . As
(cid:0)
we will show, with this threshold our model reproduces several observational features of the
galaxy clustering.
Finally, we have computed the spatial correlation function from the simulations. Using
all particles the BSI model yields a slope of . - . with a correlation radius of h
Mpc.
(cid:0)
A biasing procedure as described above changes the slope to . - ., and the correlation
radius increases to (. - .)h
Mpc (depending on the box length). In our simulation the
(cid:0)
standard CDM model yields a slope of . which is too steep, while its correlation radius is
(cid:0)
h
Mpc for all particles. In Fig. we present the spatial correlation functions for our BSI
model.
slope=1.7
100
10
1
0.1
0.01
0.001
1
10
Figure : Two-point correlation function in simulations of BSI spectra in the h
Mpc
(cid:0)
(dashed line) and h
Mpc (solid line) box for particles in cells with overdensity (cid:23) = :
(cid:0)
and (cid:23) = :, respectively.
The angular correlation function
We estimate the angular correlation function in two ways: indirectly, from the power spec-
trum of the N -body simulation, and directly, performing a pro jection of the N -body particles
on a portion of sphere according to the luminosity function (Coles et al. , Moscardini et
al. ).
From the knowledge of P (k), one can derive the ACF w((cid:18)) via the Limber equation:
R
R
kP (k)dk
y
(cid:30)
(y)J
(k(cid:18)y)dy
w((cid:18)) =
;
()
R
(cid:25)
(
y
(cid:30)(y)dy)
where (cid:30)(y) = y
exp [(cid:0)(y=y
)
] is the selection function. Inserting the linear power spectra
(cid:0) :
(cid:3)
into Eq. () and scaling the angular correlation function to the Lick depth (y
(cid:25) h
(cid:3)
(cid:0)
Mpc) one can compare the theoretical predictions with the observational data for w((cid:18))
from the APM survey (Maddox et al. ). However, the use of linear approximation in
the calculation of w((cid:18)) is justi(cid:12)ed for large (cid:18) only. For angles up to (cid:18) (cid:24)
, the agree-
(cid:14)
ment of the theoretical predictions with the data is very good, if spectra with (cid:1) = and
(cid:0)
(cid:0)
k
= (:::)h
Mpc are considered (Gottl(cid:127)ober, M(cid:127)ucket, Starobinsky ). The nonlinear
br
behaviour can be described by using the power spectra reconstructed from the di(cid:11)erent boxes.
As already mentioned on nonlinear scales the slope becomes P (k) / k
so that the theo-
(cid:0):
retical angular correlation w((cid:18)) matches the observed small-angular behaviour w((cid:18)) / (cid:18)
(cid:0) :
very well as long as the (cid:12)nite cell size of the smallest simulation box does not smear out the
correlations. The unbiased angular correlation function calculated in this way are shown in
Fig. a and b as dotted lines. In the case of the BSI model the line can be shifted to match
the data assuming a biasing as introduced above.
We have also computed the ACF by pro jecting the particles onto a spherical surface
(cid:14)
wide and estimating w((cid:18)) directly from the angular map. The details of this procedure
have been already given in literature (e.g. Coles et al. , Moscardini et al. ), so we
only sketch the method here. We have calculated the ACF both for all particles and for
particles selected by a density threshold as reported in Section . We have replicated the
simulation box by re(cid:13)ection in such a way that it covers a cone with opening
and depth
(cid:14)
(cid:24) h
Mpc, so as to reproduce the observational cone of the Lick catalog, to which the
(cid:0)
APM data have been scaled. We need three and eight levels for boxes of L = h
Mpc and
(cid:0)
L = h
Mpc, respectively, to contain the Lick cone (we did not consider the smallest and
(cid:0)
the biggest simulations here). We need the replication procedure to get both the necessary
resolution on small scales and the Lick depth. However, by this method we get also a small
systematic at (cid:18) >
, as can be seen by testing the procedure with a Poisson distribution.
(cid:14)
Therefore, we have restricted our calculations to (cid:18) <
. On small scales we are limited by
(cid:14)
the resolution of the simulations so that we considered only (cid:18) > :
.
(cid:14)
According to the luminosity function we assign an absolute magnitude to each particle
through a random process. From the distance to the observer, located in the center of the
`original' box, we derive the apparent magnitude m.
If m (cid:20) m
= : (see Maddox
Lick
et al. ), the particle is pro jected on the surface. In Fig. we show the corresponding
distribution of particles assuming a density threshold of particles per cell (see below). Since
the largest contribution comes from the box in which the observer is situated, and from the
box directly on top of it, we expect that the e(cid:11)ect introduced by the box replication is a
minor one, at least for not very large separations. To test this, we have also calculated the
ACF pro jecting only particles with brighter limiting apparent magnitudes, m
= : and
m
= :. In this way the characteristic depth D
= dex[(cid:0) :(m
(cid:0) M
(cid:0) )] becomes
(cid:3)
(cid:3)
much smaller, and the e(cid:11)ect of box duplication is greatly reduced. Then, we shift the ACF
to the Lick scale by means of the scale dependence contained in the Limber equation (see
e.g. Groth & Peebles , Peebles , Maddox et al. ). Our tests have con(cid:12)rmed
the reliability of the method of replication.
The estimator we use for the ACF is
w((cid:18)) = F
(cid:0) ;
()
C C
C R
where C C is the number of pairs in the real angular catalog at angular separation (cid:18) , C R is
the number of pairs in the crossed real-random catalogs, and F is the ratio of densities of
the random and the real catalog (F > to reduce the Poissonian noise). The results for the
discussed models are shown in Fig. a (standard CDM model) and b (BSI model). The
squares denote the angular correlation function for all particles. While the standard CDM
simulation clearly cannot (cid:12)t the angular correlation function, we (cid:12)nd a good agreement with
the dashed curves calculated from the reconstructed power spectrum. The scattering of the
data is due to measurements in di(cid:11)erent directions and measurements in di(cid:11)erent simulation
boxes.
To introduce biasing in our BSI simulations, we follow the procedure described in Section
. For retaining the full spatial resolution we identify galactic halo candidates on the grid by
imposing discrete thresholds without any smoothing. Assuming a threshold of particles per
cell ( h
Mpc, (cid:27)=.) or particles per cell ( h
Mpc, (cid:27) = :), and pro jecting the
(cid:0)
(cid:0)
galaxies according to the galaxy luminosity function, we get reasonable angular correlation
functions in the range of :
< (cid:18) <
(triangles in Fig. b). In comparison with the CDM
Figure : Angular pro jection of particles in cells with overdensity (cid:23) = :. This map should
be compared with the APM data.
model, the selected galaxies show clearly the required extra power at separations larger than
. The thresholds correspond to values of (cid:23) (cid:25) : and . for the smaller and larger box
sizes, respectively.
Higher-order moments
If the distribution of galaxies is not Gaussian, the two-point correlation function does not
fully characterize the clustering properties. Even if the initial distribution was Gaussian, as
our in(cid:13)ationary model predicts, non-Gaussianity is induced by the non-linear gravitational
e(cid:11)ects. Indeed, the deviations from Gaussianity can be expanded in a perturbative series of
the variance of the (cid:13)uctuation (cid:12)eld (Juszkiewicz et al. , Kofman & Bernardeau ).
Theory and observations agree in (cid:12)nding traces of non-Gaussianity up to very large scales,
where (cid:24) (r) (cid:28) .
A particularly simple measure of non-Gaussianity is provided by the higher order mo-
ments of the counts in cells. Let us denote with n
the number of galaxies in the i-th cell in
i
a partition of a given volume in N cells, and with ^n its mean. Let (cid:22)
be the dimensionless
m
central moment of order m, (cid:22)
=< (n
(cid:0) ^n)
> =^n
, and let (cid:20)
be the corresponding
m
i
m
m
m
dimensionless cumulant (or connected moment). To the (cid:12)rst few orders the relation between
(cid:22)
and (cid:20)
is: (cid:20)
= (cid:22)
; (cid:20)
= (cid:22)
; (cid:20)
= (cid:22)
(cid:0) (cid:22)
; etc. (see, e.g., Cramer ). For a
m
m
Gaussian (cid:12)eld, (cid:20)
= for m > . Since the galaxies are supposed to be a discrete sampling
m
of an underlying continuous (cid:12)eld, we should subtract from the cumulants the so-called shot-
noise terms, i.e. the corresponding moments of a Poissonian distribution. This is actually
just a hypothesis, because the galaxies could be far from being a Poissonian sampling of the
(cid:12)eld, but this is what has been routinely adopted in the data elaboration. In any case, it
should be important only in the limit of very large scales, when the distribution will tend
1
1
1
1
0.1
0.1
0.1
0.1
0.01
0.01
0.01
0.01
standard CDM
standard CDM
standard CDM
standard CDM
0.001
0.001
0.001
0.001
0.1
0.1
0.1
0.1
1
1
1
1
10
10
10
10
1
0.1
0.01
0.001
0.1
1
10
Figure : Comparison of the APM angular correlation function (Maddox et al. , dots)
with the simulations: a) CDM spectrum (squares); b) spectra with broken scale invariance,
both for all particles (squares) and for the biased particles (triangles). The dashed lines are
the correlation functions calculated from the Limber equation.
to be e(cid:11)ectively Poissonian, and of very small scales, when the discrete nature of galaxies
dominate over the clustering. Let us notice another point about the moment estimators (cid:20)
.
m
It is well known that when the moments of a distribution are evaluated from a sample, they
are biased estimators of the true moments (see e.g. Cramer ). They are good estimators
only in the limit of N ! , if N is the number of independent measures (i.e. the size
of the sample). We will oversample the simulation box with a large number of randomly
placed cells (of order
for the smallest cells, down to
for the largest ones), so that
the =N correction should not be relevant: however, our cells are clearly not independent,
so the biasing of the estimators can slightly a(cid:11)ect our results (as well as the data) on very
large scales.
Generally speaking, the cumulant (cid:20)
of a (cid:13)uctuation (cid:12)eld (cid:12)ltered through a window
n
W (R) with characteristic scale R is related to the n-point reduced correlation function (cid:24)
n
via the integral equation
"
#
Z
m
Y
(cid:0)m
(cid:20)
(R) = V
d
x
W
(x
)
(cid:24)
(x
; :::x
) :
()
m
i
R
i
n
m
R
V
R
i=
Through Eq. (), a model for the higher-order correlation functions is re(cid:13)ected in a model
for the cumulants (cid:20)
. The most popular of such models is probably the hierarchical scaling,
n
according to which any function (cid:24)
is proportional to products of n (cid:0) two-point functions
n
summed over all possible distinct tree graphs. Then, the following simple scaling relation is
established among the cumulants (cid:20)
:
n
(cid:20)
= S
(cid:20)
;
()
m
m
(cid:17)
m
where (cid:17)
= m (cid:0) and where the S
are constants (except for a weak scale dependence
m
m
introduced by the data windowing). On scales much larger than the correlation length of
the (cid:13)uctuation (cid:12)eld, the scaling relation holds for general random (cid:12)elds at the lowest non-
trivial order in the variance. It is interesting that it seems to give a fairly good description
of data down to quite small scales, perhaps even in the strongly non-linear regime (although
possibly with a di(cid:11)erent set of S
's). On the other hand, higher-order perturbation theory
m
in gravitational clustering (along with assumption of initial Gaussianity) leads to a predic-
tion of the constants S
in the small variance regime, i.e. at large scales (Peebles ,
m
Juszkiewicz & Bouchet , Bernardeau , ). After smoothing, the constants S
m
can be estimated as function of the linear power spectrum. Their value can depend also on
the bias mechanism (Fry & Gazta~naga ), and marginally on the redshift space distortion
(Fry & Gazta~naga ).
Observationally, the scaling relation describes the data of several surveys from a few
megaparsecs to h
Mpc and beyond (Saunders et al. ; Loveday et al. ; Bouchet
(cid:0)
et al. ; Gazta~naga , ). Values for S
up to m = are available in literature.
m
At large scales they are consistent with the gravitational perturbation theory; the errorbars
however are quite large, especially for m > . In N -body simulations, the agreement with
the gravitational perturbation theory is very good on large scales, while on small scales the
e(cid:11)ects of (cid:12)nite volume and of discreteness make the results more controversial (e.g. Lahav
, Colombi et al. ).
We determined the second, third and fourth order cumulant for our models, in real
space. We compared the variance/scale relation, the skewness/variance relation and the
kurtosis/variance relation with the above mentioned observational data, in particular with
100
100
10
10
1
1
0.1
0.1
0.01
0.01
1000
1000
100
100
10
10
1
1
0.1
0.1
0.01
0.01
BSI model
BSI model
Stromlo-APM
Stromlo-APM
1
1
10
10
100
100
G94 (APM)
G94 (APM)
0.001
0.001
10
10
1
1
variance
variance
0.1
0.1
1000
1000
100
100
10
10
1
1
0.1
0.1
0.01
0.01
0.001
0.001
100
100
10
10
G94 (APM)
G94 (APM)
10
10
1
1
variance
variance
0.1
0.1
1
1
100
100
10
10
1
1
variance
variance
0.1
0.1
Figure : Higher order moments for the BSI simulation with threshold (cid:23) (cid:25) in real space.
Open circles represent data from the h
Mpc box, (cid:12)lled circles from the h
Mpc
(cid:0)
(cid:0)
box. Clockwise from top left: variance vs. scale, compared with the Stromlo-APM data of
Loveday et al. ( ) (in this plot and in the following one the errors on our data are smaller
than the symbol size); skewness vs. variance, compared with the APM data of Gazta~naga
( , G ; dashed line); kurtosis vs. variance, compared with the results from G which
encompass scales from a few Mpc to h
Mpc; and scaling coe(cid:14)cients S
(triangles) and
(cid:0)
S
(squares) vs. variance (open symbols for the small box, (cid:12)lled symbols for the large box).
Figure : Scaling coe(cid:14)cients S
(triangles) and S
(squares) vs. variance relative to al l par-
ticles in a h
Mpc box, compared to the S
; S
expected from gravitational perturbation
(cid:0)
theory (starred symbols).
the Stromlo-APM data of Loveday et al. ( ) and the APM analysis of Gazta~naga
(G ). The results are plotted in Fig. . The errors represent the scatter between eight sets of
(cid:0)
(cid:0)
randomly placed cells. The open circles refer to the simulation box of h
Mpc
(selection at (cid:23) = :), the (cid:12)lled circles to the h
Mpc box ((cid:23) = :). The results
(cid:0)
merge one onto the other, from to h
Mpc, and correctly reproduce the observations.
(cid:0)
However, the hierarchical coe(cid:14)cients S
; S
for the small box decline below the ones from
the large box on the larger scales. This is probably a manifestation of the boundary e(cid:11)ects:
the moments estimated from a simulation with a (cid:12)xed number of particles are systematically
smaller than the moments of the parent distribution (Colombi et al. ). It is interesting
to observe that the scaling constants seem to be arranged in a plateau on small (non-linear)
scales, while they show a continuous decrease on large scales. The decrease is due to the
fact that the e(cid:11)ective slope of the power spectrum increases for large scales. Indeed, in Fig.
we plotted the expected S
; S
for top-hat cells from gravitational perturbation theory
(Bernardeau , scaling the spherical cells to our cubic cells) and compared with the
results from the matter distribution (i.e., without bias) in the h
Mpc box, on scales
(cid:0)
from to h
Mpc. The agreement with the simulation results is reasonable in the small-
(cid:0)
variance regime; the residual discrepancy is maybe due to the conversion to cubic cells, and
to (cid:12)nite-volume e(cid:11)ects.
Conclusions
This paper reports on the linear and non-linear results of an extensive set of N -body sim-
ulations with broken scale invariance power spectrum. To our knowledge, this is the (cid:12)rst
N -body simulation with a primordial power spectrum arising from a double in(cid:13)ationary
model. The results make us believe that this broken scale invariance model is an interesting
addition to the list of models that help in reconciling CDM with observations. We compared
our simulations with various observations, such as spatial and angular correlation functions,
power spectra, higher-order moments, and we found reasonable agreement. To extend the
dynamical range, we joined the results from simulation boxes ranging from h
Mpc to
(cid:0)
(cid:0)
h
Mpc. This allowed us to (cid:12)t, for instance, the observed power spectrum over almost
two decades in wavelenght, and we could determine the scaling coe(cid:14)cients S
; S
over three
decades in the variance. The replication method we used to obtain a APM-like angular pro-
jection has been successfully tested via the limiting magnitude shifting (although we found
the method unsatisfactory for (cid:18) (cid:21)
).
The BSI power spectrum we investigated contains, in addition to the overall normaliza-
tion, two additional parameters, the height and the location of the break.
In the double
in(cid:13)ation model, the break height is connected with the ratio of the two mass parameters
characterizing the two in(cid:13)ationary stages. The break scale is sensibly dependent on the
initial value of the scalar (cid:12)eld. In Gottl(cid:127)ober, M(cid:127)uller, Starobinsky ( ) it was argued that
this initial value may arise naturally as a quantum (cid:13)uctuation. Clearly this requires fur-
ther elaboration. Here our aim was to show that the values of the two parameters which
were derived from linear analysis lead to non-linear clustering properties in accordance with
observations.
We are grateful to Michael Vogeley for providing us with the data sets of the CfA power
ACKNOWLEDGMENTS
spectrum and to Steve Maddox for the APM angular correlation function. Ron Kates and
J(cid:127)org Retzla(cid:11) helped us in preparing the N -body simulations. L.A. wishes to thank the
Astrophysical Institute of Potsdam for the friendly hospitality during the early stages of
preparation of this work, and St(cid:19)ephane Colombi for very fruitful discussions. The work of
L.A. at Fermilab has been supported by DOE and NASA under grant NAGW-. L.A. also
acknowledges CNR (Italy) for (cid:12)nancial support. S.G. acknowledges support from Fermilab
for his stay at Batavia, where the (cid:12)nal version of this paper was prepared.
References
Bernardeau F. Ap. J. ,
Bernardeau F. , preprint CITA /
Bouchet F.R., Strauss M.A., Davis M., Fisher K.B., Yahil A., & Huchra J.P. , Ap. J.
,
Bond, J.R., Efstathiou, G., , Ap. J. , L.
Catelan P., Coles P., Matarrese S. & Moscardini L. Mon. Not. R. Ast. Soc. ,
Coles P. et al. Mon. Not. R. Ast. Soc. ,
Coles P. & Frenk C.S. Mon. Not. R. Ast. Soc.
Coles P. & Jones B. Mon. Not. R. Ast. Soc.
Colombi S., Bouchet F.R., & Schae(cid:11)er R., , Astron. Astrophys. ,
Cramer H. , Mathematical Methods of Statistics (Princeton: Princeton Univ. Press)
Efstathiou, G., Kaiser, N., Saunders, W., Lawrence, A., Rowan-Robinson, M., Ellis, R.S.,
and Frenk, C.S., , Mon. Not. R. Ast. Soc. , P
Fisher, K., Davis, M., Strauss, M., Yahil, A. Huchra, J.P., , Ap. J. , .
Fry J.N. & Gazta~naga E. Ap. J. , ,
Fry J.N. & Gazta~naga E. Ap. J. , ,
Gazta~naga E. , Ap. J. , , L
Gazta~naga E. , Mon. Not. R. Ast. Soc. ,
G(cid:19)orski, K.M., Hinshaw, G., Banday,A.J., Bennett, C.L., Wright, E.L., Kogut, A., Smoot,G.F.,&
Lubin, P. , Ap. J. , L
Gottl(cid:127)ober, S., M(cid:127)ucket, J.P., , Astron. Astrophys. , .
Gottl(cid:127)ober, S., M(cid:127)uller, V., Starobinsky, A.A., , Phys. Rev. D, .
Gottl(cid:127)ober, S., M(cid:127)ucket, J.P., Starobinsky, A.A., , Ap. J. October.
Groth E.J. & Peebles P.J.E. Ap. J.
Juszkiewicz R., & Bouchet F.R. , in Proc. of the nd DAEC Meeting on the Distribution
of Matter in the Universe, eds. G.A. Mamon, & D. Gerbal,
Juszkiewicz R., Weinberg D. H., Amsterdamski P., Chodorowski M. & Bouchet F. ,
preprint (IANSS-AST / )
Kaiser N. Ap. J. L
Kates, R., Klypin, A., Kotok, N., Astron. Astrophys. , .
Kates, R., Gottl(cid:127)ober, S., M(cid:127)ucket, J.P., M(cid:127)uller, V. Retzla(cid:11), J., , in preparation
Kofman L. & Bernardeau, F. , preprint IFA- -
Lahav O., Itoh M, Inagaki S. & Suto Y. , Ap. J. ,
Loveday J., Efstathiou G., Peterson B. A. & Maddox S.J. , Ap. J. , , L
Maddox, S.J., Efstathiou, G., Sutherland, W.J., and Loveday, J., , Mon. Not. R. Ast.
Soc. , P.
Moscardini L. et al. Ap. J. , L
Peacock J. A. Mon. Not. R. Ast. Soc. , P
Peebles, P.J.E., , The Large-Scale Structure of the Universe (Princeton, Princeton Uni-
versity Press)
Saunders W., Frenk C., Rowan-Robinson M., Efstathiou G., Lawrence A., Kaiser N., Ellis
R., Crawford J., Xia X.-Y., Parry I. , Nature , ,
Smoot G.F., Bennett C.L., Kogut A., Wright E.L. et al. , , Ap. J. , L.
Vogeley, M., Park, C., Geller, M. Huchra, J.P., , Ap. J. , L.
Wright, E.L., Smoot, G.F., Bennett, C.L., Lubin, P.M. , Ap. J. , ,.
This figure "fig1-1.png" is available in "png"(cid:10) format from:
http://arxiv.org/ps/astro-ph/9408104v1
|
astro-ph/9905348 | 2 | 9905 | 2000-01-12T20:00:07 | Secular Evolution of Hierarchical Triple Star Systems | [
"astro-ph"
] | We derive octupole-level secular perturbation equations for hierarchical triple systems, using classical Hamiltonian perturbation techniques. By extending previous work done to leading (quadrupole) order to octupole level (i.e., including terms of order alpha^3, where alpha = a_1/a_2 < 1 is the ratio of semimajor axes) we obtain expressions that are applicable to a much wider range of parameters. For triple systems containing a close inner binary, we also discuss the possible interaction between the classical Newtonian perturbations and the general relativistic precession of the inner orbit. In some cases we show that this interaction can lead to resonances and a significant increase in the maximum amplitude of eccentricity perturbations. We establish the validity of our analytic expressions by providing detailed comparisons with the results of direct numerical integrations of the three-body problem obtained for a large number of representative cases. We also discuss applications of the theory in the context of several observed triple systems of current interest, including the millisecond pulsar PSR B1620-26 in M4, the giant planet in 16 Cygni, and the protostellar binary TMR-1. [short version] | astro-ph | astro-ph | Secular Evolution of Hierarchical Triple Star Systems
Eric B. Ford1, Boris Kozinsky2, & Frederic A. Rasio3
Physics Department, Massachusetts Institute of Technology Cambridge, MA 02139
ABSTRACT
We derive octupole-level secular perturbation equations for hierarchical triple
systems, using classical Hamiltonian perturbation techniques. Our equations describe
the secular evolution of the orbital eccentricities and inclinations over timescales
long compared to the orbital periods. By extending previous work done to leading
(quadrupole) order to octupole level (i.e., including terms of order α3, where
α ≡ a1/a2 < 1 is the ratio of semimajor axes) we obtain expressions that are applicable
to a much wider range of parameters. In particular, our results can be applied to
high-inclination as well as coplanar systems, and our expressions are valid for almost all
mass ratios for which the system is in a stable hierarchical configuration. In contrast,
the standard quadrupole-level theory of Kozai gives a vanishing result in the limit of
zero relative inclination. The classical planetary perturbation theory, while valid to
all orders in α, applies only to orbits of low-mass objects orbiting a common central
mass, with low eccentricities and low relative inclination. For triple systems containing
a close inner binary, we also discuss the possible interaction between the classical
Newtonian perturbations and the general relativistic precession of the inner orbit.
In some cases we show that this interaction can lead to resonances and a significant
increase in the maximum amplitude of eccentricity perturbations. We establish the
validity of our analytic expressions by providing detailed comparisons with the results
of direct numerical integrations of the three-body problem obtained for a large number
of representative cases. In addition, we show that our expressions reduce correctly to
previously published analytic results obtained in various limiting regimes. We also
discuss applications of the theory in the context of several observed triple systems
of current interest, including the millisecond pulsar PSR B1620−26 in M4, the giant
planet in 16 Cygni, and the protostellar binary TMR-1.
0
0
0
2
n
a
J
2
1
2
v
8
4
3
5
0
9
9
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
Subject headings: celestial mechanics, stellar dynamics -- planetary systems --
binaries: general -- planets and satellites: general
1Present address: Princeton University, Department of Astrophysical Sciences, Peyton Hall, NJ 08544;
[email protected]
[email protected]
3Alfred P. Sloan Research Fellow; [email protected]
-- 2 --
1.
Introduction
About one third of all binary star systems are thought to be members of larger multiple
systems. Most of these are hierarchical triples, in which the (inner) binary is orbited by a third
body in a much wider orbit (see Tokovinin 1997a,b for recent results and compilations). Secular
perturbations in triples result from the gravitational interaction between the inner binary and the
outer object, possibly coupled to other processes such as stellar evolution, tidal effects, or, for
compact objects, general relativistic effects. In strongly hierarchical triples, the two orbits never
approach each other closely, and an analytic, perturbative approach can be used to calculate
the evolution of the system. One particularly important perturbation is that of the orbital
eccentricities. As the two orbits torque each other and exchange angular momentum, their
eccentricities will undergo periodic oscillations over secular timescales (i.e., very long compared
to the orbital periods). For non-coplanar systems, corresponding oscillations occur in the orbital
inclinations. In contrast, according to canonical perturbation theory, there is no secular change in
the semimajor axes, since the energy exchange between the two orbits averages out to zero over
long timescales (see, e.g., Heggie 1975).
For triple systems that begin their life near the stability limit, the result of an eccentricity
increase can be catastrophic, leading to a collision between the two inner stars, if they started as
a close pair, or, more typically, to the disintegration of the triple. This disintegration proceeds
through a phase of chaotic evolution whose outcome is the ejection of one of the three stars
(typically the least massive body) on an unbound trajectory, while the other two are left in a more
tightly bound binary. A striking example of this process was revealed by the recent HST/NICMOS
observations of the TMR-1 system in the Taurus star-forming region (Terebey et al. 1998). The
HST images reveal a faint companion, most likely a giant planet or brown dwarf, that appears to
have been ejected from its parent protostellar binary system. More indirect observational evidence
is provided in the form of binary systems with anomalously high space velocities. In particular,
the disintegration of short-lived triples formed in dense star forming regions may lead to binary
OB runaway stars with very large peculiar velocities, such as HD 3950 (Gies & Bolton 1986).
The stability of triple systems has been the subject of many theoretical studies. Most
recently, Eggleton & Kiseleva (1995 and references therein) performed numerical experiments and
provided an empirical stability criterion in terms of a critical ratio Ymin between the periastron
distance of the outer orbit to the apastron distance of the inner orbit. For systems containing
three nearly equal masses, one finds Ymin ≃ 3 − 6 depending on initial phases, eccentricities and
inclinations. Holman & Wiegert (1999) study the stability of planets in binary systems, both for
planets orbiting close to one of the two stars, and for planets orbiting outside the binary. All these
stability analyses are based on numerical integrations of the three-body problem that are limited
to 104 − 106 periods of the outer orbit. In some cases, however, the secular evolution timescale of
the triple can be much longer than this, and therefore systems that remain stable for the duration
of the numerical integration may in fact turn out to be unstable over secular timescales. Analytic
results such as those derived here can therefore help determining more accurate stability criteria,
-- 3 --
e.g., by integrating the secular evolution equations and verifying that the stability ratio remains
> Ymin over the entire cycle of secular perturbations.
Hierarchical triple star systems can play an important role in the dynamical evolution of
dense star clusters containing primordial binaries. The cores of globular clusters, for example, are
thought to contain a small but dynamically significant population of triple systems formed through
dynamical interactions between primordial binaries (McMillan, Hut, & Makino 1991). Both stable
and unstable triples can form easily through exchange and resonant interactions between binaries.
In direct N -body integrations of the cluster dynamics, marginally stable or unstable triples can
represent a significant computational bottleneck, since they require very long integrations of the
orbital dynamics in order to resolve the outcome of the interaction (see, e.g., Mikkola 1997).
Direct observational evidence for the dynamical production of triple systems in globular
clusters is provided by the millisecond pulsar system PSR B1620−26 (Rasio, McMillan, & Hut
1995; Ford et al. 2000). This radio pulsar is a member of a hierarchical triple system located in the
core of the globular cluster M4. The inner binary of the triple contains the ≃ 1.4M⊙ neutron star
with a ≃ 0.3M⊙ white-dwarf companion in a 191-day orbit (Lyne et al. 1988; McKenna & Lyne
1988). The triple nature of the system was first proposed by Backer (1993) in order to explain the
unusually high residual second and third pulse frequency derivatives left over after subtracting a
standard Keplerian model for the pulsar binary. The pulsar has now been timed for eleven years
since its discovery (see Thorsett et al. 1999 for the most recent update). These observations have
not only confirmed the triple nature of the sytem, but they have also provided tight constraints
on the mass and orbital parameters of the second companion. Theoretical modeling of the latest
timing data (now including five pulse frequency derivatives) and preliminary measurements of
the orbital perturbations of the inner binary have further constrained the mass of the second
companion, and strongly suggest that it is a giant planet or a brown dwarf of mass ∼ 0.01 M⊙ at
a distance of ∼ 50 AU from the pulsar binary (Joshi & Rasio 1997; Ford et al. 2000).
Our treatment of the secular perturbations in this paper is based on classical celestial
mechanics techniques and assumes that all three bodies are unevolving point masses. Whenever
the stellar evolution time of one of the components becomes comparable to any of the orbital
perturbation timescales computed here, the evolution of the triple can be affected significantly
through mass losses, or mass transfer. For a recent discussion of stellar evolution in triples, see
Iben & Tutukov (1999), who study the production of Type Ia supernovae from the mergers of
heavy white dwarfs inside hierarchical triples. Mikkola & Tanikawa (1998) have studied the
episodic mass transfer triggered by large eccentricity oscillations of the inner binary in the secular
evolution of the triple system CH Cygni. Eggleton & Verbunt (1989) have discussed the possible
relevance of triple star evolution for the formation of low-mass X-ray binaries.
Our work focuses on triple systems containing well-separated components, in which the
orbital perturbation timescales are short compared to any tidal dissipation time. For triple
systems containing a close inner binary, tidal dissipation in the inner components provides a sink
-- 4 --
of energy and angular momentum which can change substantially the character of the secular
perturbations. For a recent discussion of tidal dissipation in triple systems containing a close inner
binary, see the paper by Kiseleva, Eggleton, & Mikkola (1998). Bailyn & Grindlay (1987) have
discussed the combined effects of tidal interaction and mass transfer for compact X-ray binaries
in hierarchical triples. Mazeh & Shaham (1979) were the first to point out that the combination
of tidal dissipation and secular eccentricity perturbations in triples could sometimes lead to a
substantial orbital shrinking of the inner binary.
One possible additional perturbation effect that we do take into account in this work is the
general relativistic precession of the inner orbit if the inner binary contains compact objects. An
example is provided by the PSR B1620−26 triple system, in which the inner binary contains a
neutron star and a white dwarf. When the precession periods from general relativity and from
Newtonian perturbations in the triple become comparable, a type of resonant effect is possible
which leads to increased magnitudes for the orbital perturbations. A similar resonant effect has
been mentioned by Soderhjelm (1984) for triples where the inner binary precesses under the
influence of a rotationally-induced quadrupole moment in one of the stars.
Our paper is organized as follows.
In §2 we present a derivation of the octupole-order
secular perturbation equations and we compare our results to those obtained in the quadrupole
approximation and in classical planetary perturbation theory. In §3 the analytic results are
compared to direct numerical integrations, and the effects of varying all relevant parameters are
explored. In §4 we discuss the effects of the general relativistic precession of the inner orbit on
the secular evolution of the triple, using the PSR B1620−26 system as an example. In §5 possible
applications of our results to other observed triple systems are briefly discussed.
2. Analytic Secular Perturbation Theory
In this section we present a simple analytic treatment of the long-term, secular evolution of
hierarchical triple systems using time-independent Hamiltonian perturbation theory in which the
small parameter is the ratio of semimajor axes. We discuss essential aspects of the lowest-order
(quadrupole-level) approximation, which has been widely used to study hierarchical stellar triples.
Then we extend the approximation to the octupole level and compare our results with results from
the quadrupole and other approximations to show that the octupole-level equations derived here
are valid for a far greater range of parameters.
2.1. Summary of Previous Work
Our derivation of the octupole-level secular perturbation equations is based on classical
perturbation methods of celestial mechanics. Studies of the long-term behavior of the solar
system led Lagrange and Laplace to the creation of the first classical perturbation theory. Their
-- 5 --
approach is applicable to a small class of planetary configurations with parameters similar to
those of the solar system. The lunar problem was successfully attacked in the end of the last
century by Delaunay who was the first to apply the method of canonical transformations to
long-term perturbations. This method possesses much greater generality and was used to study
a broad spectrum of problems. Brown (1936) was the first to apply canonical averaging to
stellar triples, and he obtained the transformed quadrupole Hamiltonian. Kozai (1962) made use
of the quadrupole approximation while studying the long-term motion of asteroids and noted
several important properties of this approximation. Harrington (1968) obtained quadrupole-level
expressions similar to Kozai's for general hierarchical systems of three stars. Soderhjelm (1984)
derived octupole-level equations in the limit of low eccentricities and inclinations. In particular, he
demonstrated that the quadrupole approximation fails in this regime because the octupole term in
the Hamiltonian becomes dominant. Finally, Marchal (1990) averaged the octupole Hamiltonian
keeping all terms up to third order in α and some terms of order α7/2. His Hamiltonian truncated
at third order is identical to the one used in this paper.
In the process of completing this work, we became aware of related ongoing work by other
groups. In particular, Krymolowski & Mazeh (1999) have derived octupole-order perturbation
equations following the same method used here. They retain some additional terms, of order
α7/2, which were also partly included in Marchal's (1990) Hamiltonian. Based on a few numerical
integrations that Krymolowski & Mazeh (1999) provide for a fairly strongly coupled system
(α = 0.1), it appears that these higher-order terms have a negligible effect on the perturbations,
although they can lead to slightly shorter periods of eccentricity oscillations for systems with low
relative inclination. Eggleton (2000) has used a perturbation method based on the variation of the
Runge-Lenz vector (see Heggie & Rasio 1996) to derive an extension of Kozai's theory to octupole
order. Similar work has been done by Georgakarakos (2000), who concentrates on systems where
the inner orbit is nearly circular.
2.2. Octupole Theory
A hierarchical triple system consists of a close binary (m0 and m1) and a third body (m2)
moving around the inner binary on a much wider orbit. To describe this structure it is convenient
to use Jacobi coordinates, which are defined as follows. The vector r1 represents the position of
m1 relative to m0, and r2 is the position of m2 relative to the center of mass of the inner binary
(See Fig. 1). This coordinate system naturally divides the motion of the triple into two separate
motions, and makes it possible to write the Hamiltonian as a sum of two terms representing the
two decoupled motions and an infinite series representing the coupling of the orbits. Let the
subscripts 1 and 2 refer to the inner and outer orbits, respectively. The coupling term is written as
a power series in the ratio of the semi-major axes α ≡ a1/a2, which serves as the small parameter
in our perturbation expansion. The complete Hamiltonian of the three-body system is given by
-- 6 --
(Harrington 1968)
k2m0m1
2a1
+
F =
k2(m0 + m1)m2
2a2
+
k2
a2
∞
Xj=2
αjMj(cid:18) r1
a1(cid:19)j(cid:18) a2
r2(cid:19)j+1
Pj(cos Φ),
(1)
where k2 is the gravitational constant, Pj's are the Legendre polynomials, Φ is the angle between
r1 and r2, and
Mj = m0m1m2
mj−1
0 − (−m1)j−1
(m0 + m1)j
.
(2)
We shall deal with the expansion only up to third order in α.
Let us define a set of canonical variables, known as Delaunay's elements, that provide a
particularly convenient dynamical description of our three-body system. The angle variables are
chosen to be
l1, l2 = mean anomalies
g1, g2 = arguments of periastron
h1, h2 = longitudes of ascending nodes
and their conjugate momenta
L1 =
m0m1
m0 + m1qk2(m0 + m1)a1
L2 =
G1 = L1q1 − e2
1
m2(m0 + m1)
m0 + m1 + m2qk2(m0 + m1 + m2)a2,
G2 = L2q1 − e2
2,
H1 = G1 cos i1
H2 = G2 cos i2,
where e1, e2 are the orbital eccentricities and i1, i2 are the orbital inclinations.
The usual canonical relations represent the equations of motion:
dLj
dt
=
∂F
∂lj
dlj
dt
= −
∂F
∂Lj
,
dGj
dt
=
∂F
∂gj
dgj
dt
= −
∂F
∂Gj
,
dHj
dt
=
∂F
∂hj
dhj
dt
= −
∂F
∂Hj
,
(3)
(4)
(5)
(6)
(7)
(8)
(9)
(10)
(11)
where j = 1, 2. Note that H2 is the z-component of the angular momentum contributed by the
perturbing body, the z-axis being the direction of the total angular momentum, perpendicular to
-- 7 --
m2
r2
m1
Φ
r1
m0
Fig. 1. -- Diagram illustrating the coordinate system used to describe the hierarchical triple system.
-- 8 --
the invariable plane of the system (See Fig. 2). Equations (9) -- (11) appear to have six degrees of
freedom, but they can be reduced to four by the theorem of elimination of nodes (Jeffrys & Moser
1966). The Hamiltonian contains h1 and h2 only in the combination h1 − h2, and it is symmetric
with respect to the orientation of the line of nodes when the invariable plane is chosen as a
reference plane. In other words, the Hamiltonian is symmetric with respect to rotations about the
total angular momentum vector H. Thus, H1 and H2 enter the Hamiltonian only as H1 + H2 = H
and can be eliminated from the Hamiltonian using the relations
H1 =
H2 =
H 2 + G2
2H
H 2 + G2
2H
2
1 − G2
2 − G2
1
(12)
(13)
Using the new canonical elements we can write the first four terms of the Hamiltonian (1) as
Foct = F0 + F1 + F2 + F3
=
β0
2L2
1
+
β1
2L2
2
1
L6
+ 8β2 L4
2!(cid:18) r1
2!(cid:18) r1
a1(cid:19)3(cid:18) a2
r2(cid:19)4
a1(cid:19)2(cid:18) a2
r2(cid:19)3
(5 cos3 Φ − 3 cos Φ),
(3 cos2 Φ − 1)
+2β3 L6
1
L8
where the mass parameters are
,
β0 = k4 (m0m1)3
m0 + m1
β1 = k4 (m0 + m1)3m3
2
m0 + m1 + m2
(m0 + m1)7
,
β2 =
β3 =
k4
16
k4
4
(m0 + m1 + m2)3
(m0 + m1)9
(m0 + m1 + m2)4
m7
2
(m0m1)3 ,
m9
2(m0 − m1)
(m0m1)5
.
(14)
(15)
(16)
(17)
(18)
(19)
Each term in the series is labeled according to the degree of the Legendre polynomial associated
with it (same as the power of r1/r2). Following the standard nomenclature associated with
multipole expansions we shall call the Hamiltonian containing the first three terms (up to
j = 2) "quadrupole" and the third-order Hamiltonian (15) "octupole." The first two terms
in expression (15) describe the unperturbed motion of the inner and outer binaries, and the
higher-order terms describe the coupling. The quadrupole Hamiltonian contains the perturbation
of order α2, the octupole Hamiltonian extends this to order α3.
The complete Hamiltonian (1) contains the full description of the system. However, we
are going to restrict our study to the long-term, secular behavior of the system, by averaging
over short-period effects. Even though equation (15) is already an approximation of the full
-- 9 --
H
H2
G2
G1
H1
i
2. -- Diagram illustrating the relationships between the canonical variables and angular
Fig.
momenta.
invariable plane
-- 10 --
Hamiltonian, it contains information about short-period perturbations that needs to be eliminated.
In particular the angle Φ depends on the mean anomalies. Further simplification is achieved
through a canonical transformation of variables, called the von Zeipel transformation. Its essence
is to replace the Delaunay elements with a set of new canonical coordinates and momenta that rid
the Hamiltonian of the dependence on l1 and l2. The perturbed action variables are still periodic
functions of the perturbed angle variables, but the former are no longer linear functions of time.
The goal is to find such a set of action-angle variables that the perturbed Hamiltonian will be a
function only of the action variables. In the end we can think of the Hamiltonian as describing the
interaction between two weighted elliptical rings instead of point masses in orbits.
It is important to note that we do not simply average the Hamiltonian with respect to these
variables, since this would destroy the canonical structure of the equations of motion, but instead
proceed in a more cautious and intricate manner. We start by requiring that the new Hamiltonian
be equal to the old one, since changing variables does not change the energy, and expanding both
sides of the equality as Taylor series in α. Then we go order by order to identify the terms in
the transformed Hamiltonian, using the result of the previous order calculation in each step. The
theory behind the Von Zeipel method is very well presented and illustrated by Goldstein (1980,
section 11-5) and Hagihara (1972). Additionally, Harrington (1968, 1969) has applied this method
to the quadrupole Hamiltonian. We followed exactly the same prescription but all the way to
third order. Here we present only the results of the von Zeipel averaging procedure, omitting the
laborious algebraic details.
Let us define the following convenient quantities
θ = cos i =
H 2 − G2
2G1G2
1 − G2
2
,
(20)
where H = G1 + G2 and H = H is given by initial conditions and i = i1 − i2 is the mutual
inclination. The angle ϕ between the directions of periastron is given by
The doubly-averaged Hamiltonian is given by
cos ϕ = − cos g1 cos g2 − θ sin g1 sin g2.
¯Foct = C2 · {(2 + 3e2
+C3 · e1e2nA cos ϕ + 10θ(1 − θ2)(1 − e2
1) sin g1 sin g2o,
1)(3θ2 − 1) + 15e2
1(1 − θ2) cos 2g1}
where
C2 =
k4
16
(m0 + m1)7
m7
2
(m0 + m1 + m2)3
(m0m1)3
L4
1
2G3
2
L3
,
C3 =
15
16
k4
4
(m0 + m1)9
(m0 + m1 + m2)4
m9
2(m0 − m1)
(m0m1)5
L6
1
2G5
2
L3
,
(21)
(22)
(23)
(24)
-- 11 --
B = 2 + 5e2
1 − 7e2
1 cos 2g1,
A = 4 + 3e2
1 −
5
2
(1 − θ2)B.
(25)
(26)
Note that the Hamiltonian (22) does not contain any dependence on l1 or l2 because these
variables have been integrated out as a result of the canonical transformation. The actual
differences between the original and the transformed variables are small (of order α or smaller)
and periodic. Variables appearing in equation (22) are only approximations of the old variables
defined in equations (5) -- (8), and they can be thought of as the averages of the old variables. This
Hamiltonian is equivalent to the Hamiltonian given by Marchal (1990; see his eqs. [252] -- [255]).
The absence of l1 and l2 from the transformed Hamiltonian implies that L1 and L2 are
constants of the motion, which in turns implies that, in our approximation, the transformed
semi-major axes a1 and a2 are constant. Thus, the only secularly changing parameters in this
model are e1, e2, g1, g2, and i, which is coupled to e1 and e2 by relation (20). The equations of
motion are derived from the Hamiltonian (22) using the canonical relations
dei
dt
=
∂ei
∂Gi
∂ ¯Foct
∂gi
,
dgi
dt
∂ ¯Foct
∂Gi
.
= −
After regrouping terms, the octupole-level secular perturbation equations follow:
(27)
(28)
(29)
(30)
and
dg2
dt
dg1
dt
de1
dt
= C2 · 6(cid:26) 1
θ
1(3 − 5 cos 2g1)i(cid:27)
1)o − 5θB cos ϕi
1
θ
+
G2
G1 h4θ2 + (5 cos 2g1 − 1)(1 − e2
1 − θ2)i +
G2 h2 + e2
G1(cid:19)h sin g1 sin g2nA + 10(3θ2 − 1)(1 − e2
−C3 · e2(e1(cid:18) 1
1 − e2
e1G1 h sin g1 sin g2 · 10θ(1 − θ2)(1 − 3e2
−
1 − e2
G1 n30e1(1 − θ2) sin 2g1o + C3 · e2
1) + cos ϕ(3A − 10θ2 + 2)i),
1 − e2
G1 n35 cos ϕ(1 − θ2)e2
1)(1 − θ2) cos g1 sin g2 − A(sin g1 cos g2 − θ cos g1 sin g2)o,
1(3 − 5 cos 2g1)i +
G1 h2 + e2
−10θ(1 − e2
1
1
1
= C2 ·
1 sin 2g1
= C2 · 3(cid:26) 2θ
+C3 · e1( sin g1 sin g2" 4e2
+ cos ϕ"5Bθe2(cid:18) 1
G1
+
θ
G2(cid:19) +
2 + 1
e2G2
G2 h4 + 6e2
10θ(1 − θ2)(1 − e2
4e2
2 + 1
e2G2
A#),
1 + (5θ2 − 3)(2 + e2
1) − e2(cid:18) 1
θ
1(3 − 5 cos 2g1))i(cid:27)
1))#
G2(cid:19) (A + 10(3θ2 − 1)(1 − e2
G1
+
(31)
-- 12 --
de2
dt
= −C3 · e1
2
1 − e2
G2 n10θ(1 − θ2)(1 − e2
1) sin g1 cos g2 + A(cos g1 sin g2 − θ sin g1 cos g2)o.(32)
This system of coupled nonlinear differential equations describes the octupole-level behavior of
hierarchical triples and can be easily integrated numerically to determine the secular evolution
of a hierarchical triple for any initial configuration. We found that for calculations with small
eccentricities it is better to use the transformed set of variables e1 sin g1, e1 cos g1, e2 sin g2, and
e2 cos g2. These equations can be integrated numerically much more rapidly than the full equations
of motion.
2.3. Comparison with Quadrupole-Level Results
The quadrupole results of Kozai (1962) and Harrington (1969) have found numerous
applications in recent studies of planetary, pulsar, and stellar systems. Mazeh & Shaham (1979)
calculated the quadrupole-level, long-term periodic behavior of triples and used their results to
study high-amplitude eccentricity modulations of systems with large initial inclination. Holman
et al. (1997) used the quadrupole-level theory of Kozai to analyze the dynamical evolution of the
planet in the binary star 16 Cygni. Rasio, Ford, & Kozinsky (1997) used the same theory to study
long-term eccentricity perturbations in the PSR B1620−26 pulsar system.
The quadrupole Hamiltonian can be obtained from expression (22) by dropping the C3 term:
Fq = C2 · {(2 + 3e2
1)(3θ2 − 1) + 15e2
1(1 − θ2) cos 2g1}.
(33)
Note that Fq is independent of g2, meaning that, in the quadrupole approximation, G2 is a
constant of the motion, and so is e2 by equation (27). Therefore, to quadrupole order in secular
perturbation theory, there is no variation in the eccentricity of the outer orbit. This is a well-known
result (see, e.g., Marchal 1990, Sec. 10.2.3). Now there is only one degree of freedom left, so the
evolution of the inner eccentricity is described by
dg1
dt
with
= C2 · 6(cid:26) 1
G1 h4θ2 + (5 cos 2g1 − 1)(1 − e2
1 − θ2)i +
θ
G2 h2 + e2
1(3 − 5 cos 2g1)i(cid:27) ,
de1
dt
= C2 ·
1
1 − e2
G1 n30e1(1 − θ2) sin 2g1o.
(34)
(35)
The advantage of using the quadrupole approximation is that it can describe the secular
behavior of systems with high relative inclination and a wide range of initial eccentricities,
regimes not covered by the classical planetary perturbation theory. As in the classical theory,
the quadrupole-level perturbation equations (34) -- (35) can be solved exactly for the period and
amplitude of oscillation (Kozai 1962). The period of eccentricity oscillations is given approximately
by
-- 13 --
Pe ≃ P1(cid:18) m0 + m1
m2
a1(cid:19)3
(cid:19)(cid:18) a2
(cid:16)1 − e2
2(cid:17)3/2
,
(36)
where P1 is the orbital period of the inner binary (Mazeh & Shaham 1978). This expression
should be multiplied by a coefficient of order unity which can be obtained using Weierstrass's zeta
function as shown by Kozai (1962).
The secular evolution can be visualized with the help of phase-space diagrams of e1 vs cos g1.
An example is provided in Figure 3. Each contour corresponds to an initial condition with a
certain value of the total angular momentum H. Since G2 is fixed, e1 is coupled to i through
equation (20), so the relative inclination oscillates with the same period as e1. The up-down
symmetry forces similar behavior in g1-e1 phase space for both g1 ∈ [−π, 0] and g1 ∈ [0, π]. One
obvious feature is the existence of two regimes: libration and circulation. The libration island
generally appears when the initial inclination is greater than some critical value, which for most
systems is around icrit ≃ 40◦. Kozai (1962) calculated that for α ≤ 0.10, 38.960◦ ≤ icrit ≤ 39.231◦.
From the shape of the large libration island we see that e1 can grow from a very small initial value
to a very large maximum. Holman et al. (1997) approximate the maximum inner eccentricity as
where io is the initial relative inclination. Libration can occur for low inclinations as well, but this
does not lead to large eccentricity oscillations.
e1,max ≃ (cid:16)1 − (5/3) cos2 io(cid:17)1/2
,
(37)
Several erroneous features of the quadrupole approximation are worth noting. For an initial
condition with e1 = 0, no evolution occurs at all. This is especially significant in a case where there
is no libration island, since in that case the eccentricity perturbation would appear to approach
zero continuously as the initial eccentricity is decreased. Similarly, in the coplanar case (θ = 1),
the theory predicts no evolution of eccentricity. From the classical planetary perturbation theory
(§2.4), which assumes low eccentricities and inclinations, we know that this is not correct. These
features also contradict the octupole-level results (§2.2), as well as the results of direct numerical
integrations (§3.2). We conclude that the quadrupole approximation fails for low inclination and
for low inner eccentricity. However, it remains qualitatively applicable in the high-inclination
regime.
The octupole-level theory has more degrees of freedom and covers most regimes of hierarchical
triple configurations. The perturbation equations (29) -- (32) indicate that there are no additional
conserved quantities apart from the obvious ones (total angular momentum and total energy). In
contrast to the quadrupole theory, the quantities e2 and g2 now vary with time and the behavior
is much harder to visualize. We can notice striking qualitative differences between the two
theories by looking at phase-space diagrams. An example is provided in Figure 4. Trajectories
are no longer closed, and transitions between libration and circulation occur, since the angular
momentum of the outer orbit now evolves with time. Thus, we now have more than one frequency
in the secular oscillations.
-- 14 --
1
0.9
0.8
0.7
0.6
1
0.5
e
0.4
0.3
0.2
0.1
0
−1
−0.8
−0.6
−0.4
−0.2
0
cos(g1)
0.2
0.4
0.6
0.8
1
Fig. 3. -- Phase space trajectories obtained in the quadrupole approximation for a system with
m2/m1 = 10−3, m3/m1 = 1, α−1 = 100, e2 = 0.9, and initial values of e1 ranging from 0.02 to 0.9.
The libration contours were obtained by setting the initial value of g1 = 90◦.
-- 15 --
1
0.9
0.8
0.7
0.6
1
0.5
e
0.4
0.3
0.2
0.1
0
−1
−0.8
−0.6
−0.4
−0.2
0
cos(g1)
0.2
0.4
0.6
0.8
1
Fig. 4. -- Phase-space trajectory of the same system as in Figure 3 with initial e1 = 0.02, but with
the octupole terms included in the integration. (Note that this corresponds to a single trajectory
in Fig. 3.)
-- 16 --
Figure 5 demonstrates that a system can have a qualitatively different behavior from what
is expected in the quadrupole approximation. While the quadrupole theory predicts periodic
variations of constant amplitude, according to the octupole equations (and in agreement with the
results of a direct numerical integration) the amplitude grows very close to unity. This leads to a
very small periastron distance and the possibility of a tidal interaction or collision between the two
inner stars. Thus, one must exercise great caution when modeling systems using the quadrupole
approximation. Ignoring octupole-level terms can lead to completely invalid results.
2.4. Comparison with Classical Planetary Perturbation Theory
For application to many problems in the context of the Solar System, a classical perturbation
theory was developed many years ago that applies to low-eccentricity, low-inclination orbits of
planets around a central star (one dominant mass). This theory does not assume that the ratio
of semimajor axes is small, and therefore it provides results valid to all orders in α. A detailed
account of the planetary theory can be found in Brouwer & Clemence (1961, Chap. 16). For
an excellent pedagogic summary, see Dermott & Nicholson (1986). Rasio (1994, 1995) used the
classical theory to study the eccentricity perturbations in the PSR B1620-26 triple system in the
limit of coplanar orbits, and derived simple approximate expressions for the period and amplitude
of eccentricity oscillations in various limits.
Here we will only consider the variations of the eccentricities, since the results for inclinations
are very similar. It turns out that the inclination evolution is decoupled from the eccentricity
evolution, and so the two can be solved separately. Because eccentricities are very small and can
vanish, it is better to use the variables
h1 = e1 sin g1
h2 = e2 sin g2
k1 = e1 cos g1
k2 = e2 cos g2.
(38)
(39)
Since angular momentum is conserved and the mutual inclination stays constant to first order,
the two eccentricities vary 90◦ out of phase. The linear system of equations describing the secular
evolution of eccentricities in planetary perturbation theory is
dh1
dt
dk1
dt
dh2
dt
= +A11k1 − A12k2
= −A11h1 + A12h2
= −A21k1 + A22k2
(40)
(41)
(42)
-- 17 --
Fig. 5. -- These results illustrate the potential danger of using the quadrupole approximation. The
inner eccentricity is shown as a function of time for a system with m1/m0 = 10−3, m2/m1 = 1,
α−1 = 100, initial eccentricities e1 = 0.05 and e2 = 0.9, and an initial relative inclination i = 70◦.
Time is given in years assuming a1 = 1 AU and m0 = 1 M⊙. The solid line is from the integration
of the octupole-level perturbation equations, while the dashed line is from a direct numerical
integration of the three-body system (see §3.1). In the quadrupole approximation all oscillations
would have the same amplitude as the first shown here. Notice how the eccentricity oscillations
in fact increase in amplitude, making the inner periastron separation quite small. At some point
other effects such as general relativistic precession and tidal dissipation (not to mention a collision
between the two inner stars) could become significant.
-- 18 --
dk2
dt
= +A21h1 − A22h2,
(43)
where the A's are defined in terms of Laplace coefficients, which we truncate at third order in α
to obtain
A11 =
3
4
k
a3/2
1
a3
2
m2
√m0 + m1
,
A12 =
15
16
k
a5/2
1
a4
2
m2
√m0 + m1
A21 =
15
16
k
a3
1
a9/2
2
m1
√m0 + m2
,
A22 =
3
4
k
a2
1
a7/2
2
m1
√m0 + m2
,
,
(44)
(45)
where a1 and a2 are the averaged semimajor axes measured from m0 to m1 and m2, respectively.
Now we rewrite the system in terms of familiar quantities to find
de1
dt
de2
dt
= A12e2 sin g,
= −A21e1 sin g,
dg
dt
= A22 − A11 + A12
e2
e1
cos g − A21
e1
e2
cos g,
(46)
(47)
(48)
where g = g2 − g1. Indeed the use of this variable is convenient, since for coplanar orbits there is
no well-defined line of nodes, and only the relative longitudes of perihelia are important.
Upon expanding the octupole equations (29) -- (32) to first order in e1 and e2 we obtain an
identical linear system of differential equations, but with the A's replaced by
B11 =
B12 =
B21 =
B22 =
12C2
L1
=
3
4
k
4C3
L1
4C3
L2
=
=
15
16
15
16
12C2
L2
=
3
4
k
k
k
a3/2
1
a3
2
a5/2
1
a4
2
a3
1
a9/2
2
a2
1
a7/2
2
m2
,
√m0 + m1
m2(m0 − m1)
(m0 + m1)3/2 ,
m0m1(m0 − m1)√m0 + m1 + m2
(m0 + m2)3
m0m1√m0 + m1 + m2
(m0 + m1)2
.
(49)
(50)
(51)
(52)
,
It is easy to see that, in the limit where m0 ≫ m1 and m0 ≫ m2, the two sets of equations
coincide, as they should. This establishes the accuracy of our analytic results in this limit.
In general, the two sets of equations differ in the dependence of the coefficients on the masses.
Although the two theories use different coordinate systems (Jacobi vs heliocentric), this alone does
-- 19 --
not explain the difference. Instead, one must remember that the classical theory was derived from
Lagrange's planetary equations (see Brouwer & Clemence 1961), which assume that the disturbing
functions (proportional to m1 and m2) are small. Therefore the approximation is valid only if
m0 ≫ m1, m2. This can also be seen by considering the m0 = m1 case, for which Heggie & Rasio
(1996, App. B) proved that the variation of e1 vanishes to all orders in α if the initial e1 = 0. In
contrast, the classical planetary perturbation equation (46) would predict a nonzero perturbation
of e1 for this case (since A12 6= 0, while our coefficient B12 = 0 for m0 = m1).
Our octupole-level analytic results do not depend on any assumption about the three masses,
as long as the system can be modeled as a hierarchical triple. The octupole equations predict
constant eccentricities in the case where m0 = m1 and i = 0. This happens because the odd-power
coefficients are proportional to m0 − m1 in the Hamiltonian (1), so the leading terms vanish.
There is no reason to expect the octupole approximation to work for this very particular case.
However, in §3.6 we show explicitly by comparison with direct numerical integrations that the
mass-dependences of our equations (29) -- (32) are valid for wide ranges of both mass ratios.
3. Comparison with Direct Numerical Integrations
We have performed extensive numerical integrations of hierarchical triple systems using
both our octupole-order secular perturbation equations (hereafter OSPE) and direct three-body
integrations. In this section we present a sample of results that establish the validity and accuracy
of our analytic results, and at the same time illustrate the dependences of the perturbations on
different parameters.
3.1. Numerical Methods
For the numerical integration of the OSPE we change variables from (e1, g1, e2, g2) to
(e1 sin g1, e1 cos g1, e2 sin g2, e2 cos g2) to remove singularities associated with the longitude of
pericenter for circular orbits. We perform the numerical integration of the OSPE using the
Burlisch-Stoer integrator of Press et al. (1992). Energy and angular momentum are automatically
conserved, since the semimajor axes are considered constant and the relative inclination of the
orbits varies to conserve angular momentum. We present results obtained using an accuracy
parameter EPS = 10−8. We found that reducing the integration step sizes did not make a
significant numerical difference for several test systems.
We have compared the results of the OSPE integrations with direct three-body integrations
using a fixed timestep mixed-variable symplectic (hereafter MVS) integrator (Wisdom & Holman
1991) available in the software package SWIFT (Levison & Duncan 1991). For most integrations,
we used a timestep of P1/40, where P1 is the orbital period of the inner binary. Energy and
angular momentum were typically conserved to one part in 106 and 1012, respectively. For some
-- 20 --
high-eccentricity systems we reduced the timestep to a value as small as P1/600. This MVS
integrator was designed for systems in which m0 is much larger than both m1 and m2. For
systems with m1/m0 >∼ 0.1 we made use of a newer MVS integrator modified to accommodate
arbitrary mass ratios, kindly provided to us by Jack Wisdom and Matt Holman. In a number of
test calculations with this newer integrator, we varied the timestep systematically to verify that
our results are not affected by numerical errors.
To complement integrations using the MVS integrator, we also used SWIFT's Burlisch-Stoer
(BS) integrator in a few test runs. This integrator is valid for arbitrarily strong interactions
between any pair of bodies, but is not well suited for very long integrations. We have only
performed a small number of these tests, since the BS integrations require a much longer computer
time. Energy and angular momentum in BS integrations were typically conserved to one part in
105. Most BS integrations were stopped after one full oscillation of e1, which we used to determine
the "maximum eccentricity perturbation" of the inner orbit, although, as pointed out in §2.3, the
true long-term secular evolution of the eccentricity will not, in general, be strictly periodic. A
typical BS run lasting for ∼ 105P2 took about 400 CPU hours on a MIPS R10000 processor, while
the same run using the MVS integrator would only take about 2 CPU hours.
The numerical integrations all started with initial values of the inner and outer arguments of
pericenter of 0◦ and 180◦, respectively. This choice leads to the maximum eccentricity induced
in the inner binary in both the planetary limit (see §2.4) and the quadrupole approximation (see
§2.3). We have performed additional integrations to verify that the remaining angles (longitudes
of ascending node and initial anomalies) do not significantly affect the secular evolution of the
system. For large inclinations, the initial argument of pericenter of the inner binary is important
in determining whether the system will undergo circulation or libration, if the inner orbit has a
significant initial eccentricity (see Fig. 3). However, if the inner orbit is nearly circular initially,
then the initial values of the angles are of little importance since the inner orbit can switch
from circulation to libration and vice versa. For coplanar orbits, the magnitude of the angular
momentum of the inner orbit increases when g = g2 − g1 < 0 and decreases when g > 0. For small,
nonzero inclinations this can still serve as a guide when considering the effect of varying the angles.
3.2. Eccentricity Oscillations
First, we investigate the dependence of the maximum eccentricity perturbation of the inner
orbit on the initial relative inclination (see Fig. 6). We see that, as expected (Sec. 2.3), for small
inclinations, i <∼ 40◦, the perturbations are dominated by the octupole term, while for higher
inclinations the quadrupole-level perturbations dominate. In both regimes the OSPE results match
the direct numerical integrations very well. Near the transition, numerical integrations (both
OSPE and MVS) show a beat-like pattern of eccentricity oscillations suggesting an interference
between the quadrupole and octupole terms. Note that the results of Figure 6 are for a system
with m1 ≪ m0 and m2 ≪ m0, for which the analytic results from the classical planetary
-- 21 --
Fig. 6. -- Maximum eccentricity of the inner orbit after a single oscillation, as a function of the
relative inclination. Here m1/m0 = 10−3, m2/m0 = 0.01, α−1 = 100, e2 = 0.05, and the initial
e1 = 10−5. The squares are from MVS integrations, and the double dashes on either side are from
OSPE integrations with varying initial longitude of periastron. The horizontal line indicates the
amplitude of the eccentricity oscillations calculated analytically in the planetary theory (§2.4). The
solid curve indicates the amplitude of eccentricity oscillations calculated analytically according to
the quadrupole-level theory for i >∼ 40◦.
-- 22 --
perturbation theory (Sec. 2.4) can be applied for small eccentricities and inclinations. We see that
the agreement with both MVS and OSPE integrations is excellent for i <∼ 30◦.
We now discuss in some more detail the evolution of systems in the low- and high-inclination
regimes.
3.2.1. Large-inclination Regime
Figure 7 illustrates the evolution of e1, g1, e2, g2, and i obtained from a numerical MVS
integration for a typical system with large relative inclination. For large inclination, the secular
quadrupole-level perturbations dominate the evolution. In the quadrupole approximation the inner
eccentricity undergoes periodic oscillations, while the outer eccentricity remains constant. Indeed,
we see in Figure 7 that e1 undergoes approximately periodic oscillations of large amplitude (with
corresponding oscillations in i), while e2 remains approximately constant. The small-amplitude
(about 10%) fluctuations in e2 are due mainly to the smaller, octupole-level perturbations.
Deviations from strict periodicity in the variation of e1 and i are also caused by octupole-level
perturbations. The period of a quadrupole eccentricity oscillation is a function of the mass ratios
and the outer eccentricity (see §2.3). Our numerical integrations of the OSPE reveal that the
most significant corrections to this period come from the variable time spent at low eccentricities.
Equation (35) implies that the time derivative of e1 is small when the inner orbit has a small
eccentricity. Then the octupole (and higher-order) perturbations can become important, causing
significant variation in the time a system will spend with a small inner eccentricity. This effect
can be seen clearly in Figure 7.
The OSPE do not correctly describe the evolution of a systems starting with e1 = 0. However
for any system with arbitrarily small but nonzero e1, the inner orbit can switch back and forth
between libration and circulation in the e1-g1 plane, achieving the full range of eccentricities. In
our MVS integrations the short-period variations (averaged out in secular perturbation theory)
provide the necessary perturbations to allow for the full eccentricity oscillations, even if the initial
e1 = 0.
3.2.2. Small-inclination Regime
For small inclinations (i <∼ 40◦), the secular octupole-level perturbations dominate and both
e1 and e2 typically undergo very small-amplitude fluctuations, as does the relative inclination
(Fig. 8). In the octupole approximation angular momentum is periodically transferred from one
orbit to the other. In this regime, the special case of initially circular orbits is stable to eccentricity
oscillations. Short-period perturbations will still cause small fluctuations in both eccentricities, but
since the inclination does not undergo large oscillations, the angular momentum transferred from
-- 23 --
Fig. 7. -- Typical evolution of the eccentricities, longitudes of periastron, and relative inclination
for a system in the high-inclination regime. Here m1/m0 = 10−3, m2/m0 = 0.01, α−1 = 100, the
initial inclination i = 60◦, and the initial eccentricities e1 = 10−5 and e2 = 0.05. Time is given
in years assuming a1 = 1 AU and m0 = 1 M⊙. These results were obtained using numerical MVS
integrations.
-- 24 --
one orbit to another is limited by angular momentum conservation, preventing large eccentricities
from developing in either orbit.
For inclinations approaching ∼ 40◦, the octupole-level interaction still leads to a noticeable
amplitude oscillation superimposed onto the quadrupole-level result. For a small range of
inclinations the two eccentricity oscillations can become comparable leading to a secular evolution
with a period and amplitude larger than either of the two oscillations in isolation.
3.3. Dependence on the Ratio of Semimajor Axes
Numerical results illustrating the dependence of the maximum eccentricity perturbation of the
inner orbit on the ratio of semimajor axes are shown in Figure 9. Some small deviations between
the OSPE and MVS results appear for small a2/a1, where higher-order secular perturbations
may be significant. As predicted by the quadrupole-level approximation, the amplitude of the
eccentricity oscillations becomes independent of α for high inclinations.
3.4. Dependence on the Initial Eccentricity
Figure 10 shows the dependence of the maximum inner eccentricity on its initial value. For
low inclinations increasing the inner eccentricity nearly adds to the maximum induced eccentricity.
In the high-inclination regime increasing the initial inner eccentricity does not affect the maximum
inner eccentricity significantly until the two become comparable.
3.5. Dependence on the Outer Eccentricity
Figure 11 shows the effect of varying the outer eccentricity. The OSPE and MVS integrations
agree precisely for moderate eccentricities, but show discrepancies for both very large and very
small e2.
For low i and e2 <∼ 10−2, short-period eccentricity variations become important. These are not
included in the OSPE since they were averaged out of the Hamiltonian. Formally, i = e1 = e2 = 0
is a fixed point, since it implies de1/dt=0 in the OSPE. For low inclinations and eccentricities, the
short-period eccentricity oscillations determine the maximum eccentricity, since this fixed point is
stable to the small-amplitude short-period perturbations. However, for large relative inclinations,
the initial condition e1 ≃ e2 ≃ 0 will lead to large amplitude oscillations as discussed in §3.2.1.
Thus, the small-amplitude, short-period oscillations not included in the OSPE allow the system
to explore the full range of allowed eccentricities.
For large e2, some discrepancies may be caused by inaccuracies in the MVS integrations: the
-- 25 --
Fig. 8. -- Typical evolution of a system in the low-inclination regime. All parameters are as in
Fig. 7, except that the initial inclination i = 15◦. These results were obtained using numerical
MVS integrations.
-- 26 --
Fig. 9. -- Maximum eccentricity e1 as a function of the ratio of semimajor axes α−1 = a2/a1.
The integrations are for a system with m1/m0 = 10−3, m2/m0 = 0.01, and initial eccentricities
e1 = 10−5 and e2 = 0.05. The various symbols (lines) are from MVS (OSPE) integrations of
systems with various relative inclinations: 0◦ (open triangles, solid line), 15◦ (open squares, dotted
line -- here quasi-indistinguishable from the solid line, cf. subsequent figures), 30◦ (open circles,
short-dashed line), 45◦ (solid triangles, long-dashed line), 60◦ (solid squares, short-dash-dotted
line), 75◦ (solid circles, long-dash-dotted line), and 89◦ (stars, short-dash-long-dashed line).
-- 27 --
Fig. 10. -- Maximum change in e1 as a function of its initial value. These integrations are for
m1/m0 = 10−3, m2/m0 = 0.01, e2 = 0.05, and α−1 = 100. The symbols and lines are as in Fig. 9.
-- 28 --
fixed timestep implies that periastron passages may not be fully resolved. We have performed
additional MVS integrations with a smaller timestep (shown in Fig. 11) and a smaller number
of BS integrations to verify that most of the discrepancy is indeed caused by inaccuracies in the
MVS integrator and not the OSPE. However, for sufficiently large e2, the disagreement remains.
As periastron passages begin to resemble close dynamical encounters, the averaging over orbits
becomes invalid, and the OSPE are no longer applicable. In this limit where the outer orbit is
nearly parabolic, it may be better to treat each periastron passage as a separate encounter. The
results of Heggie & Rasio (1996) may be used to calculate analytically the eccentricity perturbation
of the inner binary after each encounter.
3.6. Dependence on the Mass Ratios
First, we investigate the dependence of the maximum induced e1 on m2 (Fig. 12). The MVS
and OSPE integrations are in excellent agreement, except for very large m2. For sufficiently
large m2, the binding energy of m1 to m2 becomes comparable to its binding energy to m0, and
the inner orbit deviates significantly from a Keplerian orbit, making the basic assumption of a
hierarchical triple invalid. As discussed in §3, the MVS integrator was not designed for large
m2/m0. However, we have performed a number of test integrations, both BS and MVS (with a
smaller timestep), and found that the MVS integrations are generally accurate even for m2 >∼ m0,
provided that the system is stable.
Next, we explore the effects of varying the ratio m1/(m0 + m1) (Fig. 13). The agreement
between MVS and OSPE results is very good, even when m1 ≃ m0. In the OSPE, the octupole-
level perturbations vanish when m2/m1 = 1, removing the dominant term of the expansion for
low inclinations. Therefore we did not expect the OSPE to properly model the systems with
low inclinations. Using the modified MVS integrator of Wisdom and Holman (see §3.1) for
m1/m0 >∼ 0.1, we find surprisingly good agreement between the OSPE and MVS results for both
low and high inclinations. In particular the OSPE and MVS integrations agree on the maximum
induced eccentricity in the equal-mass case, m1 = m0, which is an important case for binary stars.
Additionally, the OSPE and MVS results show similar peaks in the maximum induced eccentricity
around the resonance between the quadrupole and octupole terms. Both MVS and limited BS
integrations also indicate that the vanishing of the induced eccentricity for low-inclination systems
when m0 = m1 is real. (Unfortunately, these systems are very time-consuming to integrate
numerically with a BS integrator, prohibiting us from doing a more thorough investigation.) For
example, for the initial conditions m0 = m1 = m2, e1 = e2 = 0, α−1 = 100, and i = 0, we observed
only short-term eccentricity fluctuations of magnitude ∼ 10−12. Thus, we conclude that the OSPE
results are accurate for all values of the inner mass ratio m1/(m0 + m1).
-- 29 --
Fig. 11. -- Maximum e1 as a function of the initial outer eccentricity, e2, for a system with
m1/m0 = 10−3, m2/m0 = 0.01, and initial e1 = 10−5. For e2 < 0.6 we used α−1 = 100 as in
previous figures, while for e2 > 0.6 we increased the value to α−1 = 300 (to avoid close interactions
with m2). The symbols and line styles are as in Fig. 9.
-- 30 --
Fig. 12. -- Maximum e1 as a function of the mass of the outer body, m2, for a system with
m1/m0 = 10−3, α−1 = 100, and initial eccentricities e1 = 10−5 and e2 = 0.05. The symbols and
line styles are as in Fig. 9.
-- 31 --
Fig. 13. -- Maximum e1 as a function of the mass ratio of the inner binary, m1/(m0 + m1), for a
system with the same parameters as in Fig. 12. The symbols and line styles are as in Fig. 9.
-- 32 --
3.7. Summary and Discussion
We have performed a large number of numerical integrations (including many not shown here)
to establish the validity of our analytic results for a broad range of triple configurations. The only
significant difference we observed was in the regime where e1 ≃ 0. In that regime the system will
chaotically choose circulation or libration about an island in the (e1, g1) phase space. Since e1 = 0
creates a singularity in the OSPE, we circumvented this problem by starting runs with e1 = 10−5.
While varying the timestep affected when m1 chose to librate or circulate, it did not create any
significant difference in the ratio of circulation to libration time.
We conclude that the OSPE provide an accurate description of the secular evolution of
hierarchical triple systems (containing unevolving point masses and in Newtonian gravity)
for nearly all inclinations, initial eccentricities, and mass ratios. The OSPE may be used for
small e1, provided that e1 6= 0, since this can be unstable to large oscillations. When secular
perturbations are sufficiently small, short-period perturbations may provide the larger contribution
to the eccentricity oscillations. The OSPE are not applicable when m0/m2 < α ≡ a1/a2,
since the inner orbit is then no longer nearly-Keplerian. The OSPE also break down whenever
a2(1− e2)/a1(1+ e1) <∼ 3− 5, since the triple system is then likely unstable and its evolution will not
be dominated by secular effects. Similarly, the OSPE should not be applied when a1(1 − e1) <∼ R0,
where R0 is the radius of the larger of the two inner stars, since the tidal interaction with that star
would then be important. One should also be careful whenever e1 ≃ 1, since a small fractional
error in e1 can lead to a significant change in rp,1 ≡ a1(1− e1) which is important in differentiating
purely gravitational interactions from a strongly dissipative tidal encounter or collision between
the inner components.
4. Resonant Perturbations: The Case of PSR B1620−26
4.1.
Introduction
Hierarchical triple systems can be affected by many different types of perturbations acting
on secular timescales. In general, during a given phase in the evolution of a triple, only one type
of perturbation will be important. However, it is possible that, in some cases, two perturbation
mechanisms with different physical origins may be acting simultaneously and combine in a
nontrivial manner. In particular, whenever two perturbations are acting on comparable timescales,
the possibility exists that they will reinforce each other in a nonlinear way, leading to a kind
of resonant amplification. This is not to be confused with orbital resonances, which can lead to
nonlinear perturbations of two tightly coupled Keplerian orbits when the ratio of orbital periods
is close to a ratio of small integers (see, e.g., Peale 1976)
Perturbation effects coming from the stellar evolution of the components or from tidal
dissipation in the inner binary were mentioned briefly in §1 and will not be discussed extensively
-- 33 --
in this paper. Instead, we consider the case where the inner binary contains compact objects
and its orbit is affected by general relativistic corrections on a timescale comparable to that of
the Newtonian secular perturbations calculated in §2. Rather than basing our discussion on
hypothetical cases, we concentrate on the real example provided by the PSR B1620−26 system.
4.2. The PSR B1620−26 Triple System
PSR B1620−26 is a millisecond radio pulsar in a triple system, located near the core of the
globular cluster M4. The inner binary consists of a ≃ 1.4 M⊙ neutron star with a ≃ 0.3 M⊙
white-dwarf companion in a 191-day orbit with an eccentricity of 0.025. The mass and orbital
parameters of the third body are less certain, since the duration of the radio observations covers
only a small fraction of the outer period. However, from the modeling of the pulse frequency
derivatives as well as short-term orbital perturbation effects it appears that the second companion
is most likely a low-mass object (m2 ≃ 0.01 M⊙) in a wide orbit of semimajor axis a2 ≃ 50 AU
(orbital period P2 ≃ 300 yr) and eccentricity e2 ≃ 0.45 (Joshi & Rasio 1997; Ford et al. 2000).
The eccentricity of the inner binary, although small, is several orders of magnitude larger than
expected for a binary millisecond pulsar of this type, raising the possibility that it may have been
produced by long-term secular perturbations in the triple.
An analysis based on the classical planetary theory (i.e., for small relative inclination ignoring
general relativistic precession) shows that a second companion of stellar mass would be necessary
to induce an eccentricity as large as 0.025 in the inner binary (Rasio 1994, 1995). Such a large
mass for the second companion has now been ruled out by recent pulsar timing data, and by the
absence of an optical counterpart for the system (Ford et al. 2000).
It is reasonable to assume that the relative inclination is large, since the location of the system
near the core of a dense globular cluster suggests that the triple was formed through a dynamical
interaction between binaries (Rasio, McMillan, & Hut 1995; Ford et al. 2000). For a sufficiently
large relative inclination, we have seen (Fig. 6) that it should always be possible to induce an
arbitrarily large eccentricity in the inner binary. Therefore, this would seem to provide a natural
explanation for the anomalously high eccentricity of the binary pulsar in the PSR B1620−26
system (Rasio, Ford, & Kozinsky 1997). However, there are two additional conditions that must
be satisfied for this explanation to hold.
First, the timescale for reaching a high eccentricity must be shorter than the lifetime of the
triple system. In this case the lifetime of the triple is determined by the timescale for encounters
with passing stars in the cluster, since any such encounter is likely to disrupt the orbit of the (very
weakly bound) second companion. As discussed in detail by Ford et al. (2000), this is unlikely to
be the case in the high-inclination regime of secular perturbations, given the parameters of PSR
B1620−26 and its location near the core of M4 (or inside -- it is seen just inside the edge of the
core in projection).
-- 34 --
Second, the secular perturbation of the inner binary pulsar by its distant second companion
must be the dominant source of orbital perturbation. Additional perturbations that alter the
longitude of periastron of the inner binary can indirectly affect the evolution of its eccentricity. For
a binary pulsar, general relativity contributes a significant orbital perturbation. If the additional
precession of periastron induced by general relativity is much slower than the precession due to
the Newtonian secular perturbations, then the eccentricity oscillations should not be significantly
affected. However, if the additional precession is faster than the secular perturbations, then
eccentricity oscillations may be severely damped (Holman et al. 1997; Lin et al. 1998). In addition,
if the two precession periods are comparable, then a type of resonance could occur, leading to a
significant increase in the eccentricity perturbation.
4.3. Secular Evolution of the Eccentricity
We have used the OSPE to study the secular evolution of the inner binary eccentricity in
the PSR B1620−26 system. We integrate the system using the variables h1, h2, k1, k2 (eqs. 38 &
39), which makes it easy to incorporate the first-order post-Newtonian correction. We restrict our
attention to the one-parameter family of orbital solutions calculated by Ford et al. (2000), based
on the modeling of the four pulse frequency derivatives measured by Thorsett et al. (1999). For
each solution, the maximum induced eccentricity of the inner orbit depends only on the (unknown)
relative inclination of the two orbits. In Figure 14, we show this maximum induced eccentricity as
a function of the second companion's semimajor axis for several inclinations.
For most solutions and most values of the inclination, the maximum induced eccentricity
remains significantly smaller than the observed value of 0.025. However, for low enough
inclinations, there is a narrow range of solutions (around a2 ≃ 45) for which the observed
value can be reached. Most remarkably, these solutions are also the ones currently preferred if
one takes into account preliminary measurements of the fifth pulse derivative and short-term
orbital perturbation effects in the theoretical modeling (see Ford et al. 2000). We also see from
Figure 14 that the maximum induced eccentricity has a peak where the precession period due to
the secular perturbation of the second companion is comparable to the precession period due to
general relativity, as expected. As the inclination increases, the maximum induced eccentricity
(at the peak) becomes smaller and the peak moves towards lower values of a2. This pattern
continues for inclinations slightly larger the normal cutoff for the low-inclination limit (≃ 40◦).
For relative inclinations 50◦ <∼ i <∼ 70◦ we do not find a peak in the maximum induced eccentricity.
For inclinations >∼ 75◦ we again find a peak, which becomes smaller and moves towards larger
separations as the relative inclination of the two orbits is increased.
The maximum inner eccentricity is also limited in this case by the relatively short lifetime of
the triple system in M4, ∼ 107 − 109 yr depending on whether the system resides inside or outside
the cluster core (Ford et al. 2000). For solutions near a resonance, the inner eccentricity starts
growing linearly at nearly the same rate as it would without general relativistic perturbations.
-- 35 --
Fig. 14. -- The top panel compares several timescales in the PSR B1620−26 binary pulsar as a
function of the semimajor axis, a2, of its second companion: PGR is the general relativistic precession
period of the inner binary; PHigh i and PLow i are the periods of the eccentricity oscillations in the
high and low relative inclination regimes, respectively; τc and τhm are the expected lifetimes of
the triple in the core of M4 and at the half-mass radius, respectively. The bottom panel shows
the maximum induced eccentricity of the inner binary for several different values of the (unknown)
relative inclination. The peaks correspond to a resonance between the general relativistic precession
of the inner orbit and the Newtonian secular perturbation by the second companion.
-- 36 --
However the period of the eccentricity oscillations can be many times the period of the classical
eccentricity oscillations. Although this allows the eccentricity to grow to a larger value, the
timescale for this growth is also longer. For PSR B1620−26, Ford et al. (2000) show that, near
resonance, the inner binary achieves an eccentricity of 0.025 in a time comparable to the expected
lifetime of the triple in the core of M4. However, the location of the pulsar near the edge of the core
(in projection) suggests that it may in fact reside well outside the cluster core, where its lifetime
can be significantly longer. In summary, the resonance between general relativistic precession and
Newtonian secular perturbations by the outer companion provides a possible explanation for the
inner binary's eccentricity.
4.4. Comparison with Direct Numerical Integrations
We have conducted a few long integrations with our MVS symplectic integrator for model
systems similar to PSR B1620−26, in order to check the validity of the OSPE in the presence of a
resonance (Fig. 15). Although there is good overall agreement, we find that both the amplitude
and the width of the peak is slightly narrower in the MVS integrations. Note that the values of
the masses and semimajor axes were decreased in order to speed up the direct integrations and to
satisfy the assumptions of the well-tested MVS integrator provided in SWIFT (i.e., m1 ≪ m0 and
m2 ≪ m0). This results in smaller values for the ratio of semimajor axes, implying less accurate
results from the OSPE. Nevertheless, the OSPE integrator performs well, even near a resonance
such as the one produced by general relativistic precession in a system like the PSR B1620−26
triple.
5. Application to Other Observed Triples
5.1. The 16 Cygni Binary and its Planet
The 16 Cyg system contains two solar-like main-sequence stars in a wide orbit (separation
∼ 103 AU) and a low-mass companion orbiting 16 Cyg B in a 2.2-yr (∼ 1.7 AU) eccentric orbit
(e = 0.67). The amplitude of the observed radial velocity variations indicate that the low-mass
companion has a mass m sin i ≃ 0.6MJup, suggesting that it is a giant planet (Cochran et al. 1997).
However the large orbital eccentricity is surprising for a planet.
Holman et al. (1997) and Mazeh et al. (1997) pointed out that the secular perturbation
by 16 Cyg A could explain the large eccentricity of the planetary orbit for a sufficiently large
relative inclination. In order for the quadrupole perturbations to be effective the precession of
the longitude of periastron must be dominated by the secular perturbations of 16 Cyg A. The
general relativistic precession period (∼ 7 × 107 yr) can be significant in the eccentricity evolution
of the planet (Holman et al. 1997). Similarly, if additional companions to 16 Cyg B are found in
-- 37 --
Fig.
15. -- Maximum induced eccentricity e1 as a function of the ratio of semimajor axes,
α−1 = a2/a1. The different symbols show the results of numerical integrations with and without the
general relativistic term, using both OSPE and MVS integrators: MVS integrations are shown by
squares, OSPE integrations are shown by triangles, integrations which include general relativistic
precession are indicated by empty symbols, integrations which ignore GR are shown with solid
symbols. These results are for a system similar to PSR B1620−26, but with the masses and
semimajor axes altered to facilitate the numerical integrations: m0 = 1.4 M⊙, m1 = 5 × 10−3 M⊙,
m2 = 8 × 10−4 M⊙, e1,init = 10−4, e2,init = 0.193, a1 = 2 × 10−4 AU, and i = 0◦.
-- 38 --
larger orbits (like those recently detected around Upsilon Andromedae; see Butler et al. 1999),
these would also induce a secular change in the longitude of pericenter of the presently known
planet. Thus, additional companions could also affect the eccentricity generated by the influence
of 16 Cyg A. Such an interaction could prevent the quadrupolar secular perturbation by 16 Cyg A
from accumulating long enough to produce the observed large eccentricity.
Hauser & Marcy (1999) have combined radial velocity and astrometric data to compute a
one-parameter family of solutions, which they tabulate as a function of the line-of-sight component
z of the position vector of B relative to A. There is also a possibility that 16 Cyg A may have an
M-dwarf companion which would affect the orbital solutions for the 16 Cyg AB binary and hence
the secular perturbation timescale (Hauser & Marcy 1999). We have used their orbital solutions
for 16 Cyg A (with no M dwarf companion) to estimate the effects of secular perturbations in
this system. If we assume that the eccentricity of 16 Cyg Bb is due to quadrupolar secular
perturbations, then both general relativity and any additional planets around 16 Cyg B could
constrain the orbit of the 16 Cyg AB binary. In Figure 16 we compare timescales for eccentricity
oscillations induced by 16 Cyg A, general relativistic precession, as well as the eccentricity
oscillations induced by a hypothetical second planet around 16 Cyg B. While the period of
eccentricity oscillations is shorter than 5 Gyr (the approximate age of 16 Cyg A&B; see Ford,
Rasio, & Sills 1998, 2000) for about 75% of the orbital solutions listed by Hauser & Marcy (1999;
we actually used an extended version of their Table 4 kindly provided by H. Hauser), the period of
the eccentricity oscillations is shorter than the general relativistic precession period for only about
25% of their solutions (assuming sin i ≃ 1; with sin i ≃ 0.5 this fraction increases to about 60%).
Given the large mass ratio of 16 Cyg A to 16 Cyg Bb and the high eccentricity of the orbit
(> 0.53), the ratio C3/C2 (see Eq. 23 & 24) can approach unity. Thus, the octupole term can be
very significant in the dynamics of this system on sufficiently long timescales. As an example,
we use the z = 0 solution of Hauser & Marcy (1999) and assume that the planet was initially
on a nearly circular orbit with an initial relative inclination of 60◦. We find that the period of
eccentricity oscillations is then ∼ 3 × 107 yr (∼ 4 × 107 yr if general relativistic precession is not
included; ∼ 6 × 107 yr if neither general relativistic precession nor the octupole term is included)
and the amplitude of the first eccentricity oscillation is ≃ 0.685 (≃ 0.767 without GR; ≃ 0.764
without GR or octupole term). However, there is a longer term oscillation with a period of
∼ 7 × 108 yr and an amplitude of 0.766 (≃ 0.774 without GR) which is not present when the
octupole term is ignored. Thus, in this example, the primary effect of general relativistic precession
is to reduce the fraction of the time where the planet has a very high eccentricity.
5.2. The Protobinary System TMR-1
HST/NICMOS observations of the TMR-1 system by Terebey et al. (1998) reveal, in addition
to the two protostars (of masses ∼ 0.5 M⊙) with a projected separation of 42 AU, a faint third
body (TMR-1C) that appears to have been recently ejected from the system. The association
-- 39 --
Fig. 16. -- This plot compares the timescales for precession of the longitude of pericenter of the
planet around 16 Cyg B due to the secular perturbations by 16 Cyg A (PA,High−i and PA,Low−i
for high and low relative inclination regimes, respectively), general relativity (PGR), and secular
perturbations by a hypothetical second planet (PBc), assuming that it is coplanar with 16 Cyg Bb
and has a mass of 1 MJup. Note that PA,High−i, PA,Low−i and PGR are plotted as a function of
the binary semimajor axis, while PBc is shown as a function of the orbital radius of the additional
planet. PA,High−i and PA,Low−i were calculated for the one-parameter family of orbital solutions
given by Hauser & Marcy (1999), which do not extend below a ≃ 900 AU. τms indicates the age of
16 Cyg B.
-- 40 --
of TMR1-C with the protobinary is suggested by a long, narrow filament that seems to connect
the protobinary to the faint companion. Given the observed luminosity of TMR1-C, and using
evolutionary models for young, low-mass objects, the estimated age of ∼ 3 × 105 yr for the system
leads to a mass estimate of ∼ 2 − 5MJup. This suggests that the object may be a planet that was
formed in orbit around one of the two protostars, and later ejected from the system (Terebey et al.
1998). If the age were increased to ∼ 107 yr, the mass would increase to ∼ 15 MJup, and TMR1-C
could also have been a low-mass, brown dwarf companion to one of the stars.
If TMR-1C is indeed a planet that was ejected from the binary system, this may place
significant constraints on planet formation theory. Here we speculate about the process which may
have led to the ejection of a planet from the TMR-1 system. In the standard planet formation
theory, TMR-1C must have formed in a nearly circular orbit around one of the protostars. Secular
perturbations by the other protostar may then have driven a gradual increase in the eccentricity
of the planet's orbit, gradually pushing the system towards instability. Large apocentre distances
render perturbations by the other protostar increasingly important. The planet could then enter
the chaotic regime in which it can switch into an orbit around the other protostar, possibly
switching between stars many times before finally being ejected from the system. The expected
velocity after such an ejection is in agreement with the estimated velocity of TMR-1C (de la
Fuente Marcos & de la Fuente Marcos 1998). One concern with this scenario is that the timescale
for ejection may be short compared to the timescale for planet formation. Early in the evolution
the protostellar disk will damp the planet's eccentricity. However, as the planet becomes more
massive, the gravitational perturbation by the other protostar becomes dominant. In fact, after
the protoplanetary core has formed, it may be able to accrete more mass than in the standard
scenario, since it is no longer confined to accrete from a narrow feeding zone.
We have investigated systems similar to TMR-1, but with the low mass companion in a nearly
circular orbit around one of the stars. We assume that the two protostars have masses of 1 M⊙
and 0.5 M⊙, with a binary semimajor axis of 50 AU and a planet mass of 5 MJup. We estimate
when the triple system will become unstable by combining our models of the secular eccentricity
evolution of the binary with the stability criterion of Eggleton & Kiseleva (1995). We calculate the
amplitude of secular eccentricity perturbations and see if the system would violate the Eggleton
& Kiseleva (1995) instability criterion (Y ≡ a2(1 − e2)/a1(1 + e1) < Ymin) when the planet's orbit
reaches its maximum eccentricity. For orbits with a large relative inclination, planetary semimajor
axes from 14 AU to 8 AU are expected to become unstable as the relative inclination is varied
from 40◦ to 85◦. For nearly coplanar orbits, planetary semimajor axes from 16 AU to 3 AU will
become unstable according to this criterion, as the outer eccentricity is varied from 0 to 0.8. Thus,
it seems plausible that a protoplanet could begin to form near the critical semimajor axis and
eventually be ejected from the system after it has accreted a large amount of gas. If we assume the
initial semimajor axis of the planet to be 5 AU, then we can solve for a critical binary eccentricity,
which we find to be 0.65. The period of the eccentricity oscillations responsible for reducing the
instability parameter from ≃ 10 to ≃ 3 is about 3 × 104 yr. In the coplanar regime we can also
-- 41 --
apply the stability criteria of Holman & Wiegert (1999)4. As the eccentricity of the TMR-1 binary
increases from 0 to 0.8, the critical semimajor axis decreases from about 11 AU to 2 AU. This is
precisely the region where giant planets are expected to form. If we assume the initial semimajor
axis of the planet to be 5 AU, then we can again solve for a critical binary eccentricity, which we
find to be 0.48. The two estimates are in reasonable agreement, and both also agree with the
results of preliminary numerical simulations which we have performed for this system.
We have conducted Monte Carlo simulations to study the process of planet ejection from
protobinaries (Fig. 17). For systems with large inclinations, the most common outcome for
unstable systems is a collision of the planet with its parent star. However, for systems with a low
relative inclination, the most common outcome was for the planet to be ejected from the system.
Furthermore, we found that in many cases it can take up to ∼ 107 yr for the planet to be ejected.
Since this is longer than a typical planetary formation timescale, the scenario proposed above
appears reasonable.
5.3. Systems with Short-Period Inner Binaries
Finally, we discuss briefly some observations and related theoretical work on triple systems
containing a short-period inner binary.
be possible to observe the secular perturbations directly, since the timescale for eccentricity
modulations and orbital precession may become comparable to the timescale of observations.
Unfortunately, in these systems, other perturbation effects such as tidal dissipation are likely to
affect the secular evolution, making the theoretical analysis more difficult.
If the outer period is also relatively short, it may
5.3.1. HD 109648
HD 109648 is a triple-lined spectroscopic triple probably composed of three main-sequence
stars all with masses ∼ 1 M⊙ (Jha et al. 1999). The inner orbit has a short period, ≃ 5.5 d,
so tidal dissipation effects are likely to be important. The small but significant eccentricity
(e1 = 0.0119 ± 0.0014) of the inner orbit has been attributed to the perturbation by the outer
companion (Jha et al. 1999). This system is strongly coupled, with α ≃ 0.1, so the timescale for
eccentricity modulations is short, Pe ∼ 15 yr. Thus, the available observations, spanning over 8
years, may already have detected changes in the inner eccentricity and longitude of pericenter.
Theoretical models by Jha et al. (1999) based on current data provide a loose constraint on the
4They define the critical semimajor axis as the largest orbital radius for which planets of all initial longitudes
of periastron survived for 104 binary periods. This is different from the criterion obtained by combining secular
perturbation theory with the results of Eggleton & Kisseleva (1995). However both criteria provide an estimate of
when the triple system becomes unstable. It is reassuring that both criteria yield similar results.
-- 42 --
Fig. 17. -- Results of Monte Carlo simulations for the dynamical evolution of possible progenitor
systems of TMR-1. The systems contain a 5 MJup planet in an initially circular 10 AU orbit
around a 1M⊙ star with a 0.5 M⊙ companion star in a 50 AU orbit. All other orbital parameters
(initial phases, longitudes of pericenter, relative inclination, and binary eccentricity) were assigned
random values. Triangles correspond to cases in which the planet was ejected (escaped) and
squares correspond to cases in which the planet collided with one of the stars. The time to
ejection (triangles) or collision (squares) is shown as a function of the initial stability parameter
Y ≡ a2 (1 − e2)/a1.
-- 43 --
relative inclination of the orbits: 5.9◦ ≤ i ≤ 54◦ or 126◦ ≤ i ≤ 174.1◦. Future observations are
likely to produce tighter constraints on this and other orbital parameters for the triple system.
However, Jha et al. (1999) speculate that additional variations may also be caused by the presence
of a fourth object in a much wider orbit.
5.3.2. HD 284163
HD 284163 is a triple system in the Hyades. The inner binary consists of a 0.72 M⊙ primary
and a secondary with a minimum mass of 0.33 M⊙ in a 2.4-day orbit (Griffin & Gunn 1981; Ford
& Rasio 2000). The outer companion (of mass ∼ 0.5 M⊙) has a projected separation of 7.4 AU
(Patience et al. 1998). Theoretical and empirical evidence indicate that tidal dissipation in the
primary should have circularized the inner binary (Ford & Rasio 2000). However, the radial
velocity curves indicate a significant eccentricity, e1 = 0.057 ± 0.005 (Griffin & Gunn 1981).
The secular perturbation by the outer companion is likely responsible for inducing this observed
eccentricity. At present, however, the outer orbit is not well constrained, making further analysis
difficult.
5.3.3. β Per
This is another triple system with a short-period (2.87 d) inner binary (1.7 M⊙ + 3.7 M⊙)
which is expected to have a very nearly circular orbit on the basis of tidal dissipation theory,
but has a significantly larger observed eccentricity of 0.0653. Secular perturbations by the outer
companion (mass 1.7 M⊙ in a 1.86-yr orbit) are likely responsible for maintaining the inner
binary's eccentricity (Kisseleva et al. 1998).
Kisseleva et al. (1998) suggest that the inner binary may have originally been significantly
wider. In their scenario, quadrupole perturbations drive a eccentricity increase. As the eccentricity
increases, tidal dissipation becomes significant and removes energy from the orbit. As the orbit
shrinks, precession of the longitude of periastron due to the stellar quadrupole moments and
general relativity increase, eventually suppressing the eccentricity perturbations. The secular
decrease in the semimajor axis due to the coupling of quadrupole perturbations and tidal
dissipation is then halted near the presently observed orbit.
In this system the ratio of semimajor axes is rather small, ≃ 40. Thus the octupole-level
perturbations could play an important role in the secular evolution. In particular, this could lead to
significantly larger eccentricities in the initial orbit if other effects have not yet started to suppress
the perturbations. Thus, the range of initial conditions that could lead to such an evolution can
be much larger than would be expected by considering quadrupole-level perturbations only.
-- 44 --
We are grateful to M. Holman and J. Wisdom for providing us with a version of their MVS
integrator modified for systems with more than one massive body, and to H. Hauser for sending us
an extended table of orbital solutions for the 16 Cygni binary. We thank S. Jha for pointing out
to us the recent theoretical work of Krymolowski and Mazeh and for providing us with a draft of
his paper on HD 109648 in advance of publication. Some of the numerical simulations mentioned
in §5.2 for TMR-1 were performed at MIT by J. Madic. F.A.R. thanks the Theory Division of the
Harvard-Smithsonian Center for Astrophysics for hospitality. This work was supported in part
by NSF Grant AST-9618116 and NASA ATP Grant NAG5-8460. E.B.F. was supported in part
by the Orloff UROP Fund and the UROP program at MIT. F.A.R. was supported in part by an
Alfred P. Sloan Research Fellowship. This work was supported by the National Computational
Science Alliance under Grant AST980014N and utilized the SGI/Cray Origin2000 supercomputer
at Boston University and the Condor system at the University of Wisconsin.
-- 45 --
REFERENCES
Anosova, J. 1996, Ap&SS, 238, 223
Bailyn, C.D., & Grindlay, J.E. 1987, ApJ, 312, 748
Backer, D.C. 1993, in ASP Conf. Ser. Vol. 36, Planets around Pulsars, ed. Phillips J.A. et al. (San
Francisco: ASP), 11
Brower, D., & Clemence, G.M. 1961, Methods of Celestial Mechanics (New York: Academic)
Brown, E.W. 1936, MNRAS, 97, 116
Butler, R.P., Marcy, G.W., Fischer, D.A., Brown, T.W., Contos, A.R., Korzennik, S.G., Nisenson,
P., & Noyes, R. 1999, ApJ, submitted
Cochran, W.D., Hatzes, A.P., Butler, R.P., & Marcy, G.W. 1997, ApJ, 483, 457
de la Fuente Marcos, C. & de la Fuente Marcos, R. 1998, NewA, 4, 21
Dermott, S.F., & Nicholson, P.D. 1986, Nature, 319, 115
Eggleton, P. 2000, Evolutionary Processes in Binary and Multiple Star Systems (Cambridge:
Cambridge Univ. Press), in preparation
Eggleton, P., & Kiseleva, L. 1995, ApJ, 455, 640
Eggleton, P.P., & Verbunt, F. 1986, MNRAS, 220, 13P
Ford, E.B., Joshi, K.J., Rasio, F.A., & Zbarsky, B. 2000, ApJ, in press (January 1)
Ford, E.B., Rasio, F.A., & Sills, A.S. 1998, ApJ, 514, 411
Ford, E.B., Rasio, F.A., & Sills, A.S. 2000, in preparation
Ford, E.B., & Rasio, F.A. 2000, in preparation
Georgakarakos, N. 2000, in preparation
Gies, D.R., & Bolton, C.T. 1986, ApJS, 61, 419
Goldstein, H. 1980, Classical Mechanics, Addison-Wesley
Griffin, R.F. & Gunn, J.E. 1981, ApJ, 86, 588
Hagihara, Y. 1972, Celestial Mechanics Vol. II, Part 2, MIT Press
Harrington, R.S. 1968, AJ, 73, 190
Harrington, R.S. 1969, Celes. Mech., 1, 200
Hauser, H.M. & Marcy, G.W. 1999, PASP, 111, 32
Heggie, D.C. 1975, MNRAS, 173, 729
Heggie, D.C. & Rasio, F.A. 1996, MNRAS, 282, 1064
Holman, M., Touma, J. & Tremaine, S. 1997, Nature, 386, 254
Holman, M.J., & Wiegert, P.A. 1999, AJ, 117, 621
-- 46 --
Iben, I., Jr., & Tutukov, A.V. 1999, ApJ, 511, 324
Jeffrys, W.H., & Moser, J. 1966, AJ, 71, 568
Jha, S., Torres, G., Stefanik, R.P., Latham, D.W., & Mazeh, T. 1999, in preparation
Joshi, K.J. & Rasio, F.A. 1997, ApJ, 479, 948
Kevorkian, J. 1966, AJ, 71, 878
Kiseleva, L.G., Eggleton, P.P., & Mikkola, S. 1998, MNRAS, 300, 292
Kozai, Y., 1962, AJ, 67, 591
Krymolowski, Y., & Mazeh, T. 1999, MNRAS, 304, 720
Levison, H. & Duncan, M. 1991, Icarus, 108, 18L
Lin, D.N.C., Papaloizou, J.C.B., Bryden, G., Ida, S., & Terquem, C. 1998, to be published in
Protostars and Planets IV, eds. A. Boss, V. Mannings, & S. Russell [astro-ph/9809200]
Lyne, A.G., Biggs, J.D., Brinklow, A., Ashworth, M., & McKenna, J. 1988, Nature, 332, 45
Marchal, C. 1990, The Three-Body Problem, (Amsterdam: Elsevier)
Mathieu, R.D., & Mazeh, T. 1988, ApJ, 326, 256
Mazeh, T., Kymolowski, Y., & Rosenfeld, G. 1996, ApJ, 477, L103
Mazeh, T., & Shaham, J. 1979, A&A, 77, 145
McKenna, J., & Lyne, A.G. 1988, Nature, 336, 226; erratum, 336, 698
McMillan, S., Hut, P., & Makino, J. 1991, ApJ, 372, 111
Mikkola, S. 1997, CeMDA, 68, 87
Mikkola, S., & Tanikawa, K. 1998, AJ, 116, 444
Patience, J., Ghez, A.M., Reid, I.N., Weinberger, A.J., & Matthews, K. 1998, AJ, 115, 1972
Peale, S.J. 1976, ARA&A, 14, 215
Portegies Zwart, S.F., Hut, P., McMillan, S.L.W., Verbunt, F. 1997, A&A, 328, 143
Press, W.H., Teukolsky, S.A., Vetterling, W.T., Flannery, B.P. 1992, Numerical Recipes in C: The
Art of Scientific Computing (New York: Cambridge University Press)
Rasio, F.A. 1994, ApJ, 427, L107
Rasio, F.A. 1995, in Millisecond Pulsars: A Decade of Surprise, ASP Conference Series, Vol. 72.
Eds. A.S. Fruchter, M. Tavani, & D.C. Backer (Astronomical Society of the Pacific, San
Francisco), 424
Rasio, F.A., McMillan, S., & Hut, P. 1995, ApJ, 438, L33
Rasio, F.A., Ford, E.B., & Kozinsky, B. 1997, AAS Meeting 191, 44.16
Soderhjelm, S., 1984, A&A, 141, 232
-- 47 --
Terebey, S., Van Buren, D., Padgett, D.L., Hancock, T., & Brundage, M. 1998, ApJ, 507, L71
Thorsett, S.E., Arzoumanian, Z., Camilo, F., & Lyne, A.G. 1999, ApJ, 523, 763
Tokovinin, A.A. 1997a, AstL, 23, 727
Tokovinin, A.A. 1997b, A&AS, 124, 75
Wisdom, J., & Holman, M. 1991, AJ102, 1528
Zahn, J.-P. 1977, A&A, 57, 383; erratum 67, 162
This preprint was prepared with the AAS LATEX macros v4.0.
|
astro-ph/0107462 | 2 | 0107 | 2001-11-15T16:13:46 | Ambipolar Drift in a Turbulent Medium | [
"astro-ph"
] | The interstellar magnetic field strength and gas density are observed to be correlated, but there is a large dispersion in this relation. In particular, the magnetic field is often observed to be weaker than expected. It is usually assumed that the field is frozen into the gas, except at very low ionization fraction, in which case ion-neutral drift, or ambipolar diffusion, leads to appreciable slip of the field and tends to make it more uniform. We show that turbulence enhances the rate of ambipolar drift, and that this may explain the surprisingly low fieldstrengths sometimes seen in dense interstellar gas. | astro-ph | astro-ph |
To appear in Astrophys J. March 10, 2002
Ambipolar Drift in a Turbulent Medium
Ellen G. Zweibel
JILA & Department of Astrophysical & Planetary Sciences, University of Colorado, Boulder CO
80309
ABSTRACT
The interstellar magnetic field strength and density are observed to be correlated,
but there is a large dispersion in this relation. In particular, the magnetic field is often
observed to be weaker than expected. At low ionization fraction, ion-neutral drift, or
ambipolar diffusion, permits slip of the field relative to the neutral gas and tends to
make the field strength more uniform, but is thought to be too slow to explain the obser-
vations. The purpose of this paper is to show that ion-neutral drift is significantly faster
in a turbulent medium than in a quiescent one. We suggest that this fast ambipolar
diffusion can explain the surprisingly low magnetic fieldstrengths sometimes observed
in dense interstellar gas.
Subject headings: ISM: magnetic fields, turbulence, MHD
1.
Introduction
The interstellar magnetic fieldstrength and gas density are observed to be correlated (Troland
& Heiles 1986, Crutcher 1999). This relationship is thought to arise from so-called ideal magne-
tohydrodynamic (MHD) processes, in which the field is frozen to the ambient medium. Under
ideal MHD conditions, if the ratio of mass to magnetic flux were everywhere constant, the slope
q ≡ d log B/d log ρ of the fieldstrength - density correlation would be unity for compression normal
to B, zero for compression parallel to B, and 2/3 for isotropic compression. The observed value of
q is approximately 0.5, which is consistent with equilibrium models of self gravitating clouds which
evolved under conditions of frozen flux (Mouschovias 1976, Tomisaka, Ikeuchi, & Nakamura 1988).
However, a number of observations and upper limits on magnetic fieldstrength suggest that
B ∝ ρ0.5 is more an upper envelope than a scaling law. This is true both in atomic and molecular
gas (Bourke et al. 2001, Crutcher 1999, Heiles 2001a, Heiles & Troland 2001) Although the number
of measurements is small, and the field may be underresolved in some cases (Brogan & Troland
2001), the trend towards weak fields is clear. A different line of argument comes from numerical
simulations of molecular clouds: Padoan & Nordlund (1999) claim that models with weak fields
replicate the observations better than models with strong fields.
-- 2 --
Weak fields are difficult to reconcile with ideal MHD. This is particularly so in the case of
turbulent molecular clouds. Giant molecular clouds (GMCs) are about 40 times more dense than
the mean interstellar gas, suggesting that the magnetic field should be 6-7 times stronger than the
mean field. If GMCs are formed by flows parallel to the field then of course the field would not
be strengthened, but the mean field is not strong enough to resist compression and collimate the
flows. Furthermore, Mestel (1985) has pointed out that accumulation of the mass of a GMC by
compression in one dimension requires organized motion of the gas over nearly one kiloparsec, the
origin of which is difficult to understand.
Thus, we seek an explanation for the weakness of the magnetic fieldstrength - gas density
correlation beyond the scope of ideal MHD. At the very largest lengthscales, ideal MHD should be
an excellent approximation. Moving downwards in scale, the first non-ideal effect encountered, at
least at low ionization fraction, is ion-neutral drift, or ambipolar diffusion. At the ambipolar scale,
which is many orders of magnitude larger than the resistive scale, the magnetic field and plasma
become decoupled from the neutral material. This makes it possible to change the mass to flux
ratio without altering the magnetic topology. Ambipolar drift has been invoked as the primary
magnetic flux transport mechanism in dense, star forming gas since the classic paper by Mestel &
Spitzer (1956). However, it is thought to be too slow to be an effective transport mechanism in
diffuse gas (see §2 for quantitative estimates).
Interstellar gas is turbulent. Turbulent diffusion of quantities such as heat and angular mo-
mentum is often invoked in astrophysics as a mechanism for enhancing transport rates above their
kinetic theory values, which are usually very slow. Turbulence enhances diffusion rates by mix-
ing the relevant quantity to the small scales at which molecular diffusion operates. This leads to
a mixing time which is approximately the eddy turnover time, and is nearly independent of the
molecular diffusivity.
Whether turbulence enhances the resistive decay rate of a magnetic field is unclear, because
there is substantial evidence that magnetic forces resist stretching the field sufficiently to mix it to
the tiny scales at which resistivity operates (Cattaneo & Vainshtein 1991, Cattaneo 1994, Cattaneo,
Hughes, & Kim 1996). This paper addresses a different question, namely, whether turbulence in a
weakly ionized gas can transport magnetic field with respect to the neutral matter, without resistive
dissipation necessarily coming into play. Because the ambipolar drift scale is much larger than the
resistive scale, the feedback effects which can suppress turbulent resistivity are far less dramatic,
although they cannot always be ignored.
We use analytical methods to calculate the mixing rate. Numerical study of mixing requires
that numerical diffusion of both field and fluid be very well controlled. Analytical calculations are
useful for initial exploration of some of the basic mechanisms.
In §2, we introduce the physical model, derive an equation for the evolution of the mass to
magnetic flux ratio in a weakly ionized medium, estimate relevant timescales, and establish an
initial condition. In §3, we estimate the turbulent diffusion rate based on mixing length theory,
-- 3 --
quantify this estimate with an exact calculation of transport by a wave, and argue that enhanced
diffusion requires the introduction of small scales as well as bulk advection. In §4 we develop a
model based on exponential stretching and shrinking. This model leads to a flux redistribution
rate which is comparable to the eddy turnover rate, as expected in turbulent diffusion problems.
However, the model is two dimensional, and the fluid motions are prescribed without specifically
allowing for magnetic forces. In §5, we consider magnetic feedback. In §6, we apply the model to the
interstellar medium, and discuss the astrophysical constraints imposed by the dynamics. Section
7 is a summary and conclusion. Sections 3 and 4 are relevant to general mixing problems, such as
turbulent diffusion of a passive scalar, and the reader who is mainly interested in the astrophysical
implications could go directly to §6.
2. Basic Equations and Model
2.1. Equations for Ambipolar Drift
We consider a weakly ionized medium with magnetic field B, mass density ρ, and ion mass
density ρi ≪ ρ. We are interested in timescales much longer than the ion-neutral collision time τin,
in which case the ion-neutral drift vD ≡ vi − vn is well approximated by
vD =
(∇ × B)×B
4πρi
τin,
(1)
(Shu 1983).
The magnetic field evolves according to the magnetic induction equation for a perfectly con-
ducting medium
∂B
∂t
= ∇ × (vi × B),
(2)
Replacing vi with vD + vn, and approximating vn by the center of mass velocity v, we rewrite the
induction equation as
∂B
∂t
= ∇ × (v × B) + ∇ × (vD × B),
(3)
where vD is given by eqn. (1). The first term on the RHS of eqn. (3) can be expanded using the
identity ∇ × (v × B) = B · ∇v − v · ∇B − B∇ · v. Then, using the continuity equation
∂ρ
∂t
= −v · ∇ρ − ρ∇ · v
we derive an evolution equation for the magnetic field to density ratio B/ρ
∂
∂t
B
ρ
+ v · ∇
B
ρ
=
B
ρ · ∇v +
1
ρ∇ × (vD × B).
(4)
(5)
The left hand side of eqn. (5) is the comoving, or convective, time derivative of B/ρ. The first term
on the right hand side represents stretching of the fieldlines, and is a consequence of the frozen
-- 4 --
flux condition. The second term on the right hand side represents the evolution of B/ρ caused by
ambipolar drift.
We now restrict ourselves to two dimensional, incompressible flows perpendicular to a straight
magnetic field. This geometry captures the main effects under study, and is consistent with the
nature of turbulence in a strong, well ordered magnetic field (Strauss 1976, Sridhar & Goldreich
1994, Goldreich & Sridhar 1997). We recognize that interstellar turbulence is frequently observed
to be highly supersonic, and thus cannot be entirely incompressible. We expect that compressible
turbulence to result in magnetic flux transport just as incompressible turbulence does, but that
magnetic feedback on the turbulence is stronger in the compressible case.
Under the assumptions of incompressibility and two dimensionality, the line stretching term
vanishes, and eqn. (5) reduces to
∂
∂t
B
ρ
+ v · ∇
B
ρ
=
1
ρ∇·
B 2
4πρ
τni∇B,
(6)
where B now represents the amplitude of the magnetic field, τni ≡ τinρn/ρi is the neutral-ion
collision time, and we have used eqn. (1). Equation (6) is close to an advection - diffusion equation
for the flux to mass ratio Q ≡ B/ρ
∂Q
∂t
+ v · ∇Q =
1
ρ∇·λ∇B,
(7)
provided that we define the diffusivity λ as τniv2
A, where B/(4πρ)1/2 is the Alfven speed.
The diffusion of Q is nonlinear in the sense that λ = λ(Q, ρ). This nonlinearity can produce
sharp fronts along surfaces where B vanishes or is tightly folded (Brandenburg & Zweibel 1994),
similar to fronts created by nonlinear thermal conduction (Zel'dovich & Raizer 1966). Resistive
diffusion in these current sheets alters the mass to flux ratio as well as changing the magnetic
topology (Brandenburg & Zweibel 1995, Zweibel & Brandenburg 1997). In this paper we assume
that the relative variation of B/ρ is so weak that nonlinear effects play only a minor role in
ambipolar drift, and B remains nonsingular.
Equation (7) can be used to derive an equation for the rate of change of B within a comoving
volume V of fluid; that is, a fluid element of fixed mass. We have
d
dtZV
d3xρQ =ZV
d3x
∂
∂t
ρQ +ZS
d2xρQv · n,
(8)
where S is the surface of V . The first term on the RHS of eqn. (8) is the Eulerian change of B and
the second term accounts for the motion of V . Using eqns. (4) and (7) to expand the first term on
the RHS and applying Gauss' theorem yields
d
dtZV
d3xρQ =ZS
d2xλn · ∇B.
(9)
Since the mass within V is constant, eqn. (9) shows that the flux to mass ratio within a volume
moving with the center of mass velocity decreases if the magnetic field decreases outward on the
surface of the volume.
-- 5 --
2.2. Timescales
In the absence of flow, the characteristic diffusion time for a magnetic field of representative
strength B and scale length L ∼ B/∇B is
L2
λ
td0 ≡
=
L2
v2
Aτni
=
τ 2
A
τni
, .
where τA ≡ L/vA is the global Alfven time.
(10)
Expressing td0 in physical units reveals the magnitude of the timescale problem. We take
τni from Draine, Dalgarno, & Roberge (1983); when the ratio of ion to neutral atomic weight
Ai/An ≫ 1, τni = 6.7 × 108n−1
s. The Alfven speed vA = 2.2 × 105Bµ/(nnAn)1/2 cm s−1. The
nxi) cm2 s−1, where xi ≡ ni/nn is the ionization fraction.
diffusivity λ is then 3.2 × 1019B 2
µ/(Ann2
The drift time is
i
td0 = 3.1 × 1020 N 2
20xiAn
B 2
µ
s,
(11)
where N20 ≡ nnL is the column density in units of 1020 cm−2. For example, the systems reported
by Heiles (2001) have N20 a few tenths to a few, with Bµ typically 3. If we take An = 1.4 for gas
of cosmic composition and xi = 10−4, which is probably a conservative lower limit, then we find
td0 is of order 108 yr for these systems. Since this is much more than the 106 - 107 yr expected
lifetime of an interstellar cloud, the diffusion rate must be enhanced by a factor of 10 - 100 in order
to explain the flatness of the B − n relation.
Mixing by eddies in the neutral gas requires that the magnetic field be frozen to the flow. The
degree of freezing is measured by the ambipolar Reynolds number RAD, which is large under frozen
in conditions (Zweibel & Brandenburg 1997). For eddies of characteristic size l and speed vt,
RAD ≡
Thus, the field is frozen to the turbulent flow for
lvt
λ
.
(cid:18) l
L(cid:19)(cid:18) vt
vA(cid:19) >
τni
τA
.
(12)
(13)
Since τni/τA is expected to be small, while vt/vA is order unity, eqn. (13) implies a fair degree of
dynamic range for the sizes of eddies that can mix the field. Equation (12) can also be written as
RAD =
v2
t
v2
A
τd
τni
,
(14)
which shows that if the flow is at or above equipartition with the field (vt ≥ vA), the field is frozen
in if the neutral-ion collision time is less than the eddy turnover time.
-- 6 --
2.3.
Initial Condition
It will be useful in the following analysis to have a definite model for B. We take as an initial
condition
B(x, y, 0) = B00 +
B
′′
0 x2,
1
2
(15)
′′
where B00 and B
0 < 0 are constants. We will define λ using B00 for B0 in quantitative examples.
We view eqn. (15) as the first two terms in a Taylor expansion of a magnetic field which peaks at
x = 0, and will assume x/L ≪ 1.
We first consider pure diffusion. Motivated by the initial condition (15), we seek solutions of
eqn. (7), with v = 0, of the form
B(x, y, t) = B0(t) +
B
′′
0 x2.
1
2
Substituting eqn. (16) into eqn. (7) and using eqn. (15), we find
B0(t) = B00 + B
′′
0 λt.
Equation (17) predicts that the peak field decreases by a factor of two on a timescale
Equation (18) agrees with eqn. (10) if we define the magnetic lengthscale L by
td0 ≡ −B00/(2λB
′′
0 ).
L =(cid:18)−
B00
2B ′′
0(cid:19)1/2
.
3. Diffusion in the Presence of Waves
(16)
(17)
(18)
(19)
We begin with a mixing length argument. Consider a magnetic flux tube of width a which is
carried by a random flow u a distance l down the gradient of B. The field in the tube diffuses into
the ambient medium; thus the motion causes net transport of B. The transport is most effective
when the diffusion time across the tube is comparable to the advection time
a2
λ ∼
l
u
,
(20)
because if a2/λ ≫ l/u, the flux tubes return to their original positions with nearly the same value
of the field, while if a2/λ ≪ l/u the field diffuses too quickly to be advected by the flow.
Advection spreads B over a distance l in a time l/u. In this time, B would spread diffusively
over a distance (λl/u)1/2. By eqn. (20), this distance is just a. Advective mixing accelerates the
transport of B only if a/l < 1, meaning that the motion consists of thin fingers which travel much
-- 7 --
further than their widths (see Ottino 1989 for discussion of mixing by fingers, or tendrils). Such
fingers are not seen in models of Alfvenic turbulence, and it is not clear that they would form in
a weakly stratified gas such as the interstellar medium. Interpreted more broadly, the argument
presented here shows that turbulent mixing requires more than just advection and dispersal; it also
requires the formation of small scales.
Now, we quantify this result. Weak turbulence theory, in which turbulence is modelled as a
superposition of randomly phased waves, can be used to compute the rate of turbulent diffusion (eg.
Moffatt 1978, Gruzinov & Diamond 1994). In this so-called quasilinear approach, one partitions
quantities into mean and fluctuating parts and calculates the average effect of the fluctuations on
the mean part. We used this method in a previous study of ambipolar diffusion (Zweibel 1988).
In the present problem it is possible, as well as instructive, to solve the induction equation exactly
instead of averaging it. This confirms the argument given in §2.
We introduce a periodic flow in the x direction
v = xu sin ωt sin ky.
(21)
Motivated by the initial condition eqn. (15), we try a solution of the advection-diffusion equation
(7) of the form
B(x, y, t) = B0(t) +
B
′′
0 + B1(t)x sin ky + B2(t) cos 2ky.
(22)
1
2
Substituting eqn. (22) into eqn. (7), using eqn. (21), and equating like powers of x and Fourier
harmonics of y leads to a set of coupled ODEs for the functions B0, B1, and B2. The solution for
B0, which follows the decay of the peak magnetic field, is
B0 = B00 + λB
′′
0 t
sin 2ωt
2ω (cid:19) +
′′
B
ω2
0 u2
2 (cid:18)t −
2 (ω2 + Γ2)" Γ
ω2 + Γ2 (cid:0)1 − e−Γt cos ωt(cid:1)
ω2 + Γ2 e−Γt sin ωt#,
ωΓ
+
+
−
sin2 ωt
2
(23)
where Γ ≡ λk2.
The maximum decay rate occurs when the motion given by eqn. (21) is coherent over many
wave periods. The long time behavior of B0 is then given by
where
B0 = B00 + (λ + λt) B
′′
0 t,
Γu2
4 (ω2 + Γ2)
λt ≡
(24)
(25)
-- 8 --
represents diffusion brought about by advective transport. Equation (25) closely resembles the
turbulent diffusivity calculated from quasilinear theory (Moffatt 1978). Maximizing λt with respect
to k, we find that the maximum occurs for ω = Γ, as we asserted in the mixing length argument
following eqn. (10), and is
λt,max =
u2
8λk2 ,
(26)
where we have replaced Γ by λk2. If we express u in terms of the maximum fluid displacement
a ≡ u/ω and take the ratio of λt,max to λ, the result is
k2a2
λt,max
(27)
=
λ
.
8
Equation (27) shows that the diffusion rate is appreciably enhanced by waves only if ka ≫ 1,
meaning that the flow consists of long, thin streamers (see also Press & Rybicki 1981). We reached
the same conclusion from mixing length theory. The missing ingredient is stretching and shrinking
of scales, an intrinsic feature of turbulent flows which we incorporate in the next section.
4. Stagnation Point Flow
Hyperbolic stagnation point flow is a particularly tractable example of a flow with exponen-
tial shrinking and stretching. At hyperbolic stagnation points, the fluid flow converges in one (or
two) directions and diverges in the other direction(s), while maintaining incompressibility.
It is
well known that diffusion is accelerated in the vicinity of stagnation points, due to the shrinking
of scales in the convergent directions (Moffatt 1978, Zweibel 1998), while Zel'dovich et al. (1984)
demonstrated dynamo action by a random ensemble of stagnation points. The role of hyper-
bolic stagnation points in the mixing of scalar fields in turbulent flows has recently been reviewed
(Shraiman & Siggia 2000) with emphasis on the development of intermittency, and the high order
moments of the distribution of concentrations.
The advection-diffusion problem for fields of the form (15) is exactly soluble for stagnation
point flow. In subsection (4.1) we compute the effect of a single stagnation point. In subsection
(4.2) we superimpose the effects of a random ensemble of stagnation points, and in subsection (4.3)
we discuss the relationship between the stagnation point model and turbulent flow. Our model is
not intended to be a full theory of turbulence, but merely illustrative.
4.1. A Single Stagnation Point
We consider two dimensional, incompressible flow near a stagnation point at (ax, ay). For fields
of the form given by eqn. (15), we require ax/L ≪ 1, where L is given by eqn. (19). The flow is
(28)
vx = −γ (x − ax) ,
vy = γ (y − ay) ,
(29)
-- 9 --
where γ is a constant. It is straightforward to integrate equations (28) to find the position at time
t of a fluid parcel which is at position (x0, y0) at t = 0
x = ax + (x0 − ax) e−γt
y = ay + (y0 − ay) eγt.
The initial coordinates in terms of the coordinates at time t are
x0 = ax + (x − ax)eγt
y0 = ay + (y − ay)e−γt.
(30)
(31)
(32)
(33)
We now compute the effect of this stagnation point flow on the diffusion of the magnetic field.
With eqn. (15) as the initial condition, we look for a solution of the form
B(x, y, t) = B0(t) +
B
′′
0 x2
0,
1
2
(34)
where x0(x, y, t) is given by eqn. (32). Substituting eqn. (34) into the advection-diffusion equation
(7) yields
B0 = λB
′′
0 e2γt.
(35)
B is of the form (34) exist for two reasons. First, x0 is a constant of the motion, so for any function
f (x0)
∂f (x0)
∂t
+ v · ∇f (x0) = 0,
(36)
where v is given by eqns. (28). Second, eqn. (32) shows that x0 is a linear function of x and y, so
∇2x0 is only a function of time, and is independent of the stagnation point location (ax, ay).
The solution of eqn. (35) which fits the initial conditions is
B0(t) = B00 +
λ
2γ
B
′′
0 (cid:0)e2γt − 1(cid:1) .
(37)
The location of the peak field evolves in time to ax(1 − e−γt), but this is irrelevant because the
density ρ is shifted by the same amount. It is only diffusion which affects the mass to flux ratio.
According to eqn. (35), the peak field has dropped to half its value in the time tγ
tγ =
1
2γ
ln (1 + 2γtd0) ,
(38)
where td0 is the diffusion time in the static case, defined in eqn. (18). Equation (38) shows that
the diffusion time depends only logarithmically on the diffusivity λ. This arises because of the
exponential growth of the magnetic field gradient, as seen in eqn. (35).
The quantity 2γtd0 which appears in the logarithm in eqn. (38) is, however, a large number.
If we identify γ−1 ∼ l/vt with an eddy turnover time τd, and use eqns. (10) and (12), we can write
(39)
2γtd0 ∼ 2
τ 2
A
τdτni ∼ 2RAD
L2
l2 .
-- 10 --
At the time tγ, the x component of the drift velocity vDx is half the flow velocity vx = γx.
Beyond this time, the magnetic field is not well coupled to the flow.
4.2. An Ensemble of Stagnation Points
We compute the evolution of the magnetic field carried by Lagrangian fluid elements under
sequences of stagnation point flows oriented in random directions, each one of which endures for a
time τ (see Childress & Gilbert 1995 for a general discussion of these so-called renewing flows).
We take the flow during time (n − 1)τ < t < nτ to be
n(cid:1)1/2
vnx = γµnx + γ(cid:0)1 − µ2
y,
n(cid:1)1/2
vny = γ(cid:0)1 − µ2
x − γµny,
where −1 ≤ µn ≤ 1 and n ≥ 1, and for simplicity we have taken all flows to have the same strength
γ. Equation (40) reduces to eqn. (28) if µ = −1. These flows are curl free, and hence not of the
most general possible type. However, in the neighborhoods of stagnation points, vorticity leads to
changes of scale at an algebraic rather than exponential rate, complicating the mathematics while
having little effect on the diffusion rate (Zel'dovich et al. 1984, Zweibel 1998). Thus, vorticity is of
secondary importance to our problem, and we omit it here, although interstellar turbulence almost
certainly possesses vorticity.
(40)
(41)
Let the (x, y) coordinates of a fluid parcel at time (n − 1)τ be rn−1. Then at time nτ , the
coordinates rn can be written as
where the matrix An is
rn = An · rn−1,
cosh γτ + µn sinh γτ
sinh γτ
(cid:0)1 − µ2
n(cid:1)1/2
Inverting eqn. (42) yields rn−1 in terms of rn
sinh γτ
cosh γτ − µn sinh γτ !
(cid:0)1 − µ2
n(cid:1)1/2
rn−1 = A−1
n ·rn.
(42)
(43)
Successive backwards iteration of eqn. (43) yields the initial position r0 in terms of the coordinate
rN ≡ r(N τ, r0)
r0 = A1
(44)
−1 · A2
−1...AN
−1 · rN.
At times nτ < t < (n + 1)τ , the position r(t) is related to the postion at time nτ by an equation
similar to eqn. (42), where in the matrix A we replace γτ by γ(t − nτ ).
Since the matrices An are independent of the spatial coordinates, r0 is a linear function of
0 is a function only of time. It follows that the
rN, or, more generally, r(t). This means that ∇2x2
-- 11 --
solution B of the advection- diffusion equation (7) can still be written in the form of eqn. (34), and
that the rate of diffusion increases with time at the same rate as ∇2x2
0.
In order to estimate the rate of increase of ∇2x2
0, we generated random sequences of µn and
calculated r0 from rN according to eqn. (44). We used the result to calculate ∇2x2
0 as a function of
n, or equivalently, as a function of time, since n corresponds to the time nτ . The only adjustable
parameter in these calculations is γτ , which measures the coherence of the flow, here the renewal
interval in units of the stretching rate. We expect γτ to be O(1).
Figure (1) shows the dimensionless diffusion rate as a function of iteration number in the case
γτ = 0.5. On average, the increase in diffusion rate is well fit by an exponential, and is about 84%
Fig. 1. -- The natural log of the increase in diffusion rate, or amplification factor, with time, or the
number of iterations. Each solid curve is the average of 50 independent realizations of the iteration
process. The dashed line, which has slope 0.838, is the exponential with the same final value as
the average of the curves. The maximum amplification factor possible would occur if all stagnation
points had inflow along the x axis, and would have a slope of 1 in these units.
the rate of increase for coherent stagnation point flow given in eqn. (28). After 20 renewals the
average diffusion rate is more than 107 times larger than its original value. This is much larger than
the 2-3 orders of magnitude that we estimated in §2.2 as required to explain the B − n relation.
The mean amplification rate is relatively insensitive to the coherence parameter γτ , being about
75% of maximum if γτ = 0.1 and about 87% of maximum if γτ = 1.
However, there is substantial dispersion about the mean. Each solid curve represents 50
-- 12 --
independent sequences of iterations, and Fig. (1) shows differences between them. The standard
deviation within each set of 50 sequences is typically about 40% of the mean, and the amplification
factors for single sequences rarely grow exponentially. This is illustrated in Figure (2). The large
Fig. 2. -- The natural log of the increase in diffusion rate, or amplification factor, with time, or the
number of iterations. Each solid curve is a single independent realization of the iteration process.
The curves, which correspond to the first 5 members of a larger ensemble, show the intrinsic
variability of the amplification process.
standard deviation suggests that the diffusion rate in this model is highly intermittent, which
is characteristic of turbulent mixing. Additional evidence of intermittency is seen in the PDFs
of the distribution of amplification factors, shown in Figure (3) for two different values of γτ .
The maximum possible amplification rate of 20 imposes a cutoff on the high amplification side of
the curve for γτ = 1.0; the PDF for the case γτ = 0.1 is more symmetrical because the mean
amplification rate is well below the maximum.
4.3. The Stagnation Point Model and Turbulent Flow
The hyperbolic stagnation point model achieves fast diffusion by increasing the magnetic field
gradient at, on average, nearly exponential rates. Chaotic flow achieves fast stretching and shrinking
without hyperbolic stagnation points because the trajectories of neighboring fluid particles separate
at an exponential rate. Small differences in the rates of exponential change lead to highly intermit-
tent distributions of scalar quantities, which we saw reflected in the stagnation point model through
-- 13 --
Fig. 3. -- Normalized frequency distributions, or PDFs, of the natural logarithms of the amplifi-
cation factors for 10 iterations at γτ = 1.0 (solid curve) and 100 iterations at γτ = 0.1 (dashed
curve), so that the curves correspond to the same total time. Each curve is based on 104 random
sequences of iterations.
the wide dispersion of amplification rates (see Figures 2 & 3). The maximum rate of stretching of
a fluid element labelled by its initial position x0 is given by the Lyapunov exponent Λ(x0)
Λ (x0) ≡ max
e0
lim sup
t→∞
1
t
ln
∂x
∂x0 · e0 ,
(45)
where the e0 are the set of all possible unit vectors (see Childress & Gilbert 1995).
In order to make it plausible that the diffusion rate is enhanced by exponential shrinking and
stretching in a flow, we imagine that the diffusivity λ is so small that we can ignore it. In this
limit, the solution of the advection equation for the initial condition (15) is
B (x, t) = B (x0, 0) = B00 +
1
2
B ′′
0 x2
0 (x) .
If we restore diffusion, it appears in the advection-diffusion equation (7) as the term
(46)
(47)
λB ′′
0∇2 x2
0
2
= λB ′′
0(cid:0)x0∇2x0 + ∇x0 · ∇x0(cid:1) .
The quantity ∇x0 · ∇x0 on the right hand side of eqn. (47) is the square of the inverse of the
stretching rate. This suggests heuristically that the diffusion rate grows exponentially.
-- 14 --
Examples of 2D, chaotic, spatially periodic flow with a single lengthscale, including maps of
Lyapunov exponents and other measures of chaos, are given by Galloway & Proctor (1992), Ponty
et al. (1993) and Cattaneo et al. (1995).
5. Dynamical Feedback with Extension to 3D
In §4, we prescribed a strictly 2D flow, and neglected feedback by Lorentz forces.
In fact,
by virtue of eqn. (1), the diffusion rate cannot be enhanced without increasing the Lorentz force,
while 3D effects modify the diffusion process itself (compare eqns. (3) and (6)). In this section we
quantify the effects of magnetic feedback and derive criteria for the validity of the 2D model.
We assume the stagnation point flow given by eqns. (28) with a = 0. The magnetic field can
be written using eqns. (32), (34), and (37) as
B = B00(cid:20)1 −
1
4γtd0 (cid:0)e2γt − 1(cid:1) −
x2
4L2 e2γt(cid:21) .
(48)
We evaluate the importance of feedback by following the force on a fluid element over the mixing
time tγ. This probably overestimates the effects of magnetic feedback, because the coherence time
τ of any particular realization of the flow is expected to be less than tγ. Thus, the constraints on
the turbulence which we derive are likely to be conservative. In fact, mixing appears to take place
in fully self consistent models of Alfven wave turbulence; Maron & Goldreich (2001).
In what follows, it is useful to remember that the parameter 2γtd0, the ratio of the classical
ambipolar diffusion time to the eddy turnover time, is large; see eqn. (39). We will sometimes use
the inverse of this quantity as an expansion parameter.
5.1. Magnetic Pressure Forces
We estimate the deceleration of an element of fluid by magnetic pressure in time tγ. The x
component of magnetic pressure force Fm is
Fm = −x
∂
∂x
B 2
8π
.
The deceleration ∆vP of a fluid element over a time t is
∆vP (x0) =
dsFm (x(s), s) ,
1
ρZ t
0
(49)
(50)
where x(s) = x0e−γs is the position of the fluid element at time s and x0 is its original position.
Substituting eqn. (48) into eqn. (49) and integrating eqn. (50) to t = tγ yields to leading order in
2γtd0
∆vP (x0) = −
B 2
00x0
16πρL2γ
(2γtd0)1/2 ,
(51)
-- 15 --
where we have assumed x0/L ≪ 1.
The average velocity ¯v(x0) of the fluid element over this time is
¯v (x0) ≡
1
tγ Z tγ
0
dsv(x(s), s),
(52)
where x(s) is once again the Lagrangian position of a fluid element. For the stagnation point flow
(28),
¯v =
,
(53)
x0
tγ
to leading order in 2γtd0. Magnetic feedback on the flow is unimportant if ∆vP /¯v < 1. Combining
eqns. (51) and (53), we derive
∆vP
¯v
=
B 2
00
32πρL2γ 2 (2γtd0)1/2 ln (2γtd0) .
(54)
Using eqns. (12) and (39), eqn. ( 54) can be rewritten as
Equation (55) implies an upper limit τ P
∆vP
¯v
=
τd
τ 2
A
ln(cid:18)2
32τni(cid:19)1/2
τdτni(cid:19) .
τA (cid:18) τd
max on the eddy turnover time τd such that ∆vP /¯v ≤ 1;
τA(cid:19)1/3(cid:20)ln(cid:18) τA√2τni(cid:19)(cid:21)−2/3
τni
,
τ P
max
τA
=(cid:18)18
(55)
(56)
where, to sufficient accuracy, we have replaced τd by (32τ 2
timescale the magnetic field is still well frozen to the eddies; from eqns. (12) and (56),
Aτni)1/3 in the logarithmic factor. On this
RAD(cid:0)τ P
max(cid:1) =
1/3
v2
t
v2
A (cid:18)18
τ 2
A
τ 2
ni(cid:19)
(cid:20)ln(cid:18) τA√2τni(cid:19)(cid:21)−2/3
.
(57)
5.2.
3D Effects
We assume the turbulent motions are in the (x, y) plane, but depend weakly on z; i.e. the
characteristic wavenumber k along the field is related to the turbulent lengthscale l by kl ≪ 1.
This quasi-two dimensionality is expected to be a feature of Alfvenic turbulence in strong magnetic
fields (Strauss 1976, Goldreich & Sridhar 1997).
Since the field is fairly well frozen in even on scales l, it is very well frozen on the scale k−1,
and the transverse field B⊥ is given to a good approximation by
B⊥ = Bz(x0, t)
∂x⊥
∂z
,
(58)
-- 16 --
where x0 is the initial position of the fluid element at position x at time t.
Let us introduce a small parameter ǫ and assume that ∂z is O(ǫ) relative to the perpendicular
derivatives, and that B⊥/Bz is also O(ǫ) (this is the so-called reduced MHD ordering; Strauss 1976).
It can then be shown that the changes in the ambipolar drift terms are of order ǫ2. Therefore, we
may assume that weak three dimensionality has little effect on ambipolar drift of the vertical field.
The bent field exerts a tension force which decelerates the fluid by an amount ∆vT . We
compute ∆vT for a z dependent stagnation point flow model with
vx = −γx cos kz,
vy = γy cos kz,
a generalization of eqns. (28). The Lagrangian positions are
x = x0e−γt cos kz,
y = y0eγt cos kz.
According to eqns. (58) and (61), the x component of the field is
Bx = Bzγtkx sin kz.
The magnetic tension force Fm in the x direction is
B 2
z
4π
∂Bx
∂z
Fm =
Bz
4π
=
γtk2x cos kz.
(59)
(60)
(61)
(62)
(63)
(64)
We set z = 0 and follow a procedure similar to the derivation of eqn. (55), integrating Fm along
the path of a fluid element from t = 0 to t = tγ. We approximate Bz by B00, which overestimates
Fm. The result to leading order in (2γtd0)−1 is
Using eqn. (53), the relative deceleration is ∆vT /¯v
1
ρZ tγ
0
dtFm =
k2v2
Ax0
γ
.
∆vT =
∆vT
¯v
=
k2v2
A
2γ 2 ln (2γtd0) .
(65)
(66)
Equation (66) shows that magnetic tension has little effect on the fluid as long as the Alfven
frequency along the fieldline is less than the eddy turnover rate by the factor [ln (2γtd0)]1/2.
Equation (66) can be used to set an upper limit τ T
max on the eddy turnover time such that
the magnetic field reaches the mixing scale without decelerating the fluid. Proceeding as in the
derivation of eqn. (56) we find
τ T
max
τA
=
√2
kL (cid:20)ln(cid:18)√2kL
τA
τni(cid:19)(cid:21)−1/2
.
(67)
Equation (67) shows that tension forces are less important in long, thin structures, in which kL
can be much less than one, than they are in flattened structures such as disks. The field is well
frozen to the eddies as long as τA/τni ≫ 1.
-- 17 --
6. Application to the Galactic Magnetic Field
Little is know about interstellar turbulence beyond its gross energetics: the turbulent kinetic
energy is at or above equipartition with the magnetic energy. Energy injection by a variety of
mechanisms, combined with nonlinear processes, should lead to a turbulent spectrum over a wide
range in scales.
In principle, any weakly ionized interstellar structure which survives for several eddy turnover
times is a candidate for turbulent ambipolar diffusion. We showed in §5 that the efficiency of turbu-
lent ambipolar drift can be limited by the back reaction of magnetic forces. In the strictly 2D case,
the increase in ambipolar diffusion rate is associated with the local buildup of magnetic pressure
forces, and in the 3D case, by magnetic tension forces as well. We expressed these constraints in
terms of lower bounds on the strain rates, or inverse eddy turnover times, such that concentration
of the field occurs before deceleration of the flow. These constraints appear in eqns. (56) and (67).
Here, we express them numerically.
The critical parameters are the ratio τni/τA and the geometrical factor kL. Referring back to
§2.2 for numerical values, we have
τni
τA
= 4.8 × 10−5
Bµ
Lpcni (nnAn)1/2 ,
(68)
where L is expressed in parsecs. For example, if nn = 50 cm−3, ni = 5 × 10−3 cm−3, An = 1.4, Lpc =
1, Bµ = 3, τni/τA = 3.4 × 10−3, and eqn. (56) requires τd/τA < 0.13. If vt ∼ vA, turbulence on the
scale of a tenth of a parsec or less can mix the magnetic field down to the ambipolar diffusion scale.
On the other hand, eqn. (67) requires τd/τA < 0.09(2π/kL), which is a more severe constraint,
especially in a highly flattened structure.
There is at least one type of H I structure in which flux freezing appears to be obeyed. Mag-
netic fields in H I shells are observed to be quite strong, with magnitudes consistent with shock
compression (Heiles 1989). The same observations suggest that turbulence with Alfvenic or slightly
subAlfvenic velocities is present. If these shells were not expanding, they would appear to fulfill the
conditions for fast ambipolar drift, and their strong fields would be counterexamples to the theory.
However, expansion of the shells at speeds of order 10 - 20 km/s adds new magnetic flux faster
than it can diffuse upstream, while on the downstream side the ionization is too high for efficient
ambipolar drift. Thus, the field in the shells remains large. Recent observations by Heiles (2001b)
and Heiles & Troland (2001) of cold, moderately dense H I regions which have weak magnetic fields,
Alfvenic random velocities, and no observed association with shells appear to be better candidates
for fast ambipolar drift.
-- 18 --
7. Summary and Conclusions
Observations show that while interstellar magnetic fieldstrength and gas density are to some
extent correlated, the fieldstrength is often lower than expected. This suggests that processes
beyond ideal MHD may play a role.
The flux to mass ratio is altered by ambipolar drift, but estimates of the ambipolar diffusivity
v2
Aτni predict that ambipolar drift is important only in very dense, strongly magnetized gas with
substantial gradients on small scales. However, it is well established that turbulence can enhance
the transport rates of quantities such as entropy and angular momentum. This motivated us to
consider the effect of turbulence on the rate of ambipolar drift. Enhancement by roughly two orders
of magnitude would explain the observations.
As a first attempt on the problem, we considered the geometrically restricted case of a straight
magnetic field, with a transverse gradient, mixed by 2D, perpendicular turbulence. In this situation,
ambipolar drift is described by a nonlinear diffusion term [eqn. (6)]. For simplicity, however, we
approximated the diffusivity as linear.
We showed by a mixing length argument, and then an explicit calculation (§3) that advection
of the field by a periodic flow reduces its peak value. However, unless the motions are long and
thin, like streamers, the rate of relaxation is no faster than relaxation by ambipolar drift alone.
The missing ingredient is stretching and shrinking of scales, which in chaotic flows happens at
an exponential rate.
In §4 we modelled these exponential changes of scale by representing the
flow as a sequence of randomly oriented hyperbolic stagnation points. With this model, and a
parabolic initial condition for the magnetic profile, the advection-diffusion equation (7) can be
solved exactly. The model predicts an exponential increase of the mean diffusion rate with time
(Figure 1), although with considerable variance from point to point (Figures 2 & 3). The stagnation
point model predicts that the field diffuses on a timescale comparable to the eddy turnover time,
with only logarithmic dependence on the ambipolar diffusivity itself and on the original gradient
lengthscale.
In §5, we estimated the back reaction of magnetic pressure and tension forces on the stagnation
point flow, including weak three dimensionality. The relative deceleration of the fluid over one
mixing time is given for pressure forces by eqn. (54) and for tension forces by eqn. (66). Comparing
the deceleration time of a fluid element to the mixing time, we derived upper limits on the eddy
turnover time such that deceleration is order unity or less within a mixing time [eqns. (56) and
(67)]. As we showed in §6, these criteria can be satisfied in the interstellar medium without extreme
assumptions about the size and velocity of the turbulent eddies, although they cannot be wholly
ignored. These estimates of feedback are conservative in the sense that a fully self consistent
model of MHD turbulence can still produce fast mixing, as shown by Maron & Goldreich (2001)
for spreading of a passive scalar. This conclusion may be dependent on geometry, however, as
concluded by Kim (1997) based on a study of turbulent decay of a coplanar magnetic field caused
by 2D incompressible motions in weakly ionized fluid.
-- 19 --
The outcome of these calculations is that turbulence is likely to have a major effect on the
magnetic flux to mass ratio in the weakly ionized portions of the interstellar medium, making the
magnetic field more uniform. The model presented here applies to regions with simple magnetic
topology, filamentary structure, and no global cross-field flows. It follows that the strength of the
field is not necessarily a good indicator of the dynamical processes which determine the gas density.
Although the results presented here are consistent with the conventional wisdom that turbulent
mixing takes place in an eddy turnover time, it is important to recognize that they are not obtained
from a full model of turbulence. We plan to extend this work to more realistic models which include
compressibility, vorticity, and a self consistent treatment of magnetic forces as well as more general
magnetic geometry.
If the stagnation point model holds up in comparison with more complete
models, it could be useful in other mixing problems.
Magnetic reconnection can also change the magnetic flux to mass ratio. Lazarian & Vishniac
(1999) have argued that if the spectrum of interstellar turbulence extends to the resistive scale
then reconnection takes place at the Alfven speed. We have concentrated here on ambipolar drift
because it does not require turbulent structure on such small scales; if flux is quickly redistributed
in the fully ionized portions of the ISM then an alternative process is certainly required.
This work was initiated during the program on Astrophysical Turbulence at the ITP in Santa
Barbara in 2000. I am happy to acknowledge useful discussions with Nic Brummell, Dick Crutcher,
George Field, Carl Heiles, Fabian Heitsch, David Hughes, Eun-jin Kim, Steve Tobias, Tom Troland,
and especially the referee, Pat Diamond. Material support was provided by NSF Grants AST
9800616 and AST 0098701 to the University of Colorado and PHY 9407194 to UC Santa Barbara.
Bourke, T.L., Myers, P.C., Robinson, G., & Hyland, A.R. 2001, ApJ, in press, and astro-ph 0102469
REFERENCES
Brandenburg, A., & Zweibel, E.G. 1994, ApJ, 427, L91
Brandenburg, A., & Zweibel, E.G. 1995, ApJ, 448, 734
Brogan, C. & Troland, T. 2001, ApJ, 550, 799
Cattaneo, F. 1994, ApJ, 434, 200
Cattaneo, F., Hughes, D.W., & Kim, E. 1996, Phys. Rev. Lett., 76, 2057
Cattaneo, F., Kim, E., Proctor, M.R.E., & Tao, L. 1995, Phys. Rev. Lett., 75, 1522
Cattaneo, F., & Vainshtein, S. 1991, ApJ, 376, L21
Childress, S. & Gilbert, A.D. 1995, Stretch, Twist, Fold: The Fast Dynamo, Springer
-- 20 --
Crutcher, R.M. 1999, ApJ, 520, 706
Draine, B.T., Roberge, W.R., & Dalgarno, A. 1983, ApJ, 264, 485
Galloway, D.J. & Proctor, M.R.E. 1992, Nature, 356, 691
Goldreich, P.M., & Sridhar, S. 1997, ApJ, 485, 680
Gruzinov, A., & Diamond, P.H. 1994, Phys. Rev. Lett., 72, 1651
Heiles, C. 1989, ApJ, 336, 808
Heiles, C. 2001a, to be published in Proceedings of the 4th Tetons Summer Conference, and astro-ph
0010047
Heiles, C. 2001b, ApJ, 551, L105
Heiles, C., & Troland, T. 2001, in preparation
Kim, E. 1997, ApJ, 477, 183
Lazarian, A., & Vishniac, E.T. 1999, ApJ, 517, 700
Mestel, L. 1985, in Protostars & Planets II, eds. D.C. Black & M.S. Matthews, Univ. Arizona Press,
p. 320
Mestel, L., & Spitzer, L.S. 1956, MNRAS, 116, 505
Moffatt, H.K. 1978, Magnetic Field Generation in Electrically Conducting Fluids, Cambridge Univ.
Press
Mouschovias, T. 1976, ApJ, 207, 141
Ottino, J.M. 1989, The Kinematics of Mixing, Cambridge Univ. Press
Padoan, P. & Nordlund, A. 1999, ApJ, 526, 279
Ponty, Y.,Pouquet, A., Rom-Kedar, V., & Sulem, P.-L. 1993, in Solar & Planetary Dynamos, eds.
M. Proctor, P. Matthews, & A. Rucklidge, Cambridge Univ. Press, p. 241
Press, W.H. & Rybicki, G. 1981, ApJ, 248, 751
Shraiman, B.I. & Siggia, E.D. 2000, Nature, 405, 639
Shu, F.H. 1983, ApJ, 273, 202
Sridhar, S., & Goldreich, P.M. 1994, ApJ, 432, 612
Strauss, H.R. 1976, Phys. Fluids, 19, 134
-- 21 --
Tomisaka, K., Ikeuchi, C., & Nakamura, T. 1988, ApJ, 335, 239
Troland, T.H. & Heiles, C. 1986, ApJ, 301, 339
Zel'dovich, Y.B.& Raizer, Y.P. 1966, Physics of Shock Waves & High Temperature Hydrodynamic
Phenomena, Academic Press
Zel'dovich, Y.B., Ruzmaikin, A.A., Molchanov, S.A., & Sokoloff, D.D. 1984, J. Fluid Mech., 144, 1
Zweibel, E.G. 1988, ApJ, 329, 384
Zweibel, E.G. 1998, Phys. Plasmas, 5, 247
Zweibel, E.G. & Brandenburg, A. 1997, ApJ, 478, 563
This preprint was prepared with the AAS LATEX macros v5.0.
|
0801.0541 | 1 | 0801 | 2008-01-03T15:16:59 | Doppler Images and Chromospheric Variability of TWA 6 | [
"astro-ph"
] | We present Doppler imaging and Balmer line analysis of the weak-line T Tauri star TWA 6. Using this data we have made one of the first attempts to measure differential rotation in a T Tauri star, and the first detection of a slingshot prominence in such a star. We also show the most direct evidence to date of the existence of solar-type plages in a star other than the Sun.
Observations were made over six nights: 11-13th February 2006 and 18-20th February 2006, when spectra were taken with the UCL Echelle Spectrograph on the 3.9-m Anglo-Australian Telescope. Using least-squares deconvolution to improve the effective signal--to--noise ratio we produced two Doppler maps. These show similar features to maps of other rapidly rotating T Tauri stars, i.e. a polar spot with more spots extending out of it down to equator. Comparison of the two maps was carried out to measure the differential rotation. Cross-correlation and parameter fitting indicates that TWA 6 does not have detectable differential rotation.
The Balmer emission of the star was studied. The mean H-alpha profile has a narrow component consistent with rotational broadening and a broad component extending out to 250km/s. The variability in H-alpha suggests that the chromosphere has active regions that are cospatial with the spots in the photosphere, similar to the 'plages' observed on the Sun. In addition the star has at least one slingshot prominence 3 stellar radii above the surface - the first such detection in a T Tauri star. | astro-ph | astro-ph | Mon. Not. R. Astron. Soc. 000, 1–13 (2007)
Printed 12 August 2017
(MN LATEX style file v2.2)
Doppler Images and Chromospheric Variability of TWA 6
M. B. Skelly,1⋆ Y. C. Unruh,1 A. Collier Cameron,2 J. R. Barnes,3 J.-F. Donati,4
W. A. Lawson,5 B. D. Carter6
1 Astrophysics Group, Imperial College of Science, Technology and Medicine, London SW7 2AZ
2 School of Physics and Astronomy, University of St Andrews, Fife KY16 9SS
3 Centre for Astrophysics Research, University of Hertfordshire, College Lane, Hatfield, Herts AL10 9AB
4 Observatoire Midi-Pyr´en´ees, LATT, 14 avenue Edouard Belin, 31400 Toulouse, France
5 School of Physical, Environmental and Mathematical Sciences, University of New South Wales,
Australian Defence Force Academy, Canberra, ACT 2600, Australia
6 Faculty of Sciences, University of Southern Queensland, Toowoomba, 4350, Australia
3 January 2008
ABSTRACT
We present Doppler imaging and Balmer line analysis of the weak-line T Tauri star
TWA 6. Using this data we have made one of the first attempts to measure differential
rotation in a T Tauri star, and the first detection of a slingshot prominence in such
a star. We also show the most direct evidence to date of the existence of solar-type
plages in a star other than the Sun.
Observations were made over six nights: 11–13th February 2006 and 18–20th
February 2006, when spectra were taken with the UCL Echelle Spectrograph on the
3.9-m Anglo–Australian Telescope. Using least-squares deconvolution to improve the
effective signal–to–noise ratio we produced two Doppler maps. These show similar fea-
tures to maps of other rapidly rotating T Tauri stars, i.e. a polar spot with more spots
extending out of it down to equator. Comparison of the two maps was carried out
to measure the differential rotation. Cross-correlation and parameter fitting indicates
that TWA 6 does not have detectable differential rotation.
The Balmer emission of the star was studied. The mean Hα profile has a narrow
component consistent with rotational broadening and a broad component extending
out to ±250 km s−1. The variability in Hα suggests that the chromosphere has active
regions that are cospatial with the spots in the photosphere, similar to the 'plages'
observed on the Sun. In addition the star has at least one slingshot prominence 3R∗
above the surface – the first such detection in a T Tauri star.
Key words: stars: pre-main-sequence – stars: late-type – stars: rotation – stars: spots
– stars: individual: TWA 6 – stars: magnetic fields
8
0
0
2
n
a
J
3
]
h
p
-
o
r
t
s
a
[
1
v
1
4
5
0
.
1
0
8
0
:
v
i
X
r
a
1
INTRODUCTION
1.1 T Tauri Stars
T Tauri stars are pre-main sequence equivalents of Sun-like
stars. They are young, have low masses (0.3 – 1.0 M⊙) and
spectral types F–M (Hayashi 1966). The T Tauri phase is
considered to begin when the star becomes optically visible
and continues until hydrogen burning begins and the star
has reached the post-T Tauri phase at an age of the order
of 107 years (depending on spectral type).
fall on Hayashi
tracks on the
Hertzsprung-Russell diagram, i.e. vertical tracks of constant
temperature. As the star approaches the zero-age main se-
Very young stars
⋆ E-mail: [email protected]
quence (ZAMS) it can either continue to follow the Hayashi
track all the way onto the ZAMS (and remain fully convec-
tive), or develop a radiative core and turn onto the Henyey
track and evolve onto the ZAMS with constant luminos-
ity and increasing temperature. Which of these evolutionary
paths the T Tauri star follows is thought to depend on the
mass of the star but observations of TTS with ages close to
this turn-off will allow us to characterise this stage of stellar
evolution more fully.
T Tauri stars fall
into two subcategories – classical
and weak-line (Bertout 1989 and references therein). Ob-
servations of classical TTS (CTTS) show excess ultravio-
let and infrared emission and strong emission line activity,
characteristics which are due to the presence of a circum-
stellar disc. The star–disc interactions lead to strong, ir-
regular variability that is predominately due to accretion
2 M. B. Skelly et al
of material from the disc onto the surface. The disc has a
braking effect on the star, causing CTTS to rotate more
slowly than their discless equivalents (Bouvier et al. 1993;
Collier Cameron & Campbell 1993).
Weak-line T Tauri stars (WTTS) have weaker emission-
line activity and show less evidence of accretion. WTTS ap-
pear to be older than CTTS on average (Armitage et al.
2003), suggesting that when TTS form they have circum-
stellar discs which dissipate as the star evolves, but disc life-
times are variable, ranging from 0.1 – 10 Myr (Bouvier et al.
1997). WTTS are variable, but in contrast to CTTS the vari-
ation is usually periodic as it is due to cool spots that go in
and out of view as the star rotates. These cool spots form
when magnetic flux tubes emerge through the photosphere.
Doppler imaging (Vogt & Penrod 1983) maps spots
on stellar photospheres. Doppler imaging of rapidly rotat-
ing, magnetically active stars, such as WTTS, often finds
spots at high latitudes and even at the poles, see e.g.
Rice & Strassmeier (1996), Strassmeier (2002), in contrast
to the Sun where they are normally within ±30◦ of the equa-
tor. One of the favoured explanations is that Coriolis forces
in rapidly rotating stars cause the flux tubes to emerge at
higher latitudes (Schuessler & Solanki 1992; Caligari et al.
1994).
In the past Doppler imaging of cool stars has been
limited by its challenging signal-to-noise requirements. The
technique of least-squares deconvolution (LSD, Donati et al.
1997) has made it possible to produce Doppler images of
stars for which the signal-to-noise ratio (SNR) would be too
low if individual spectral lines were used. In LSD the under-
lying rotationally broadened line profile for each spectrum
is found by deconvolving a line list from the spectrum. This
makes it possible for late–K T Tauri stars to be imaged,
where previously the SNR requirements would have been
prohibitive.
Sunspots are accompanied by prominences, i.e. loops
of material that follow the magnetic field lines emerging
through the surface of the star. High–lying prominences,
known as 'slingshot prominences', typically appear as fast–
moving absorption transients and have been observed in a
number of young stars such as Speedy Mic (Dunstone et al.
2006) and AB Dor (Collier Cameron & Robinson 1989;
Donati et al. 1999), but not in any TTS to date. These
prominences are viewed as a tracer for large-scale magnetic
fields. If a T Tauri star was shown to be able to support
such high-lying prominences for several rotational periods it
would suggest the presence of a large-scale field that does
not decay away quickly outside the stellar radius.
1.2 Differential Rotation in Stars with Deep
Convection Zones
Fully convective stars are believed to have a distributed mag-
netic dynamo (α2), arising from turbulent motions or Corio-
lis forces, rather than a solar-type dynamo (αΩ) that is due
to both convection and differential rotation, as seen in stars
with radiative cores.
Differential rotation, when rotational velocity is a func-
tion of latitude, is expected to be less apparent with increas-
ing convection zone depth (Kitchatinov & Rudiger 1999), so
stars with larger convection zones are expected to rotate
more like a solid body. Differential rotation in the Sun gives
rise to the Ω effect that produces a large-scale toroidal field
from the small-scale poloidal field. Most models for an α2 dy-
namo rule out the possibility of there being differential rota-
tion in fully convective stars (e.g. Chabrier & Kuker 2006).
However Dobler et al. (2006) suggest fully convective stars
have significant differential rotation, although this would not
be Sun-like, and takes the form of a meridional circulation.
Observations to date have only attempted to detect
the latitudinal differential rotation seen in the Sun. A re-
cent measurement for the fully-convective M4 dwarf V374
Peg (Morin et al. 2007) found a weak differential rotation
(around 0.1 of the solar value). Differential rotation has been
observed to fall with decreasing photospheric temperature.
For example, figure 2 in Barnes et al. (2005a) plots differen-
tial rotation against temperature. The trend seems to imply
that differential rotation approaches zero below 4000K but it
is uncertain where it becomes undetectable as there has been
no measurement to date of stars with temperatures between
3800–4500K. In addition there have not been any reliable
measurements of differential rotation in T Tauri stars. Such
measurements are needed to confirm that TTS follow these
trends too.
1.3 TW Hydrae Association
Nearby young co-moving groups are a rich source of T Tauri
stars for observation. Our chosen targets lie in the TW Hy-
drae association (TWA), a group of pre-main sequence stars
lying at a distance of about 51 ± 5pc and with an age of
8 Myr (Zuckerman & Song 2004). Of the 24 likely mem-
bers of the TWA, 16 have had their periods measured in
Lawson & Crause (2005). The age of this association is an
important one as it is corresponds to the mean age of cir-
cumstellar discs and hence the time when T Tauri stars can
begin to spin up, so the members of this association are inter-
esting objects for study. The star that gives the association
its name (TW Hya) is a classical T Tauri star, spectral type
K7. Its disc has been directly imaged in Krist et al. (2000)
and its magnetic field was detected in Yang et al. (2007).
Although it is rotating fairly rapidly (P = 2.8d) it is almost
pole–on and is therefore not a suitable candidate for Doppler
imaging.
This paper concerns another member of the TW Hy-
drae association – TWA 6. It has a spectral type of K7
(Webb et al. 1999), corresponding to a temperature of about
4000K for a T Tauri star (Cohen & Kuhi 1979). Models
by Siess et al. (2000) indicate that it is fully convective or
just beginning to develop a radiative core. Photometry has
revealed a large lightcurve amplitude, with a variation of
0.49 magnitudes in the V band (Lawson & Crause 2005). Its
rapid rotation, v sin i = 55 km s−1 (Webb et al. 1999) and
period of 0.54 days (Lawson & Crause 2005) make it an ex-
cellent target for Doppler imaging. In this paper we present
Doppler images of TWA 6, and a discussion of whether it has
differential rotation. We also analyse the Balmer line emis-
sion in order to study the behaviour of the chromosphere.
2 OBSERVATIONS AND DATA REDUCTION
Observations were carried out at the Anglo-Australian Tele-
scope (AAT) using the University College London Echelle
Doppler Images and Chromospheric Variability of TWA 6
3
Spectrograph (UCLES) and the EEV2 2k × 4k detector be-
tween 11–20th February 2006. In order to detect differential
rotation observations were made with a gap of 4 nights. By
doing this two maps can be made of the star allowing differ-
ential rotation to be detected, or an upper limit placed on its
value (Petit et al. 2004). With three nights of observations
on either side of this gap full phase coverage was achievable
on these two separate occasions.
We took spectra of the targets, telluric and spectral
standard stars, and calibration frames (see tables 1 and 2).
UCLES was centred at 5500A giving a wavelength coverage
of 4400 − 7200A. A slit width of 1.0′′ was used giving a
spectral resolution of 45000.
The weather on the first three nights was reasonably
good, and although conditions and seeing were variable (see-
ing ranged from 1.2–2.2′′) in total only 3 hrs were lost. 36
good quality spectra of TWA 6 were obtained, as well as
27 spectra of our secondary target TWA 17 (Skelly et al, in
prep.). On the second set of nights the weather was of vari-
able quality and more time was lost. Of a potential 24 hours
we had 13 hours of observing time, with seeing between 1.3
– 2.2′′. Most of this time were concentrated on TWA 6 and
a further 39 spectra were obtained. An additional 1.5hr of
observing time was obtained on 17th Feb 2006 due to the
early finish of aluminising the telescope mirror.
2.1 Data Reduction
Data reduction was carried out with the STARLINK rou-
tine ECHOMOP (Mills et al. 1997). Frames were bias sub-
tracted and flatfielded, cosmic rays were removed and the
spectra were extracted. Sky subtraction was carried out by
subtracting a polynomial fitted to the inter-order gaps. The
spectra were normalised using a continuum fit to a star of
the same spectral type. The spectral type (K7) was verified
by eye by identifying the best temperature fit to the spec-
tra in comparison with the template spectra. Wavelength
calibration was carried out using ThAr arc spectra.
3 LEAST-SQUARES DECONVOLUTION
As the 36+39 spectra did not have a high enough SNR for
Doppler imaging we used least-squares deconvolution (LSD,
described in Donati et al. 1997) to extract the underlying ro-
tationally broadened profiles. LSD uses matrix inversion to
deconvolve a line list from a spectrum to leave a high signal-
to-noise profile. The program SPDECON (implemented as
described in Barnes et al. 1998) was used to do this. We used
a line list from the Vienna Atomic Line Database (VALD,
Kupka et al. 2000) for a 4000K star. Over 4000 images of
2000 spectral lines were used in the deconvolution leading
to a signal gain of ∼ 60. For the deconvolution, the lines
were weighted by counts2/variance. The effect of this is to
favour regions of the spectrum with higher SNR, and the
centres of the orders rather than the wings. The resultant
line profiles had a mean centroidal wavelength of 5530A.
There was a small consistent tilt in the profiles which
was removed by fitting a straight line between the two sec-
tions of continuum on the left and right sides. This tilt
changed depending on which spectral standard was used and
hence is most likely due to a small mismatch between the
Figure 1. The dashed line shows the mean deconvolved profile
for the first dataset (11–13th Feb 2006). There is an absorption
feature due to sunlight reflected off the Moon at 0 km s−1. The
solid line is the mean of the profiles after this feature was removed
and a slope was removed. Mean errors are shown for the data after
the feature was removed. The errors close to 0 km s−1have been
increased due to the subtraction.
temperatures of the star and the spectral standards. In ad-
dition there was a feature near 0 km s−1 in many of the
profiles. It was identified as being due to moonlight in the
spectra as it only appeared to be present when the Moon
was above the horizon (Fig 1).
As we did not have a solar spectrum we deconvolved the
hottest template spectrum (K3) we had taken with UCLES
to model the shape of this feature. We were unable to find
a pattern to its depth as it is difficult to disentangle from
variations due to surface features. We failed to find a re-
lationship between its equivalent width and the zenith dis-
tance, time or seeing, although the absorption signal was
stronger during bad weather. Hence for all spectra taken
when the Moon was up, a constant shape was subtracted
from all profiles taken within the same night (see Fig. 1).
This was calculated by inspecting the mean of all profiles,
and subtracting a profile such that the absorption feature
was not obvious in the mean. A scale factor was calculated
separately for each night. The errors for the pixels affected
were increased so that these pixels would have a lower influ-
ence on the final image produced. This was done by adding
in quadrature the error produced by SPDECON and the
amount that was subtracted from each pixel.
Each of the deconvolved profiles has been normalised
by dividing through by the mean profile. In Fig. 2 the nor-
malised profiles have been stacked and shown in greyscale.
As spots appear in emission the lighter regions in these im-
ages show the spots being carried across the stellar disc. Spot
groups are apparent at the same phases in both, particularly
around phases 0.05 and 0.55, corresponding to longitudes of
340◦ and 160◦. This suggests that the spot distribution is
similar at both epochs.
4 M. B. Skelly et al
Table 1. Table setting out the observations made on the nights of 11/02/06 to 13/02/06.
Spectral types in the 'Comments' column refer to spectral standards. The signal-to-noise
ratio is the maximum for the set of exposures.
Object
Date UT Start
Exp.
time [s]
No. of
Frames
S/N Comments Conditions
HR 3037
HR 3037
HD 34673
TWA 6
HR 3037
HD 52919
TWA 6
SAO 116640
TWA 6
TWA 6
TWA 17
TWA 6
GJ 204
GJ 3331A
TWA 6
TWA 6
HD 19007
TWA 17
HR 3037
TWA 17
TWA 17
HR 3037
TWA 17
HR 3037
TWA 17
HR 3037
HD 52919
SAO 116640
GJ 3331A
TWA 6
GJ 204
TWA 17
TWA 6
TWA 17
TWA 6
TWA 17
TWA 6
TWA 17
TWA 6
TWA 17
11
11
11
11
11
11
11
11
11
11
11
11
12
12
12
12
12
12
12
12
12
12
12
12
12
13
13
13
13
13
13
13
13
13
13
13
13
13
13
13
09:42:56
09:46:15
09:58:05
10:25:43
12:10:19
12:16:17
12:22:44
13:45:02
13:55:31
15:17:37
16:43:21
18:01:59
09:40:18
10:01:49
10:23:33
10:42:47
11:04:33
12:42:06
13:06:28
13:13:37
14:06:49
15:26:56
15:42:30
16:48:51
16:55:16
09:32:42
09:35:50
09:40:57
09:50:39
10:09:24
12:36:26
12:44:30
13:31:51
13:49:34
14:33:20
15:01:47
15:35:40
15:54:08
17:22:43
17:40:59
20
120
200
900
20
200
900
450
900
900
1200
900
120
900
900
900
200
1200
20
1200
1200
20
1200
20
1200
20
200
400
900
900
120
1200
900
1200
900
1200
900
1200
900
1200
1
1
1
6
1
1
5
1
5
5
1
4
2
1
1
1
1
1
1
2
3
1
2
1
3
1
1
1
1
9
1
2
1
2
1
2
1
4
1
4
98 Telluric
164 Telluric
68 K3
53 Target 1
105 Telluric
79 K5
63 Target 1
110 K7
58 Target 1
62 Target 1
48 Target 2
53 Target 1
58 K5
16 M2
19 Target 1
3 Target 1
40 K4
31 Target 2
90 Telluric
33 Target 2
42 Target 2
40 Telluric
28 Target 2
87 Telluric
36 Target 2
49 Telluric
90 K5
82 K7
18 M2
37 Target 1
67 K5
23 Target 2
39 Target 1
30 Target 2
42 Target 1
27 Target 2
44 Target 1
31 Target 2
49 Target 1
34 Target 2
Diffuse cloud
Diffuse cloud
Aborted 148s
Heavy cloud
Cloud
Cloud
Some fog
Seeing > 2"
Seeing > 2"
Seeing > 2"
Seeing > 2"
Seeing > 2"
4 DOPPLER IMAGING
4.1 Doppler Imaging Code
The unnormalised profiles and corresponding errors pro-
duced by SPDECON are used as inputs for the program
DoTS (Collier Cameron 2001). The program calculates the
surface temperature distribution by using χ2 minimization
to find the best fit to the data. A two-temperature model is
used to describe the star (Collier-Cameron & Unruh 1994)
and each element of the star is assigned a spot-filling factor
f . DoTS finds an optimal solution by minimising χ2 while
maximising the entropy of the image. It is not sufficient to
simply minimize χ2 as a large number of images may fit the
data, particularly when the errors are large. The maximum
entropy approach allows a unique image to be returned that
does not contain information that was not required by the
data (Skilling & Bryan 1984). DoTS also requires a look-up
table, containing the shape of the unbroadened spectral line
at a number of limb angles and at the spot and photospheric
temperatures.
4.2 Stellar Parameters
Many of the parameters of TWA 6 were known from previ-
ous work and are listed in table 3. Also listed are our new
or updated parameters as determined using χ2 minimiza-
tion. Testing the program with artificial data has shown
that nearly all of the parameters can be derived indepen-
dently by finding the value which gives the lowest χ2 (even
if the other parameters have not yet been fixed). However,
Doppler Images and Chromospheric Variability of TWA 6
5
Table 2. As table 1 for observations made on the nights of 17/02/06 to 20/02/06.
Object
Date UT Start
Exp.
time[s]
No. of
Frames
S/N Comments Conditions
HR 3037
TWA 6
GJ 529
TWA 6
BD -20 4645
HR 3037
BD -02 801
TWA 6
HR 3037
TWA 6
TWA 17
TWA 6
HR 3037
TWA 6
HR 3037
TWA 6
TWA 17
17
17
18
18
18
19
19
19
19
19
19
19
20
20
20
20
20
17:07:40
17:10:02
16:53:38
17:14:09
18:23:37
09:55:30
10:02:10
10:18:29
15:05:55
15:08:40
17:19:03
17:41:09
09:40:29
09:45:46
14:36:28
14:39:22
17:22:40
20
900
200
900
650
20
800
900
20
900
1200
900
20
900
20
900
1200
1
6
1
4
1
1
1
3
1
8
1
4
1
14
1
10
2
75 Telluric
59 Target
38 K4.5
67 Target
114 K6
121 Telluric
115 K6
42 Target 1
74 Telluric
58 Target 1
46 Target 2
54 Target 1
110 Telluric
38 Target 1
97 Telluric
51 Target 1
31 Target 2
Late start
Storms
Cloud/seeing > 2′′
Cloud/seeing > 2′′
3rd aborted 768s
Storm
Some cloud
Some cloud
Some cloud
Cirrus cloud
Cirrus cloud
Cirrus cloud
Cirrus cloud
Cirrus cloud
The minimum of the χ2 plot is not as clear as with some of
the other parameters and the curve is very shallow around
the minimum χ2, found at an inclination of 47◦. This dif-
ficulty in obtaining the inclination angle has been noted in
other work e.g. Kuerster et al. (1994) where ±10◦ is given
as a reasonable range. The errors given in table 3 were de-
termined by eye from the width of the 'trough' in the χ2 vs
sin i plot.
Using the inclination angle, period and projected equa-
torial velocity the radius was calculated. The resulting value
of 1.05R⊙ is within one sigma of the value calculated using
the distance and V - R colour of (0.8 ± 0.4)R⊙, using the
method described in Gray (1992). It should be noted that
the distance quoted in table 3 is determined using Hippar-
cos distances to other members of the TWA and is therefore
uncertain so the radius determination using the rotational
velocity and period is the more reliable one.
Assuming an age of 8 Myr, an effective temperature of
4000K and a luminosity of 0.25 L⊙ the best-fitting mass
of the star, from both the Alexander et al. (1989) and the
Kurucz (1991) opacities (D'Antona & Mazzitelli 1994), is
0.7M⊙. The luminosity was calculated using the radius and
effective temperature.
4.3 Surface Images
Fig. 3 shows the fits to the profiles for both datasets. For the
first dataset we achieve a χ2 of 1.7, in the second dataset it
is 2.6. A reduced χ2 of 1.0 is not achieved for either dataset.
The errors estimated in LSD using photon noise statistics
are often underestimated in the case of unpolarised spectra
(Donati & Collier Cameron 1997). No scaling of errors was
carried out in this work, so χ2 values greater than 1 are
expected. The normalised residuals of the fits compared to
the data were studied and there are no obvious systematic
effects increasing the χ2 or introducing artefacts into the
maps. The distortions in the profiles are very large, suggest-
Figure 2. Greyscale images of the LSD profiles divided by the
mean profile for 11–13th Feb (left) and 17–20th Feb (right). Spots
appear as emission (white). The greyscale is given in the bar at
the bottom.
for some parameters the minimization must be carried out
for two parameters simultaneously, e.g. period and differen-
tial rotation must be determined together (see section 4.5 for
further discussion of differential rotation). Of the previously
measured values only one changed by a notable amount:
the projected rotational velocity, v sin i was found to have
a value of 72 km s−1, somewhat higher than the previously
measured value of 55 km s−1 (Webb et al. 1999).
Constraining the inclination is slightly more difficult.
6 M. B. Skelly et al
Figure 3. Fits (lines) and deconvolved profiles (symbols) for TWA 6 during Feb 2006. The fits for 11–13th are shown in the two boxes
on the left, and those for the 17–20th are in the two on the right. The numbers on the right of the plots are the rotational phase at
time of observation. The error bars for the data are drawn. The larger error bars in the last 11 profiles in the second box are due to the
removal of a feature due to Moonlight in the spectra.
Table 3. Previously-known and updated parameters for TWA
6. References: [1]: Zuckerman & Song (2004), [2]: Cohen & Kuhi
(1979), [3]: Webb et al. (1999), [4]:Lawson & Crause (2005), [5]:
(D'Antona & Mazzitelli 1994).
Parameter
Spectral type
Temp. [K]
V magnitude
V − R
Distance [pc]
v sin i [kms−1]
Period [day]
Inclination [◦]
Radius [R⊙]
Mass [M⊙]
Lum. [L⊙]
Previous
Value
K7
–
12
1.19
51 ± 5
55 ± 10
0.54 ± 0.01
–
–
–
0.16+0.13
−0.08
New Value
Reference
K7
4000 ± 200
–
–
–
72 ± 1
0.5409 ± 0.00005
47+10
−8
1.05+0.16
−0.15
0.7 ± 0.2
0.25 ± 0.04
[1]
[2]
[3]
[2]
[1]
[3], this work
[4], this work
this work
this work
[5], this work
[3], this work
ing the star has high spot filling factors and sizable tem-
perature differences on the surface. This may explain the
previous identification of the TWA 6 as a binary star in
Jayawardhana et al. (2006), as the presence of a large dis-
tortion at the line–centre may cause it to be mistaken for a
double-lined binary.
The maps given by these fits are shown in Figs. 4 and
5. In all maps white represents a zero spot–filling factor and
a temperature of 4000K, black represents complete spot–
filling and a temperature of 3300K. Both maps show a polar
spot and features extending out of it down to lower latitudes
and some isolated spots near the equator. The features are
consistent between the two maps, and both are consistent
with Fig. 6 which was obtained by fitting both datasets si-
multaneously. The structure of the polar spot can be seen
more clearly in Fig. 7, which shows the star in a flattened
polar projection.
High latitude and polar spots are not observed on the
Sun, but the presence of a polar spot is common to other
young K stars e.g. LO Peg, (Barnes et al. 2005b) and V410
Tau, (Schmidt et al. 2005). In rapidly rotating stars polar
spots have been explained by Coriolis forces causing mag-
netic flux tubes to rise almost parallel to the rotational axis
(Schuessler & Solanki 1992). Magnetohydrodynamical cal-
culations in Caligari et al. (1994) confirmed that for stars
with high rotational velocities and deep convection zones
high spot emergence latitudes are expected.
Doppler Images and Chromospheric Variability of TWA 6
7
Figure 4. Doppler image produced from the data from 11–13th
Feb 2006, shown as a Mercator projection. White represents a zero
spot–filling factor and a temperature of 4000K, black represents
complete spot–filling and a temperature of 3300K. The greyscale
is shown on the right hand side, and latitude on the left. Tick-
marks along the bottom edge of the map show the phase coverage
of the data.The average filling factor f is 9.5%.
Figure 6. As Fig. 4 for the combination of both data sets. The
average filling factor is 9.5%.
Figure 5. As Fig. 4 for the second set of data (from 17–20th Feb
2006). The average filling factor is 10.7%.
Granzer et al. (2000) carried out simulations of flux
tube emergence in stars with a variety of rotation rates and
masses. According to these models, for a TTS with a mass of
0.7 M⊙ and an angular velocity 50 times that of the Sun, the
predominate flux tube emergence should be within a latitude
range 60 − 90◦. The lower latitude spots are not predicted.
In the model for rapidly rotating stars (Ω > 10 Ω⊙) with
very small radiative cores the Coriolis force is expected to
overwhelm the buoyancy forces. The flux tubes detach from
the overshoot layer altogether and the resulting flux rings
rise close to the pole, giving rise to little spot emergence be-
low 60◦. TWA 6 has spot emergence at all latitudes above
the equator, and the spot distribution looks more like that
for a 1.0 M⊙ star. However given the uncertainty in its mass
this is not unreasonable.
Figure 7. A polar projection of the Doppler map for the full
dataset for 11-20th February 2006. Dotted circles are at latitude
intervals of 30◦, the solid circle is at the equator. Longitude inter-
vals of 30◦ are also shown . The numbers around the edge show
phases from 0–1, corresponding to those given in Fig 3. Tickmarks
are placed at phases where observations were made.
4.4 Photometric Lightcurves
The Doppler images produced in section 4.3 were used
to produce monochromatic synthetic lightcurves centred at
B,V and I. The results were double peaked profiles with the
secondary peak being about one-eighth the size of the pri-
mary. Photometry is often used as a constraint on spot sizes,
e.g Vogt et al. (1999), however contemporaneous photome-
try of TWA 6 does not exist. We can qualitatively compare
our lightcurves with the photometry in Lawson & Crause
(2005), taken in March 2000. In doing this we are not in-
tending to directly compare the lightcurves, as several years
elapsed between the two sets of measurements, only to check
that our spot maps are plausible in light of previous results.
As the central wavelength of our LSD profiles is ∼
5500A comparison should be made with the V band pho-
tometry. The amplitude of the variability, 0.1 magnitudes is
around one-fifth of the value measured by Lawson & Crause
8 M. B. Skelly et al
(2005) (∆V = 0.5). At first sight this suggests that the spot
coverage when our measurements were made was rather dif-
ferent to when the photometry in Lawson & Crause (2005)
was carried out. It should however be noted that it is rather
difficult to reliably recover the amplitude of the variation us-
ing spectroscopy alone as spots at low latitudes will be given
less weight in Doppler imaging, and the amplitude will be
underestimated, see e.g. Unruh et al. (1995).
The lightcurve amplitude of ∆V = 0.5 magnitudes is
large compared to other WTTS, where amplitudes of ∼ 0.1
are more typical. A check should therefore be made to make
sure our spot/photospheric flux differences are realistic. A
simple calculation suggests that a spot coverage of approxi-
mately 40% on one hemisphere (i.e. 20% of the total surface)
can recreate the observed brightness change. By using DoTS
to produce synthetic images effects like limb darkening can
be taken into account. Using our spot temperature of 3300K
a lightcurve with an amplitude of 0.5 can be produced with
a gaussian spot at 30◦ latitude and a mean spot-filling fac-
tor of 9%. Such a spot has a full width at half maximum
(FWHM) of 13◦. The mean filling–factor in the images in
Figs. 4, 5 and 6 is 10%. This suggests that there was only
one major spot group in 2000 whereas in 2006 there were rel-
atively large mid-latitude spots apart from the main group
at phase 0.0. Hence it is likely that a combination of un-
derestimation of the variability using Doppler imaging and
a change in the spot coverage over the intervening years
causes the discrepancy. There is certainly a case for further
photometry on this object, as subsequent sections will show.
4.5 Differential Rotation
Figs. 4 and 5 show close agreement between the spots in
the two datasets. Fig. 8 shows the contours of the second
map overlaid on the greyscale image of the first. The two
images in this case were constructed using similar phases in
order to remove differences that may arise due to the differ-
ent phase coverage of the two datasets. The alignment of the
spots suggests that there is no strong differential rotation in
TWA 6. The main differences are in the spots at lower lati-
tudes, where the spots in the greyscale seem slightly larger.
However given the difficulties in determining the exact lati-
tudinal extent and weight of low-latitude features some dif-
ferences are to be expected. There is no strong evidence of
the meridional circulation of the type referred to in section
1.2.
To quantify the differential rotation more rigorously the
two images were cross-correlated at each latitude band with
a range of longitude shifts. If differential rotation of the type
observed in the Sun was occurring this would be apparent
as the band of maximum correlation would curve towards
negative longitude shifts at higher latitudes. A solar-type
differential rotation can be described by
Ω(φ) = Ω0 − ∆Ωsin2φ,
(1)
where Ω0 is the angular velocity of the star at the equator, φ
is the latitude and ∆Ω is the difference in angular velocities
between the equator and the poles. Fig. 9 plots the value of
the cross correlation in a latitude range 0 − 70◦. There is no
apparent curvature in the position of the maximum corre-
lation. An upper limit on ∆Ω can be calculated by fitting
gaussians to each latitude band in the cross correlation. At
Figure 8. The greyscale image from first dataset overlaid with
the contours of the second for the northern hemisphere. The two
images were reconstructed using similar phases.
Figure 9. Cross correlation of the first map with the second. The
solid white line shows the maximum correlation at each latitude.
The black curves are a contour map of the velocity represented
by each part of the cross-correlation image. ∆v is the velocity
resolution of the instrument (6.7 km s−1).
each 10◦ latitude band between 0 − 70◦ the mean position of
the gaussian peak is always within one pixel of 0◦ (one pixel
corresponds to 2◦). At higher latitudes it becomes more diffi-
cult to cross-correlate the images due to the large polar spot
and lower velocity ranges at those latitudes. At lower lati-
tudes the width of the gaussian can provide an upper limit
to the differential rotation. The mean width of the gaussians
between 0 − 50◦ (where most of the isolated spots are) gives
an upper limit on ∆Ω of 0.04 rad day−1.
Carrying out a similar cross-correlation, but this time
comparing longitude bands between the two images, allows
us to investigate whether the star exhibits differential ro-
tation along the meridional direction. As before, a gaussian
was fitted to each band in the correlation image (not shown),
the mean FWHM of these gaussians gives an upper limit on
this form of differential rotation of 0.02 rad day−1.
As a further constraint on the differential rotation DoTS
was used to explore the parameter space to find the values
which give the minimum χ2. The relevant parameters which
should be varied can be found in equation 1. The differential
Doppler Images and Chromospheric Variability of TWA 6
9
Figure 10. A plot of the χ2 value for a variety of angular speeds
and differential rotation values, where Ω0 and ∆Ω are as in equa-
tion 1. The χ2 values are given in the bar at the bottom. The min-
imum χ2, at the '+' in the plot, is at Ω = 11.616 rad day−1 (cor-
responding to a period of 0.54091 d) and ∆Ω = 0.0 rad day−1.
The smallest contour is at a χ2 of 2.376, and approximately cor-
responds to the 3σ confidence level.
rotation of the star can be parameterised using ∆Ω, but
Ω0 must be varied simultaneously, equivalent to varying the
period and differential rotation. The result of this is shown
in Fig. 10, which is a two-dimensional plot of χ2 against
differential rotation and the angular velocity. We find that
the minimum χ2 is at period = 0.54091 day and differential
rotation = 0.00 rad day−1, further evidence that the star
does not have differential rotation.
By drawing contours of ∆χ2 confidence levels can be
placed on the values of Ω0 and ∆Ω obtained. When us-
ing χ2 minimization to find a parameter value there is a
67% probability of the true value lying in the range where
∆χ2 6 1 (Press 2002). The 1σ confidence level therefore lies
at ∆χ2 = 1. As our minimum χ2 is more than 1.0 all the
χ2 values must be divided by the minimum χ2. The surface
is then multiplied by the number of degrees of freedom in
the fit, i.e. the number of spectral data points used to create
the Doppler image. The 1σ confidence level indicates that
differential rotation lies within the range ±0.003 rad day−1.
5 BALMER LINE ANALYSIS
Balmer lines are a useful probe of the circumstellar environ-
ment of TTS. Hα emission is formed in the stellar chromo-
sphere, disc and wind when hydrogen atoms are ionised by
photons from the stellar surface. When a prominence lies in
front of the stellar disc from our point of view most of the
Balmer line photons will scattered out of our line-of-sight,
so prominences in front of the stellar disc appear in absorp-
tion. When the prominence is off the stellar disc the photons
will be scattered into our line-of-sight and the prominence
appears in emission.
Fig. 11 shows the mean Hα and Hβ profile shapes for
the spectra in the first and second datasets. The raw profiles
Figure 11. Mean Hα profiles (top) and Hβ (bottom). In each
plot the upper set of lines shows the normalised mean profile for
the first dataset (solid line) and second dataset (dashed line). The
lower set of lines are the mean profile with a rotationally broad-
ened absorption profile subtracted off. The template spectrum
was rotationally broadened to the v sin i of TWA 6 (72 km s−1),
shifted to an appropriate radial velocity, scaled and subtracted.
The telluric lines have been divided out. For Hα the unsubtracted
and subtracted lines have an equivalent width of 5A and 4.2A re-
spectively, while for Hβ it is 1.2A and 1.0A.
are shown on top and underneath a broadened absorption
spectrum has been subtracted off. The absorption spectrum
was created by broadening the normalised template K7 spec-
trum to a v sin i of 72 km s−1. The telluric lines have been
removed.
Both Hα and Hβ profiles have a narrow component with
a FWHM of 110 km s−1and 113 km s−1
respectively.
This in good agreement with the expected FWHM of a ro-
tationally broadened spectral line for a star with a v sin i of
72 km s−1, which is 112 km s−1 (Gray 1992) if a limb dark-
ening of 0.88 is assumed (Al-Naimiy 1978). The subtracted
Hα profiles have additional broad wings. These are more
prominent on the blue side. The subtracted Hβ profiles by
contrast appear to only have a narrow component.
The presence of a broad component on the Hα emission
is supported by the calculation of the variance profile, shown
in Fig. 12. As described in Johns & Basri (1995) this is is
calculated using
Vλ = hPn
i=1(Iλ,i − Iλ)2
(n − 1)
i1/2
.
(2)
10 M. B. Skelly et al
Figure 12. The normalised variance profiles for Hα. The solid
line is the variance profiles for the first dataset and the dashed line
is for the second dataset. This shows variance out to ±250 km s−1.
The small peak at ∼ −350 km s−1 is due to the presence of an
absorption line.
Fig. 12 shows that the Hα emission varies out to
±250 km s−1, well beyond the v sin i of 72 km s−1. A
similar calculation was carried out for Hβ but this showed
negligible variability off the stellar disc and has not been
shown here.
Figs. 13 and 14 are greyscale images of Hα divided by
a mean profile, at narrow and wide velocity scales, shown
with the LSD profiles for comparison. In the left and centre
images of Fig 13 there is a strong similarity between the LSD
and Hα profiles. The bright regions coincide in these images,
suggesting that the spots on the photosphere correspond
with active Hα regions. The equivalent Hβ image has similar
features, although it is noisier and has not been shown. The
SNR in Hα is lower than for the LSD as the LSD profiles
have been produced using about 2000 lines; we are thus not
able to resolve fine detail in the Hα images that we can see
in the LSD image.
The correspondence is less clear between phases 0 – 0.4
in Fig. 13 than in phases 0.4 – 1.0. There is a dark feature at
around 50 km s−1 in the Hα which is making the contrast
more difficult to see. In Fig. 14 there is also similarity, be-
tween phases 0 – 0.4 some of the bright regions coincide. At
phases 0.4 – 0.6 in the Hα there are dark features masking
any possible correspondence but there is some correspon-
dence between phases 0.6 – 1.0.
Fig. 15 shows the LSD image with the contours of
the Hα on top. Both images have been smoothed. The
tick marks show the direction of decreasing intensity. The
straight horizontal edges in the contour plot are where the
phase coverage was incomplete. The alignment between the
bright regions is apparent in the left hand image for phase
0.4 – 1.0. In the right hand image alignment appears to be
present between phases 0.6 – 1.0 only. Note that we have as-
sumed Sun-like rotation, so a phase of 0.4 corresponds to a
longitude of 216◦. Hence the increased Hα emission appears
to associated with the group of spots between 60 − 220◦.
The behaviour implied by Fig 13 is similar to the
Sun but has not often been observed in other stars.
While low-lying plages have been inferred previously, e.g.
Donati & Collier Cameron (1997), Fern´andez et al. (2004),
we believe that this is the first observation of such a clear
alignment of phases and velocities, indicating that the plages
are overlying the spots. This is in contrast to observations of
earlier type stars such as AB Dor and Speedy Mic where the
dominant signal is from higher-lying prominences (known as
'slingshot prominences').
Also shown on the right-hand sides Fig 13 and 14 is Hα
variability within a velocity range of ±500 km s−1. There
are features well outside the v sin i of the star, suggesting
that slingshot prominences can be seen in emission off the
stellar disc. In particular there is a bright emission feature in
Fig 13 that is can be fitted with a sinusoid, shown in Fig. 16.
In this figure the image has been smoothed and the contrast
has been enhanced, and a sinusoid has been overplotted.
The maximum velocity of this sinusoid is 285 km s−1. If we
assume the prominence is corotating with the star then it is
lying at a height of about 4R∗. The star has a corotation ra-
dius of 2.4R∗. Jardine & van Ballegooijen (2005) show that
a star of this corotation radius can support prominences out
to 4.8R∗. The model assumes that the prominences are em-
bedded in the wind rather than an extended corona. Sling-
shot prominences have not been detected before in T Tauri
stars – their presence suggests that the star can support
stable structures beyond corotation.
Other explanations for this large velocity range should
also be considered. For example, there may be infalling gas,
close to the surface but with a maximum velocity component
along our line of sight of around 285 km s−1. The observed
velocity will vary sinusoidally as the star rotates. Unless the
gas is falling directly onto the equator the magnitude of the
velocity would not be the same at phases separated by 0.5,
as they appear to be here. However if it was on the equator
we would not see it for part of the rotation as the inclination
of the star would cause it to go out of sight.
The spectra plotted in Fig. 13 were taken over 3 nights,
corresponding to about 5.5 rotational periods. All the spec-
tra used in Fig. 13 were taken on either the first or third day
of observations. Most taken on the third day lie in the phase
range 0.5 – 0.7 where the brightening is less obvious, which
implies that the prominence was only present on the first
night. However there are two spectra from the third night in
the phase range 0.7 – 1.0 which do support the presence of
a prominence, in particular the line at phase ∼ 0.75 shows
a brightening at velocities ∼ 300 km s−1. The prominence
may be less apparent at phase 0.5 – 0.7 because it is on the
far side of the star at that time and is obscured by circum-
stellar material.
In Fig. 14 there are features outside v sin i
that may
be evidence of more slingshot prominences, particularly be-
tween phases 0.6 – 1.0. However they do not seem to be
present at other phases and there is not enough information
to justify identifying them as prominences. Also, as they are
not at the same phases as the prominence in Fig. 13 it is
clear that this prominence has not survived the 5 days be-
tween the two sets of observations, suggesting lifetimes of
the order of a few days.
6 CONCLUSIONS
Spectra of TWA 6 were taken at AAT/UCLES during two
epochs separated by approximately nine rotational periods.
Doppler Images and Chromospheric Variability of TWA 6
11
Figure 13. Greyscale images for the first set of data (covering
11 – 13th Feb 2006). From left to right, the normalised decon-
volved profiles from Fig 2; the normalised Hα at velocity range
±90km s−1 (bin width 4.2 km s−1) and the normalised Hα at
velocity range ±500km s−1 (bin widths 17 km s−1). The scale of
the centre and right images is given by the bar at the bottom,
while the scale of the left image is as in Fig. 2. The similarity
between the left and central images should be noted, the bright
regions correspond, particularly from phase 0.4 – 1.0. In the right-
hand image there is a sinusoidal feature with an amplitude of
285 km s−1, see also Fig, 16.
Figure 15. LSD profiles shown in greyscale, overlaid with the
contours of the Hα profiles, for 11–13th Feb (left) and 17–20th
Feb 2006 (right). The LSD and Hα profiles are as in Fig. 13 and
14 but they have been smoothed.
Figure 14. As Fig. 13 for the second set of data, covering 17–
20th Feb 2006. The scale bar refers to the centre and right images,
the left image is as in Fig. 2.
Figure 16. Hα variability overplotted with a sinusoid indicating
the position of a slingshot prominence. The maximum velocity is
285 km s−1. The vertical lines are at ±v sin i, i.e. ± 72 km s−1.
Using these spectra and Doppler imaging two maps of its
spot distribution were produced. The spot distribution of
TWA 6 is qualitatively similar to other T Tauri stars, e.g LO
Peg, (Barnes et al. 2005b) and V410 Tau, (Schmidt et al.
2005), with the predominate spot emergence at high lati-
tudes and spots all the way to the equator.
In the Sun photospheric spots are often coincident with
'plages', active regions in the chromosphere, usually ob-
served using a Hα filter. The Hα emission of TWA 6 in-
dicates that the active chromospheric regions are cospatial
with the photospheric spots - similar to the behaviour of
the Sun. This supports previous findings that Hα emis-
sion in young stars increases when spot coverage is larger
(Donati & Collier Cameron (1997), Fern´andez et al. (2004))
but here we have, for the first time, a strong indication that
the plages are overlying the spots.
TWA 6 appears to support slingshot prominences, the
first such observation in a T Tauri star. At least one promi-
12 M. B. Skelly et al
nence is observed, lying at a velocity of 285 km s−1 which
corresponds to a radial distance of 4R∗ from the axis. There
is evidence that the prominence survives for three days after
the initial detection - suggesting that the structure is stable
over at least 5.5 rotational periods. The presence of a stable
slingshot prominence at a height of 3R∗ from the surface
suggests that the star has a large-scale field. Further obser-
vations of TWA 6 are justified in order to fully characterise
its prominence system, as well as of other stars of similar
age and spectral type to determine whether the behaviour
set out in this paper is typical.
The star has no differential rotation (upper limit
0.003 rad day−1), in line with the previously observed trend
that differential rotation decreases with decreasing effective
temperature and increasing rotational velocity. Earlier dif-
ferential rotation measurements have been for older objects,
here we have extended the observations to younger stars.
The null–measurement is consistent with expectations for
the spectral type and angular speed of TWA 6 and indi-
cates that T Tauri stars exhibit solid–body rotation.
ACKNOWLEDGMENTS
Thanks to D. Mortlock and N. Dunstone for useful discus-
sions; the staff at the AAT for their help during the observ-
ing run and the anonymous referee for helpful comments.
The Vienna Atomic Line Database provided the spectral
line lists used, and the data reduction was carried out us-
ing Starlink software. M. Skelly would like to acknowledge
the financial support of STFC. WAL acknowledges financial
support from UNSW@ADFA Faculty Research Grants.
REFERENCES
Al-Naimiy H. M., 1978, Astrophys. Sp. Sci. , 53, 181
Alexander D. R., Augason G. C., Johnson H. R., 1989,
Collier Cameron A., 2001,
in Boffin H. M. J., Steeghs
D., Cuypers J., eds, Astrotomography, Indirect Imaging
Methods in Observational Astronomy Vol. 573 of Lecture
Notes in Physics, Berlin Springer Verlag, Spot Mapping
in Cool Stars. pp 183–+
Collier Cameron A., Campbell C. G., 1993, Astron. Astro-
phys. , 274, 309
Collier
Cameron A., Robinson R. D.,
1989,
Mon. Not. R. Astron. Soc. , 236, 57
Collier-Cameron A., Unruh Y. C., 1994, Mon. Not. R. As-
tron. Soc. , 269, 814
D'Antona F., Mazzitelli I., 1994, Astrophys. J. Suppl. , 90,
467
Dobler W., Stix M., Brandenburg A., 2006, Astrophys. J. ,
638, 336
Donati J.-F., Semel M., Carter B. D., Rees D. E., Collier
Cameron A., 1997, Mon. Not. R. Astron. Soc. , 291, 658
Donati J.-F., Collier Cameron A., 1997, Mon. Not. R. As-
tron. Soc. , 291, 1
Donati J.-F., Collier Cameron A., Hussain G. A. J., Semel
M., 1999, Mon. Not. R. Astron. Soc. , 302, 437
Dunstone N. J., Barnes J. R., Cameron A. C., Jardine M.,
2006, Mon. Not. R. Astron. Soc. , 365, 530
Fern´andez M., Stelzer B., Henden A., Grankin K., Gameiro
J. F., Costa V. M., Guenther E., Amado P. J., Rodriguez
E., 2004, Astron. Astrophys. , 427, 263
Granzer T., Schussler M., Caligari P., Strassmeier K. G.,
2000, Astron. Astrophys. , 355, 1087
Gray D. F., 1992, The Observation and Analysis of Stellar
Photospheres. Cambridge University Press
Hayashi C., 1966, Ann. Rev. Astron. Astrophys. , 4, 171
Jardine M., van Ballegooijen A. A., 2005, Mon. Not. R. As-
tron. Soc. , 361, 1173
Jayawardhana R., Coffey J., Scholz A., Brandeker A., van
Kerkwijk M. H., 2006, Astrophys. J. , 648, 1206
Johns C. M., Basri G., 1995, Astron. J. , 109, 2800
Kitchatinov L. L., Rudiger G., 1999, Astron. Astrophys. ,
344, 911
Astrophys. J. , 345, 1014
Krist J. E., Stapelfeldt K. R., M´enard F., Padgett D. L.,
Armitage P.
J., Clarke C.
J., Palla F.,
2003,
Burrows C. J., 2000, Astrophys. J. , 538, 793
Mon. Not. R. Astron. Soc. , 342, 1139
Kuerster M., Schmitt J. H. M. M., Cutispoto G., 1994,
Barnes J. R., Cameron A. C., Donati J.-F., James D. J.,
Marsden S. C., Petit P., 2005a, Mon. Not. R. Astron. Soc. ,
357, L1
Barnes J. R., Cameron A. C., Lister T. A., Pointer G. R.,
Still M. D., 2005b, Mon. Not. R. Astron. Soc. , 356, 1501
Barnes J. R., Collier Cameron A., Unruh Y. C., Donati
J. F., Hussain G. A. J., 1998, Mon. Not. R. Astron. Soc. ,
299, 904
Bertout C., 1989, Ann. Rev. Astron. Astrophys. , 27, 351
Bouvier J., Cabrit S., Fernandez M., Martin E. L.,
Astron. Astrophys. , 289, 899
Kupka F. G., Ryabchikova T. A., Piskunov N. E., Stempels
H. C., Weiss W. W., 2000, Baltic Astronomy, 9, 590
Kurucz R. L., 1991, in Crivellari L., Hubeny I., Hummer
D. G., eds, NATO ASIC Proc. 341: Stellar Atmospheres
- Beyond Classical Models New Opacity Calculations. pp
441–+
Lawson W. A., Crause L. A., 2005, Mon. Not. R. As-
tron. Soc. , 357, 1399
Mills D., Webb J., Clayton M., 1997, Starlink User Note
Matthews J. M., 1993, Astron. Astrophys. , 272, 176
152.4
Bouvier J., Forestini M., Allain S., 1997, Astron. Astro-
phys. , 326, 1023
Caligari P., Schussler M., Stix M., Solanki S. K., 1994, in
Caillault J.-P., ed., Cool Stars, Stellar Systems, and the
Sun Vol. 64 of Astronomical Society of the Pacific Confer-
ence Series, Distribution of Magnetic Flux on the Surface
of Rapidly Rotating Stars. pp 387–+
Morin J., Donati J. ., Forveille T., Delfosse X., Dobler W.,
Petit P., Jardine M. M., Cameron A. C., Albert L., Manset
N., Dintrans B., Chabrier G., Valenti J. A., 2007, ArXiv
e-prints, 711
Petit P., Donati J.-F., Collier Cameron A., 2004, As-
tronomische Nachrichten, 325, 221
Press W. H., 2002, Numerical recipes in C++ : the art of
Chabrier G., Kuker M., 2006, Astron. Astrophys. , 446,
scientific computing. ISBN : 0521750334
1027
Rice J. B., Strassmeier K. G., 1996, Astron. Astrophys. ,
Cohen M., Kuhi L. V., 1979, Astrophys. J. Suppl. , 41, 743
316, 164
Doppler Images and Chromospheric Variability of TWA 6
13
Schmidt T., Guenther E., Hatzes A. P., Ries C., Hart-
mann M., Ohlert J. M., Lehmann H., 2005, Astronomische
Nachrichten, 326, 667
Schuessler M., Solanki S. K., 1992, Astron. Astrophys. ,
264, L13
Siess L., Dufour E., Forestini M., 2000, Astron. Astrophys. ,
358, 593
Skilling J., Bryan R. K., 1984, Mon. Not. R. Astron. Soc. ,
211, 111
Strassmeier K. G., 2002, Astronomische Nachrichten, 323,
309
Unruh Y. C., Collier Cameron A., Cutispoto G., 1995,
Mon. Not. R. Astron. Soc. , 277, 1145
Vogt S. S., Hatzes A. P., Misch A. A., Kurster M., 1999,
Astrophys. J. Suppl. , 121, 547
Vogt S. S., Penrod G. D., 1983, Publ. Astron. Soc. Pac. ,
95, 565
Webb R. A., Zuckerman B., Platais I., Patience J.,
White R. J., Schwartz M. J., McCarthy C., 1999, Astro-
phys. J. Lett. , 512, L63
Yang H., Johns-Krull C. M., Valenti J. A., 2007, Astron. J. ,
133, 73
Zuckerman B., Song I., 2004, Ann. Rev. Astron. Astro-
phys. , 42, 685
|
astro-ph/0602136 | 1 | 0602 | 2006-02-06T21:04:48 | A stellar companion in the HD 189733 system with a known transiting extrasolar planet | [
"astro-ph"
] | We show that the very close-by (19 pc) K0 star HD 189733, already found to be orbited by a transiting giant planet, is the primary of a double-star system, with the secondary being a mid-M dwarf with projected separation of about 216 AU from the primary. This conclusion is based on astrometry, proper motion and radial velocity measurements, spectral type determination and photometry. We also detect differential proper motion of the secondary. The data appear consistent with the secondary orbiting the primary in a clockwise orbit, lying nearly in the plane of the sky (that is, nearly perpendicular to the orbital plane of the transiting planet), and with period about 3200 years. | astro-ph | astro-ph | Draft version April 10, 2018
Preprint typeset using LATEX style emulateapj v. 12/14/05
6
0
0
2
b
e
F
6
1
v
6
3
1
2
0
6
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
A STELLAR COMPANION IN THE HD 189733 SYSTEM WITH A KNOWN TRANSITING EXTRASOLAR
PLANET
G´asp´ar ´A. Bakos1,2, Andr´as P´al3,1, David W. Latham1, Robert W. Noyes1, Robert P. Stefanik1
Draft version April 10, 2018
ABSTRACT
We show that the very close-by (19 pc) K0 star HD 189733, already found to be orbited by a
transiting giant planet, is the primary of a double-star system, with the secondary being a mid-
M dwarf with projected separation of about 216 AU from the primary. This conclusion is based
on astrometry, proper motion and radial velocity measurements, spectral type determination and
photometry. We also detect differential proper motion of the secondary. The data appear consistent
with the secondary orbiting the primary in a clockwise orbit, lying nearly in the plane of the sky (that
is, nearly perpendicular to the orbital plane of the transiting planet), and with period about 3200
years.
Subject headings: stars: low-mass, brown dwarfs -- stars: individual: HD 189733 -- stars: individual:
HD 189733B -- planetary systems -- binaries (including multiple)
1. INTRODUCTION
2. EVIDENCE THAT HD 189733 HAS A PHYSICAL
Of the ∼ 170 exoplanets in 146 planetary systems
known at the present time4, the majority (∼ 160)
are revealed only by the reflex velocity of their par-
ent stars, (e.g. Mayor & Queloz 1995; Marcy & Butler
1996), yielding their periods, eccentricities, semi-major
axes, and m sin i values. However, a small number (9)
of the known exoplanets transit their host stars, so that
the inclination ambiguity is removed and -- assuming a
mass and radius for the primary -- their actual mass and
radius, hence their mean density may be determined.
A number of the known exoplanets have been found
to occur in multiple stellar systems. The small sam-
ple of 22 such systems form an important class of ob-
jects (for references, see: Eggenberger et al. 2004b, 2005;
Mugrauer et al. 2004a,b, 2005a). Until now, however, no
planet found in a multiple stellar system also transits its
parent star. In this Letter, we note that the parent star
of the recently-discovered transiting extra-solar planet
HD 189733b (Bouchy et al. 2005, hereafter B05) is itself
a member of a double-star system.
In §2 we present the evidence that the star HD 189733
has a physical companion star. Specifically, §2.1 shows
that it has a common proper-motion companion; §2.2
shows that the companion's radial velocity is the same
as HD 189733 within uncertainties; and §2.3 shows that
the companion is a red dwarf star which must lie at ap-
proximately the same distance as HD 189733. For these
combined reasons the co-moving companion, now labeled
HD 189733B, almost certainly must be a true physical
companion.
In §3 we detect a differential proper mo-
tion, and obtain an initial estimate of the orbital motion
of HD 189733B about HD 189733. Finally, §4 discusses
some implications of this finding, and avenues for future
work.
Electronic address: [email protected]
1 Harvard-Smithsonian Center for Astrophysics (CfA), 60 Gar-
den Street, Cambridge, MA 02138, USA
2 Hubble Fellow
3 Eotvos Lor´and University, Department of Astronomy, H-1518
Budapest, Pf. 32., Hungary
4 http://vo.obspm.fr/exoplanetes/encyclo/index.php
COMPANION STAR
HD 189733 is a close-by (D = 19.3 ± 0.3pc) K0
dwarf, with mass of 0.82 ± 0.03M⊙ and other properties
as described in B05. The 2MASS survey (Cutri et al.
2003) lists a nearby red star, 2MASS 20004297+2242342
(J = 10.12 ± 0.04, H = 9.55 ± 0.08, KS = 9.32 ± 0.03,
J − KS = 0.8), some 3.7 magnitudes fainter in KS and 5
magnitudes fainter in V (estimated from J, H, KS). This
star has an angular separation of 11.2′′ from HD 189733,
lying 10.2′′ W and 4.6′′ to the S.
HD 189733 is not listed in the Washington Double Star
Catalogue (WDS; Mason et al. 2001), but here we will
show that 2MASS 20004297+2242342 is in fact its phys-
ical companion; in anticipation of that result we hence-
forth denote the companion as HD 189733B. Note that
the latter name is distinct from HD 189733b, the name
given by B05 to the planetary companion to HD 189733.
Our conclusion that the two stars form a bound binary
system is based on the following evidence.
2.1. Common Proper Motion
HD 189733 has a relatively high proper motion of
µα = −2.49 ± 0.68 mas yr−1 and µδ = −250.81 ± 0.53
mas yr−1(Hipparcos; Perryman et al. 1997). To deter-
mine whether HD 189733 and HD 189733B share com-
mon proper motion, and hence may be members of the
same physical system, we have inspected the follow-
ing archival material: i) the digitized Palomar Observa-
tory Sky Survey (POSS) scans5 (POSS-I, 1951 R-band;
POSS-II, 1990 R-band, 1992 B-band, and 1996 I-band);
ii) the HST QuickV survey (1982, R-band),
iii) the
2MASS6 2000 J-, H- and K-band scans (Skrutskie et al.
2000).
in November 2005, we used the
TopHAT telescope of the HAT Network at the Fred
Lawrence Whipple Observatory (FLWO), to obtain I-
band images. We also acquired 8 short exposure I-band
frames in December 2005 with KeplerCam on the 1.2m
telescope at FLWO.
In addition,
5 http://archive.stsci.edu/cgi-bin/dss form
6 http://irsa.ipac.caltech.edu/
2
Bakos et al.
1951 POSS−I, R
1982 QuickV, R
1992 POSS−II IR, I
1999 2MASS K
N
G
30"
W
X
B
X
D
E
B
G
B
Fig. 1. -- HD 189733 in four different epochs on registered frames. The known proper motion to the South is well visible. The upper
dashed line shows the declination in 1951, and the lower dashed line for 1999. The companion HD 189733B is not visible on the POSS-I
frame (left panel) because HD 189733 is over-saturated and the scan resolution is not adequate. HD 189733B (indicated by B) is visible
on the rest of the panels, displaced to the SW by ∼ 11′′ from HD 189733; it is seen increasingly better from left to right, as the companion
is relatively brighter at increasingly longer wavelengths (R, I and K bands). "X" denotes a faint star visible on POSS-I and the QuickV
scans, before the Southward moving HD 189733B merges with it. "D" and "E" mark two faint stars that are visible on all of the frames,
but one of the artificial filter glints ("G") merges with their position on the 2MASS frame.
We used our home-grown fihat software environment
(P´al et al. 2006, in preparation) to find sources on all
the above images (∼50 isolated, non-saturated stars),
cross-match them by rejecting outliers, determine the as-
trometric mapping between the images, and transform
them to the same reference system. Visual inspection of
the registered frames show i) the prominent Southward
proper motion of HD 189733 (in harmony with the Hip-
parcos values), ii) a much fainter co-moving companion
that we identify as HD 189733B (for details, see Fig. 1).
The companion is clearly separated from HD 189733 on
the 2MASS J, H and K scans (Fig. 1, right panel),
and is listed in the 2MASS point source catalogue as
2MASS 20004297+2242342. Although not illustrated on
Fig. 1, the co-movement is also demonstrated by the
TopHAT and FLWO 1.2m I-band frames. No other co-
moving companion is detected on these frames.
To quantify the proper motion of HD 189733B, we have
carried out astrometry on the QuickV, POSS-II, 2MASS,
TopHAT and FLWO1.2m observations. We used the
2MASS catalogue as astrometric reference, where the
quoted position uncertainty of bright, isolated sources
is 120mas. We used the previously established pixel cen-
ters of ∼ 50 un-saturated stars that we found by fit-
ting Gaussian profiles. By running our fihat/fistar
star-finder algorithm on artificially generated frames, we
found typical errors of the centroid positions to be on
the order of 0.05pix (corresponding to 0.08′′ on QuickV,
0.05′′ on POSS-II and 2MASS, and 0.1′′ on TopHAT and
0.07′′ on FLWO1.2m). We then derived the second or-
der astrometric mappings between the X,Y coordinates
and the 2MASS astrometric reference (HD 189733(B)
were omitted from the fit), and used this to transform
the pixel coordinates of the frames to the ICRS (α, δ;
Seidelmann & Kovalevsky 2002) system used by 2MASS.
The rms around the fit was ∼ 0.2′′ for QuickV, ∼ 0.18′′
for POSS-II, ∼ 0.05′′ for 2MASS, ∼ 0.3′′ for TopHAT,
and 0.1′′ for the FLWO1.2m, respectively.
The derived proper motion of HD 189733B is µα,B =
−4.1±9 mas yr−1, µδ,B = −264±12 mas yr−1(see Fig. 2),
which are within 2σ of the Hipparcos proper motion of
HD 189733 itself. Therefore, it is clear that, within un-
certainties, HD 189733 and HD 189733B share a com-
mon proper motion and so are likely to comprise a bound
system.
]
c
e
s
c
r
a
[
0
A
R
-
A
R
6
5
4
3
2
1
0
-1
-2
-15 -10
-5
0
5
-15 -10
-5
0
5
Epoch - Epoch0 (2MASS H 1999) [yr]
Epoch - Epoch0 (2MASS H 1999) [yr]
6
5
4
3
2
1
0
-1
-2
]
c
e
s
c
r
a
[
0
c
e
D
-
c
e
D
Fig. 2. -- Proper motion of HD 189733B in RA (left) and Dec
(right) relative to the 2MASS 1999 position. The two panels are
on the same scale. The central solid lines show the linear fit to
the data; the two other solid lines show the same fits using slope
and intersection parameters differing by ±1σ. For reference, the
dashed line shows the relative proper motion of HD 189733 from
Hipparcos, offset by 1.0 arcsec for clarity.
2.2. Radial Velocity
A common radial velocity is a further indication that
the two stars are physically associated. In order to test
this possibility, we obtained spectroscopic observations of
both the primary star and the suspected companion, us-
ing the Center for Astrophysics (CfA) Digital Speedome-
ter (DS; Latham 1992) at the 1.5m Tillinghast telescope
of FLWO, Arizona.
Seven DS observations have been made of the star
HD 189733 dating back to 1995, with the two most re-
cent being 2005 Dec 10 and Dec 17. For each of these
a radial velocity was obtained on the CfA Native Sys-
tem velocity reference (Stefanik et al. 1999). The mean
and standard deviation of these measurements is Vrad
= −2.38 ± 0.20 km s−1, with no significant evidence for
a long-term velocity variation over the past 10 years.
For HD 189733B, two DS observations were made, on
2005 Dec 10 and 17, yielding a mean radial velocity
Vrad,B = −3.1 ± 1.0 km s−1. Thus, within observational
uncertainties the difference in measured velocities is con-
sistent with the two stars being physical companions.
A stellar companion in the HD 189733 system
3
2.3. Characteristics of the Star HD 189733B
There is still a remote chance that the apparent com-
panion could be a distant giant star with high tangential
velocity or a very close-by low luminosity and velocity
sub-dwarf, that happens to have the same sky position,
radial velocity, and proper motion.
While the DS has the primary goal of radial-velocity
measurements, correlation of the spectrum with template
spectra based on the Kurucz (1993) models also yields in-
formation on Teff, log g and stellar rotation v sin i. We
obtained log g = 4.5 ± 0.3 -- a value appropriate to a
main sequence M dwarf. Correlation with observed spec-
tra of M dwarfs with spectral type ranging from M0.0 to
M5.5 gives best correlation near M3.5. We also observed
HD 189733B with the FAST spectrograph on the FLWO
1.5m telescope. Visual comparison of the resulting spec-
tra to known M dwarf spectra suggests a spectral type of
M4V. Based on its spectral type, plus apparent magni-
tude, we then infer a distance to HD 189733B consistent
with the 19.3 pc distance to HD 189733 .
6
6.5
7
7.5
H
M
8
HD189733B
8.5
9
9.5
HD196050B
0.4
0.35
0.3
HD114729B
0.25
0.2
HD213240B
HD75289B
0.175
0.15
0.13
0.11
0.1
10
0.7 0.75 0.8 0.85 0.9 0.95 1 1.05 1.1
J-K
Fig. 3. -- Location of HD 189733B on the (Baraffe et al. 1998)
5Gyr isochrone with other M-dwarf binary companions of stars
with known planets, plotted from Eggenberger et al. (2004b). Stel-
lar masses are labeled along the isochrone in units of M⊙.
We independently estimated the spectral type of
HD 189733B from 2MASS photometry.
Because of
the slightly overlapping profile of the close-by, bright
HD 189733 we performed aperture photometry of
HD 189733B on the 1999 2MASS scans after subtract-
ing off the Gaussian profile of HD 189733, and using
∼ 50 isolated stars with original 2MASS photometry
as reference. This analysis yields J = 10.147 ± 0.02,
H = 9.551 ± 0.03, and KS = 9.318 ± 0.02. These val-
ues are within 0.03mag of the 2MASS Point Source Cat-
alogue values, but we find smaller errors, and slightly
redder J − K color index. We transformed our mea-
sured J − KS value (namely 0.829 ± 0.03), to J − K on
the Bessell and Brett (BB) system following Carpenter
(2005), to obtain (J − K)BB = 0.864 ± 0.04. Then we de-
rived the absolute magnitude MH (8.13 ± 0.045) from
our measured H magnitude and an assumed distance
equal to that of HD 189733 and plotted these values on
the color-magnitude diagram of Mugrauer et al. (2005a),
which also plots a 5-Gyr isochrone from Baraffe et al.
(1998). The position on the color magnitude diagram
above
analyses we
From the
(Fig. 3) corresponds to a stellar mass of 0.175 to 0.2M⊙,
which according to Cox (2000) corresponds to an M
dwarf with spectral type of about M5. The good fit to
the isochrone supports our assumption that the distance
to HD 189733B is similar to that to HD 189733.
conclude
that
HD 189733B is an M dwarf with spectral type in
the range M3.5 to M5, with common proper motion,
radial velocity, and distance to HD 189733. The relative
positions (from astrometry, and assuming equidistance),
and relative velocities of the two stars (from proper
motion and radial velocity data), along with the 1M⊙
total system mass, are consistent within 2σ with the
system being gravitationally bound. Although it would
formally be possible for an interloper M dwarf star to be
passing very close to HD 189733 at the current epoch,
with a relative speed so small as to make it almost
gravitationally bound, this possibility is so remote that
we conclude that the two stars do indeed form a bound
system with projected separation about 216 AU.
3. ORBITAL CHARACTERISTICS OF THE SECONDARY
If the true separation of HD 189733 and HD 189733B
is close to the projected separation, and the orbit is cir-
cular, and the total system mass is ∼1M⊙ (0.82M⊙ for
HD 189733 from B05 and 0.2M⊙ for HD 189733B from
§2.3) then the orbital period is ∼3200 years, correspond-
ing to ∼2km s−1 orbital motion. For a face on orbit this
would yield an observable 22 mas yr−1 differential proper
motion in addition to the co-movement.
In an attempt to detect this, we used two pairs of
2MASS images in the H and K bands obtained in 1999
and 2000 (one of these is the right-hand image in Fig. 1),
with mean epoch 2000.073, and compared them with the
8 high resolution I-band frames of the field taken with
the FLWO 1.2m telescope at epoch 2005.970. We note
that although the other data were useful in confirmation
of the common proper motion (as shown in §2.1), because
of the saturated HD 189733 image (POSS, Quick-V) and
low S/N (TopHAT) they were not used in this precise as-
trometry aimed at refining the differential proper motion.
The astrometry of HD 189733 over the 5.897 year base-
line yields a proper motion of µα = −8 ± 5 mas yr−1 and
µδ = −245 ± 8 mas yr−1. Because this agrees with the
more precise Hipparcos value within uncertainties (§2.1),
we adopt that value for the proper motion of HD 189733.
For HD 189733B we find µα,B = −3 ± 5 mas yr−1 and
µδ,B = −272 ± 5 mas yr−1, which imply a differential
proper motion relative to HD 189733 of ∆µα = −1 ± 5
mas yr−1 and ∆µδ = −21.2 ± 5 mas yr−1(Fig. 4).
Formally, the data indicate a detection of relative
proper motion at the 4σ level. The position of the com-
panion, and its direction and magnitude of relative mo-
tion, are consistent with orbital motion in a clockwise
orbit roughly in the plane of the sky.
However, there could be underlying systematic effects
due to the different instruments and bandpasses used
for the earlier-epoch 2MASS data and the later-epoch
FLWO1.2 data. We estimated one of these, namely
the effect of stellar profile merging of HD 189733B
with HD 189733 that is different on the 2MASS and
FLWO1.2m frames. By subtracting off the Gaussian pro-
file of the primary, we found that the derived proper mo-
tion of HD 189733B changed by only 2mas yr−1. Nev-
4
Bakos et al.
ertheless, while the detection of orbital motion roughly
in the plane of the sky seems secure, it is premature to
derive specific orbital parameters for the system.
-4.70
-4.75
-4.80
-4.85
-4.90
)
c
e
s
c
r
a
n
i
,
h
t
r
o
N
,
3
3
7
9
8
1
D
H
m
o
r
f
(
Y
D
-4.95
+10.00
2MASS
FLWO1.2
+10.05
+10.20
DX (from HD189733, West, in arcsec)
+10.15
+10.10
+10.25
Fig. 4. -- Differential proper motion of HD 189733B relative to
HD 189733. Open squares and circles indicate individual 2MASS
(epoch 2000.073) and FLWO1.2 (epoch 2005.970) data points re-
spectively. Filled square and circle represent the mean of those
respective data points, with error bars representing 1 standard
deviation of those means. Dashed line depicts a circular orbital
path about HD 189733 in the plane of the sky, passing through
the 2MASS position. The triangle depicts the position at epoch
2005.973 for such an orbit.
4. DISCUSSION
Several studies have been done on the characteristics
of close-in planets orbiting the primary star of a mul-
tiple star system. Thus Eggenberger et al. (2004b, and
references therein) found a tendency for massive plan-
ets (m sin i greater than 2MJ) to occur preferentially in
multiple-star systems. However, HD 189733b is a low-
mass planet (1.15MJwithout sin i ambiguity), and yet it
is in a multiple stellar system, thus weakening the dis-
tinction between single and multiple stars.
HD 189733b exceeds all other known extrasolar plan-
ets in binary systems in its proximity to its parent
star (apl=0.031 AU). However, a number of planets in
multiple-star systems are nearly as close (e.g. Tau Boo
b, apl=0.05 AU; HD 75289b, apl=0.046 AU).
The data suggest (see §3) that the binary orbit is
likely to be nearly face on, i.e. the orbital plane would
be nearly orthogonal to the orbital plane of the planet
HD 189733b, which by virtue of its transit we know to
have an inclination of nearly 90◦. A detailed calcu-
lation based on the relative velocities and positions of
the two stars shows that the orbit of HD 189733B and
HD 189733b cannot be coplanar at the 4-σ level (P´al
et al, in preparation). When additional transiting plan-
ets are discovered in multiple star systems, it should be
possible to study the relation of the system architecture
(e.g. mass, semi-major axis of the stellar secondary) to
planet properties in better detail.
Because the HD 189733 system is so close-by, it should
be possible to do excellent astrometry over the next few
years with modern high precision astrometric techniques
from ground or space. The changing radial velocity signal
of either or both of HD 189733 and HD 189733B might
also be detectable after a few years of monitoring with
current high-precision radial velocity devices. High pre-
cision proper motion and radial velocity measurements
should lead to a good characterization of the orbital char-
acteristics of the binary system and their relation to the
orbit of the transiting planet.
Finally, we note that because this system is only 19
pc from Earth and hence both stellar components are
unusually bright for their spectral types, many addi-
tional follow-on observations requiring high resolution
spectroscopy or high precision photometry will be fea-
sible. This should permit a full characterization of the
system including the possible detection of additional low
mass components.
This work was funded by NASA grant NNG04GN74G.
Work by G. ´A. B. was supported by NASA through grant
HST-HF-01170.01-A Hubble Fellowship. A. P. wishes to
acknowledge the hospitality of the Harvard-Smithsonian
Center for Astrophysics, where part of this work has been
carried out. Work of A. P. was also supported by Hungar-
ian OTKA grant T-038437. D. W. L. and R. P. S thank
the Kepler mission for support through NASA Coopera-
tive Agreement NCC2-1390. We acknowledge the use of
the 2MASS survey frames, and the Palomar Sky Survey
digital scans. We thank Perry Berlind at FLWO for tak-
ing the FAST spectra of HD 189733B, and Emilio Falco
for the follow-up exposures using the 1.2m telescope.
A. N. Cox. 2000, Allen's astrophysical quantities, 4th ed. Publisher:
New York: AIP Press; Springer, 2000
Baraffe, I., Chabrier, G., Allard, F., & Hauschildt, P. H. 1998,
Latham, D. W. 1992, ASP Conf. Ser. 32: IAU Colloq. 135, 32, 110
Marcy, G. W., & Butler, R. P. 1996, ApJ, 464, L147
Mason, B. D., Wycoff, G. L., Hartkopf, W. I., Douglass, G. G., &
REFERENCES
A&A, 337, 403
Bessell, M.S., Castelli, F., & Plez, B. 1998, A&A, 333, 231
Bouchy, F., et al. 2005, A&A, submitted, astro-ph/0510119
Carpenter, J. 2005, ApJ, 121, 2851
Charbonneau, D. et al. 2005, astro-ph/0508051
Cutri, R. M., et al. 2003, VizieR Online Data Catalog, 2246,
Eggenberger, A., Udry, S., Mayor, M., Beuzit, J.-L., Lagrange,
A. M., & Chauvin, G. 2004, ASP Conf. Ser. 321: Extrasolar
Planets: Today and Tomorrow, 321, 93
Eggenberger, A., Udry, S., & Mayor, M. 2004, A&A, 417, 353
Eggenberger, A., Mayor, M., Naef, D., Pepe, F., Queloz, D., Santos,
N. C., Udry, S., Lovis, C. astro-ph/0510561
Kervella, P., Th´evenin, Di Folco, E., & S´egransan, D. 2004, A&A,
426, 297
Kurucz, R. 1993, ATLAS9 CD-ROM No. 13. Cambridge, MA,
SAO, 1993., 13
Worley, C. E. 2001, AJ, 122, 3466
Mayor, M., & Queloz, D. 1995, Nature, 378, 355
Mugrauer, M., Neuhauser, R., Seifahrt, A., Mazeh, T., & Guenther,
E. 2005, A&A, 440, 1051
Mugrauer, M., & Neuhauser, R. 2005, MNRAS, 361, L15
Mugrauer, M., Neuhauser, R., Mazeh, T., Guenther, E., &
Fern´andez, M. 2004, Astronomische Nachrichten, 325, 718
Mugrauer, M., Neuhauser, R., Mazeh, T., Alves, J., & Guenther,
E. 2004, A&A, 425, 249
Perryman, M. A. C., et al. 1997, A&A, 323, L29
Seidelmann, P. K., & Kovalevsky, J. 2002, A&A, 392, 341
Skrutskie, M. F., et al. 2000, VizieR Online Data Catalog, 1, 2003
Stefanik, R. P., Latham, D. W., & Torres, G. 1999, ASP
Conf. Ser. 185: IAU Colloq. 170, 185, 354
|
astro-ph/0203040 | 2 | 0203 | 2002-06-24T09:27:19 | Radiation from the Relativistic Jet: a Role of the Shear Boundary Layer | [
"astro-ph"
] | Recent radio and optical large scale jets' observations suggest a two-component jet morphology, consisting of a fast central spine surrounded with a boundary layer with a velocity shear. We study radiation of electrons accelerated at such boundary layers as an option for standard approaches involving internal shocks in jets. The acceleration process in the boundary layer yields in a natural way a two component electron distribution: a power-law continuum with a bump at the energy, where energy gains equal radiation losses, followed by a cut-off. For such distributions we derive the observed spectra of synchrotron and inverse-Compton radiation, including comptonization of synchrotron and CMB photons. Under simple assumptions of energy equipartition between the relativistic particles and the magnetic field, the relativistic jet velocity at large scales and a turbulent character of the shear layer, the considered radiation can substantially contribute to the jet radiative output. In the considered conditions the synchrotron emission is characterized by a spectral index of the radio-to-optical continuum being approximately constant along the jet. A characteristic feature of the obtained broad-band synchrotron spectrum is an excess at X-ray frequencies, similar to the one observed in some objects by Chandra. As compared to the uniform jet models, the velocity shear across the radiating boundary region leads to decrease and frequency dependence of the observed jet-counterjet radio brightness asymmetry. We conclude that a careful investigation of the observational data looking for the derived effects can allow to evaluate the role of the boundary layer acceleration processes and/or impose constraints for the physical parameters of such layers in large scale jets. | astro-ph | astro-ph |
Radiation from the Relativistic Jet:
a Role of the Shear Boundary Layer
L. Stawarz and M. Ostrowski
Obserwatorium Astronomiczne, Uniwersytet Jagiello´nski,
ul. Orla 171, 30-244 Krak´ow, Poland
[email protected]
ABSTRACT
Recent radio and optical
large scale jets' observations suggest a two-
component jet morphology, consisting of a fast central spine surrounded with
a boundary layer with a velocity shear. We study radiation of electrons accel-
erated at such boundary layers as an option for standard approaches involving
internal shocks in jets. The acceleration process in the boundary layer yields in a
natural way a two component electron distribution: a power-law continuum with
a bump at the energy, where energy gains equal radiation losses, followed by a
cut-off. For such distributions we derive the observed spectra of synchrotron and
inverse-Compton radiation, including comptonization of synchrotron and CMB
photons. Under simple assumptions of energy equipartition between the relativis-
tic particles and the magnetic field, the relativistic jet velocity at large scales and
a turbulent character of the shear layer, the considered radiation can substan-
tially contribute to the jet radiative output.
In the considered conditions the
synchrotron emission is characterized by a spectral index of the radio-to-optical
continuum being approximately constant along the jet. A characteristic feature
of the obtained broad-band synchrotron spectrum is an excess at X-ray frequen-
cies, similar to the one observed in some objects by Chandra. As compared to
the uniform jet models, the velocity shear across the radiating boundary region
leads to decrease and frequency dependence of the observed jet-counterjet radio
brightness asymmetry. We conclude that a careful investigation of the observa-
tional data looking for the derived effects can allow to evaluate the role of the
boundary layer acceleration processes and/or impose constraints for the physical
parameters of such layers in large scale jets.
Subject headings: acceleration of particles -- galaxies: jets -- radiation mechanisms:
nonthermal
-- 2 --
1.
Introduction
Jet radial stratification was proposed in order to interpret some radio and optical ob-
servations of extragalactic jets. The existence of the jet boundary layer with a velocity shear
was first suggested by the observations of jets in FR I sources. Owen et al. (1989) concluded
that the jet morphology in M87 is dominated by the boundary layer more than by the shock
structure. The constancy of the radio-to-optical spectral index along the jet and the similar-
ity of radio and optical maps with emissivity peaking near the jet's surface strongly suggested
the acceleration of radiating electrons in situ at the boundary region. Later polarimetry of
M87 jet (Perlman et al. 1999) confirmed its complex structure, which contains knots with a
transverse magnetic field dominating optical emission, and a shear layer which is bright in
the radio band with a strong polarization suggesting a highly ordered magnetic field parallel
to the jet's axis. Jet-counterjet surface brightness asymmetries and magnetic field structures
of the other FR I radio galaxies, 3C 31 and 3C 296 (Laing 1996 and Hardcastle et al. 1997,
respectively), were interpreted in terms of a model in which the jets consisted of a relativistic
spine surrounded by a cylindrical layer with a velocity shear. A study of optical counterparts
of the radio jets in BL Lac objects PKS 2201+044, PKS 0521-365, and 3C371 (Scarpa et al.
1999) emphasized the problem of the reacceleration of electrons radiating within the opti-
cal band in the equipartition magnetic field, as the observed constant synchrotron spectral
index along the jet could not be explained by the first-order Fermi shock acceleration. A
map of the radio spectral index of another well known blazar, Mkn 501, revealed a boundary
region of its jet with an unexpectedly flat spectrum (Edwards et al. 2000). Once again, this
suggested action of some reacceleration mechanism at the jet surface.
Some FR II radio sources also reveal a spine - shear boundary layer jet morphology.
For instance, radio observations of 3C353 (Swain et al. 1998) were interpreted in terms
of a Doppler-hidden relativistic spine surrounded by the slower moving boundary layer.
Polarimetry has shown no magnetic field component transverse to the jet axis in the boundary
layer, in agreement with previous observations of radio linear polarization of quasars (e.g.
Cawthorne et al. 1993). The magnetic field being parallel to the flow velocity is expected in
the presence of a strong velocity shear (e.g., Kahn 1983). VLBI observations of another FR
II object 1055+018 (Attridge et al. 1999) revealed a fragmentary but distinct boundary layer
with the mean longitudinal magnetic field present where interaction with the surrounding
matter takes place (i.e., where the jet bends). Optical and radio observations of the quasar
3C 273 were interpreted in terms of a two component jet model, in which a fast-moving
proper jet was surrounded by the slow-moving radio cocoon (Bahcall et al. 1995). Smooth
changes of the radio-to-optical spectral index along the jet and a lack of correlation between
the optical flux and the spectral index (Jester et al. 2001), strongly suggest reacceleration of
the radiating electrons taking place within the whole jet body. Jester et al. also emphasized
-- 3 --
the discrepancy between the light-travel time along the jet in 3C 273 and the lifetime of
electrons responsible for its synchrotron optical emission. This discrepancy could not be
removed by relativistic beaming or the sub-equipartition magnetic field.
A difference between FR I and FR II radio sources is ascribed to respective jet power and
the importance of its interaction with the surrounding medium (Ghisellini et al. 1998). In
general, FR I jets are thought to decelerate and spread out on the tens-of-kpc scale with their
apparent magnetic field dominated by the component made perpendicular due to propagating
shocks (e.g., Laing 1996; Laing et al. 1999). On the contrary, jets in FR II sources seem to
remain relativistic and well collimated even if they are far away from the nucleus. At the
same time, their apparent parallel magnetic field configuration results from the interaction
with the ambient medium, thereby creating the shear layer with an approximately constant
(limited) thickness and dominating the observed polarization (Cawthorne et al. 1993; Swain
et al. 1998). The relativistic velocity on a large scale leading to apparent superluminal
motions is also present in intermediate objects like M87 (Biretta et al. 1999).
The broad band spectral properties of BL Lacs and FR I radio galaxies give us another
argument in support of a jet stratification. By comparing the multiwavelength observations
of these two classes of AGNs in a framework of a unification scheme, Chiaberge et al. (2000)
found evidence for a significant boundary layer emission, dominating the radiative jet output
in FR Is. It is because FR I nuclei are over-luminous as compared to the predictions of the
unification scheme, and because the ratios of radio to optical luminosities in BL Lacs are
inconsistent with the ones observed in FR I's. The inferred velocities of such boundary
regions are significantly slower as compared to the velocities of the central spines, but still
relativistic, in order to explain anizotropic emission observed from the FR Is' cores.
One should note, that our knowledge of physical conditions at the relativistic jet bound-
ary is still insufficient. Several types of viscosity for the relativistic jet with a shear layer
which confines the beam due to viscous interaction with the ambient medium were inves-
tigated by Baan (1980). Cosmic ray viscosity in a boundary layer with a velocity shear
which influences the hydrodynamics of the jet and particle distribution along its surface
was studied by Earl et al. (1988) and Jokipii et al. (1989). Analogous radiation viscosity
was investigated by Arav and Begelman (1992). The issue of jet stability with respect to
magnetohydrodynamic Kelvin-Helmholtz (KH) instabilities of a boundary layer in a vortex
sheet approximation was studied by Ferrari et al. (1978, 1980, 1981); Hardee (1979, 1983)
and Birkinshaw (1984), among others. The influence of the finite thickness shear layer was
considered by Ferrari et al. (1982) and Birkinshaw (1991). More recently, Hanasz and Sol
(1996, 1998) investigated the growth of KH instabilities in two component relativistic and
nonrelativistic jets. Such instabilities at the interface of a fluid beam and an external medium
-- 4 --
can influence the apparent jet morphology (Ferrari et al. 1978; Hardee 1979; Benford et al.
1980) and generate turbulent MHD waves which accelerate particles (Ferrari et al. 1979;
Eilek 1982). Ferrari et al. (1979) demonstrated that a small amplitude Alfv´en turbulence
developed by the KH instabilities in extended radio lobes can accelerate electrons to high
energies via the wave-particle resonant interactions. Eilek (1982) showed that MHD tur-
bulence generated by the surface instabilities and accelerating the electrons is restricted to
a narrow turbulent edge of the jet. During the last few years, the numerical modeling of
relativistic jets provided some additional information on the nature of such transition layers.
3D simulations revealed the presence of a shear layer with a high specific internal energy and
highly turbulent, subsonic and thin cocoon surrounding a relativistic flow (Aloy et al. 1999).
Simulations of radio emission from such jets (Aloy et al. 2000), show a radial structure in
both intensity and polarization. One should note that the reason for the generation of the
turbulent shear layer in those simulations was numerical viscosity, which only qualitatively
mimics the real viscous processes.
Till the present time no one performed a more thorough study of both acceleration
processes acting at jet boundaries and the resulting observational effects. Our present paper
is intended to study the possibility that the turbulent boundary layers of large scale jets can
substantially contribute to the radiative jet output. The modifications of the standard models
of jet radiation arise from particle acceleration acting continuously within the considered
region and kinematic effects due to the velocity shear. In the present study, we neglect a
possible separate and different energetic electron population present within the jet spine.
We concentrate on the boundary layer radiation, in order to evaluate a role of the electron
acceleration processes in this layer and to constrain hardly known physical parameters of this
region. Below, in Section 2 we discuss an acceleration mechanism creating at the jet boundary
a power-law particle distribution with a much harder component at its high energy end. The
radiation of such electrons through the synchrotron, synchrotron self-Compton and external
Compton processes, modified by the kinematic effects connected with the jet velocity shear,
are derived in Section 3. Consequences of the considered acceleration mechanism in the
context of multiwavelength large scale jets observations are discussed in Section 4. Finally,
a few conclusions are presented in the last Section 5.
2. Acceleration of cosmic ray electrons
Particle acceleration in a velocity shear layer was proposed in the early eighties by
Berezhko with collaborators and then gradually developed (c.f. a review by Berezhko 1990).
The action of such a mechanism at relativistic shear layers occurring at side boundaries
-- 5 --
of relativistic jets was discussed by Ostrowski (1990, 1998, 2000). The mechanism was
considered to provide high energy cosmic rays and to be an important factor influencing
the dynamics of relativistic jets in extragalactic radio sources, thereby forming cosmic rays
cocoons. Ostrowski and Sikora (2001) proposed that such an acceleration mechanism could
provide a cosmic ray proton population being a substantial pressure factor in FR II radio
source lobes. Below we discuss the application of the model of Ostrowski (2000) to the
electron acceleration in the large scale jets.
2.1. Maximum electron energy
Let us assume that a thickness of the transition layer between the considered tens-
of-kiloparsecs scale jet and the surrounding medium, D, is comparable to the jet radius
Rj ≈ 1 kpc (c.f. Owen et al. 1989; Swain et al. 1998, see also Aloy et al. 1999), and that
the magnetic field B frozen into tenuous plasma at and outside the boundary, is parallel
on average to the flow velocity, U. The scattering on the magnetic field irregularities in
a turbulent medium with a perpendicular (≡ radial) mean velocity gradient results in the
acceleration of energetic particles injected into the boundary layer.
In our simple approach the efficiency of the acceleration process in the turbulent shear
layer depends on the velocity structure of the layer and the particle mean free path λ. In
the considered case of the mean magnetic field aligned along the jet in a highly turbulent
boundary layer, a particle gyroradius, rg, can be taken for λ, and a relatively inefficient cross-
field radial diffusion provides a particle escape mechanism from the acceleration region. For
extremely high energy particles with rg > D, the jet boundary can be approximated as a
surface of a discontinuous velocity change. Such a particle moving near the jet boundary
interacts with the magnetic field perturbations and ocassionally crosses the discontinuity
thereby increasing its energy on average. Numerical simulations for 'typical' jet parameters
(Ostrowski 1998) reveal that protons can reach energies above 1019 eV in this mechanism.
For particles with lower energies, rg < D, the acceleration within a finite thickness shear layer
has to be considered. In this case, we deal with two different acceleration processes, the well-
known second order Fermi acceleration in the turbulent medium and the acceleration due
to cosmic ray viscosity in a velocity shear at the boundary. The cosmic ray viscosity in the
context of particle acceleration was discussed by Berezhko in the early eighties (see Berezhko
1990 and references therein) and then by Earl et al. (1988) and Jokipii et al. (1989) in the
case of nonrelativistic flows, and by Ostrowski (2000) for relativistic flow velocities. Below,
we consider these mechanisms to provide relativistic electrons radiating through synchrotron
and inverse Compton processes. Because of rapid energy radiation losses at high energies,
-- 6 --
the electrons are always expected to satisfy the required condition rg ≪ D. Then, the
acceleration time scale in a highly turbulent shear layer can be estimated (Ostrowski 2000)
as
Tacc ∼ (cid:16) rg
c (cid:17)
c2
D(cid:1)2
V 2 +(cid:0) rg
,
U 2
(2.1)
where V is a characteristic velocity of the magnetic turbulence, U is the jet velocity and we
put rg for λ. The first term in the denominator represents the ordinary Fermi mechanism
and the second one represents the viscous acceleration. The Fermi process dominates when
the particle energy is small enough to satisfy rg < D(V /U).
We consider the jet velocity to be comparable to that of the light velocity, c, even on large
scales. The velocity V is comparable to the Alfv´en velocity, VA, for a subsonic turbulence and
possibly a few orders of magnitude smaller than U. MHD turbulent modes in a shear layer
can be dominated by Alfv´en waves, because the other types of waves are damped much more
rapidly. In order to estimate the Alfv´en velocity, VA = B / √4πρ, we assume that the mass
density ρ of the jet and its extended lobes is dominated by nonrelativistic protons, ρ = mpn
(Sikora and Madejski 2000). With the anticipated parameters for the boundary layer of the
tens-of-kpc scale jet -- B ≈ 10−5 G and the proton number density n ≈ 10−4 cm−3
(c.f.
Ferrari et al. 1979) -- one gets VA ≈ 2 · 108 cm/s. Then, with the condition rg < D(V /U)
satisfied and V = VA, the characteristic acceleration time scale (2.1) for electrons ('e') reads
as
Tacc,e ∼
rg
c (cid:16) c
V (cid:17)2
= 5 · 103 γ B−1
µG V −2
8
[s],
(2.2)
where γ is the electron Lorentz factor, V8 ≡ VA/108 cm s−1 and BµG is the magnetic field
in micro-Gauss units. Instead of n, we will use VA to parametrize the acceleration process,
because the mass density is not an observed quantity. One should note, that the equation
(2.2) introduces an optimistic scenario for the acceleration process involving the large am-
plitude MHD turbulence at the scale of rg, leading to short scattering free path λ ∼ rg. The
time scale for particle escape from the acceleration region due to cross-field diffusion can be
evaluated as :
Tesc ∼
D2
κ⊥
= 6 · 1023 γ −1 η−1 BµG D2
kpc
[s],
(2.3)
where κ⊥ is the cross-field diffusion coefficient, κ⊥ = η rg c/3, with η ≤ 1 being a numerical
scaling factor, and D is the acceleration region (boundary layer) thickness (Dkpc ≡ D/1 kpc).
Particle escape determines the electron energy spectrum, when the electrons satisfy the
condition Tacc,e ≈ Tesc , i.e., when rg ≈ (3/η)1/2D(V /c). For the considered boundary layer
parameters and a strong turbulence condition η ≈ 1, radiative cooling becomes important
at much lower electron energies. We consider radiative losses due to the synchrotron and
-- 7 --
the inverse-Compton radiation (in the Thomson regime), to yield the loss time scale
Tloss ∼
6πmc
σT γB2(1 + X)
= 8 · 1020 γ −1 B−2
µG (1 + X)−1
[s],
(2.4)
where X is a ratio of the photon field energy density to the magnetic field energy density.
Therefore, the maximum electron energy in the acceleration process can be estimated by
comparing the time scale (2.2) with the scale for radiative losses, (2.4). For a value of
X ≤ 1, possibly valid at the considered large distance from the active galactic nucleus (see
next section), the maximum electron energy given by the condition Tacc,e = Tloss is
γeq ≈ 4 · 108 V8 B−1/2
µG .
(2.5)
In our model, a rapid radiative cooling of electrons with such high maximum energy is
compensated by acceleration taking place along the jet within the whole boundary layer
volume. For lower turbulence amplitude, the acceleration process can proceed slower to yield
kpcB3
a lower value for γeq. One may note, that at γ = γeq the ratio Tesc/Tacc ≈ 7 · 102D2
µG
(≈ 7 · 105 for the considered jet parameters).
2.2. Electron energy distribution
A simple time dependent model of Ostrowski (2000) uses a regular acceleration term
to represent continues second-order Fermi acceleration plus the not-considered here oblique
shocks possibly formed at the jet side boundary. After switching on the injection of seed
electrons with the low initial Lorentz factor γ0, a spectral evolution of the distribution
function ne(γ, t) for particles accelerated in the shear layer begins as a power-law with a
growing energy cut-off. Then, either it forms a stable high energy cut-off when particle escape
becomes substantial at some γ < γeq, or it evolves into a characteristic shape consisting of two
different components: a power-law part at low energies, ne(γ) ∝ γ −σ, finished with a hard
component modeled here as a nearly monoenergetic peak at γ = γeq due to the accelerated
particles' pile-up caused by losses (Fig. 1). Normalization of the electron energy distribution
depends on an unknown electron injection rate and is a free parameter of the model.
In
principle, the particle high energy component can grow to large values, limited only by the
inefficient side way particle diffusion. However, when the energetic particle pressure becomes
comparable to the ambient medium pressure, it can modify the transition layer structure.
Then, some non-diffusive cosmic ray transport (turbulent diffusion, convective motions) or
other non-linearities of the process can stabilize the average spectral distribution at the
boundary on long time scales, or result in its locally intermittent character. Below, because
of the very nature of the acceleration process, we consider the stationary avarage electron
-- 8 --
spectrum formed within the boundary layer consisting of two basic components: a flat power-
law component at lower energies and a high energy bump. Using the Haeviside function, Θ,
and the Dirac delta function, δ, at energies above the injection energy, γ > γ0, the considered
electron distribution (cf. Fig. 1) can be expressed by an analytic formula
ne(γ) = a γ −σ Θ(γeq − γ) + b δ(γ − γeq),
(2.6)
where a and b are normalization parameters, and we put σ = 2. In the following discussion,
we use numerical values for a and b providing equipartition between the magnetic field energy
density uB ≡ B2/8 π and the energy densities of both electron spectral components:
Z ∞
γ0
(γ mc2) a γ −2 Θ(γeq − γ) dγ = Z ∞
γ0
(γ mc2) b δ(γ − γeq) dγ =
1
2
uB,
(2.7)
where γeq > γ0. With assumed B = 10−5 G and maximum electron energy much higher than
the injection energy, γeq ∼ 108 ≫ γ0 ∼ 10 − 100, the equipartitional normalization of the
power-law electron distribution and the number density of the monoenergetic peak electrons
are a ≈ 10−7 cm−3 and b ≈ 10−14 cm−3, respectively.
Let us mention, that an analogous model was discussed by Schlickeiser (1984), who con-
sidered similar pile-up mechanism for the relativistic electron distribution and studied evo-
lution of the formed monoenergetic electron peak by solving stationary and time-dependent
diffusion equation in the momentum space. In his model, the first-order shock acceleration
accompanied by the second-order turbulent Fermi acceleration creates the pile-up growing
bump at the energy, where the radiative losses balance the energy gains. Schlickeiser assumed
the mean free path of the electron being independent on its energy, contrary to the consid-
ered by us λ ∝ γ. It was shown, that the stationary two-component electron spectrum (the
power-law plus the monoenergetic peak) forms in a presence of efficient shock acceleration,
when the escape time scale is much longer than the Fermi acceleration scale. After switching
off the shock acceleration, at the late phase of the evolution the high energy electron peak
evolves into the ultrarelativistic Maxwell-Boltzmann distribution with the absolute height
of the bump decreasing as exp(−t/Tesc).
Below, we assume that the (quasi -- ) stationary averaged spectrum of electrons formed at
the jet boundary layer can be represented by the distribution (2.6). As long as the diffusive
particle escape to the sides is inefficient, the discussed acceleration process will create a high
energy particle pile-up in a natural way (cf. Stawarz and Ostrowski 2001). The distributions
produced within the jet boundary layer can vary in time and space reflecting the changes of
the local conditions due to the jet interaction with the surrounding medium and action of
non-linear effects in the acceleration process. However, the form of the averaged observed
electron spectrum must consist of the two aforementioned components: the more or less exact
-- 9 --
power-law and the broadened by the fluctuating background high energy bump preceding
the distribution cut-off. Without a detailed acceleration model available, depending also on
the unknown velocity structure of the boundary layer, we use the simple form (2.6) to discuss
the role of these two spectral components for a particular example of comparable energies
stored in both components.
3. Radiation of shear layer electrons
We consider two radiation processes taking place at the relativistic jet boundary -- the
synchrotron (SYN) emission of ultrarelativistic electrons with the energy spectrum (2.6),
and their inverse Compton (IC) cooling. In the case of magnetic bremsstrahlung, we assume
randomly orientated magnetic fields within the turbulent medium of the transition layer.
A synchrotron self-absorption is neglected (i.e., the assumed optical depth for this process
τsyn ≡ µsynl ≪ 1, where µsyn is a synchrotrotron self-absorption coefficient and l is an
emitting region size). Therefore, in our calculations we limit the considered frequencies
somewhat arbitrarily to the 'safe' range ν ≥ νabs = 1010 Hz, where self-absorption effects
are unlikely to occur in the considered large scale jets. Additionally, the present analytic IC
scattering calculations are limited to the Thomson regime, assuming a step scattering cross
section σ = σT for γ ǫi < 1 and 0 otherwise, where ǫi is the seed photon energy expressed in
units of the electron rest energy, mc2. A high energy spectral cut-off due to absorption of
gamma rays via photon-photon pair creation on Cosmic Infrared Background (CIB) photons
is expected to occur for distant sources (e.g., Renault et al. 2001). As we do not consider
such processes, the obtained spectral distributions above TeV energies are the ones derived
for close vicinity of the radiating jets and cannot be directly compared to observations.
Considering the source of the target photons in a tens-of-kpc (or larger) scale jets, we neglect
a possibility of comptonization of the galactic narrow line radiation and IR dust emission
(Celotti et al. 2001), and we study the synchrotron self-Compton (SSC) radiation and the
Compton scattering of external CMB photons (EC). With the assumed B = 10−5 G, the
magnetic field energy density at the jet boundary is uB = B2/8π ≈ 4 · 10−12 erg/cm3. The
energy density of the CMB radiation in a source frame moving with a bulk Lorentz factor
cmbΓ2 ≈ 4 · 10−13 Γ2 erg/cm3,
Γ (neglecting the cosmological redshift correction) is ucmb = u∗
In the jet rest frame,
where an asterisk denotes quantities in the observer's rest frame.
ucmb can be of the same order of magnitude as uB, and both processes can be significant
even for moderate Γ. One should note that in such a case, the estimate (2.5) for γeq still
holds. Celotti et al. (2001) considered also comptonization of the blazar ('bl') emission by
electrons in the boundary region of the kpc-scale conical jet, with a low average boundary
layer Lorentz factor Γ. The energy density of such radiation as seen in the boundary layer
-- 10 --
bl Γ2
kpc Γ−2 erg/cm3, where L′
pc / 4πc z2 Γ2 ≈ 3 · 10−10 z−2
bl ∼ 1043erg/s
local rest frame is ubl = L′
is the intrinsic synchrotron blazar luminosity, Γpc ∼ 10 is the bulk Lorentz factor of the
parsec-scale jet, and z is a distance from the nucleus (zkpc ≡ z/kpc). The blazar emission
dominates over the magnetic field energy density at distances zkpc < 10/Γ, and over the
CMB field at the scales zkpc < 30/Γ2. Hence, for the studied large scale jets (z ≥ 10 kpc)
and the relativistic flows (Γ ≥ 2) the blazar emission is neglected. In order to derive analytic
formulae for emission we are forced to make a simplifying approximation of the isotropic
distribution of the synchrotron radiation in the source local rest frame.
In reality, the
shear layer synchrotron radiation distribution is anizotropic due to kinematic effects of the
nonuniform relativistic flow. A numerical modeling of such anizotropic radiation field will
be presented in a forthcoming paper.
3.1. Synchrotron radiation
The synchrotron emissivity of an isotropic ultrarelativistic electron distribution ne(γ)
averaged over a randomly orientated magnetic field, B, is
jsyn(ν) =
√3e3B
mc2 Z R(cid:18) ν
c1γ 2(cid:19) ne(γ)
4π
dγ,
(3.1)
where c1 = 3eB / 4πmc and R(x) is given by a combination of the modified second order
Bessel functions (Crusius and Schlickeiser 1986) as
R(x) =
x2
2
K4/3(cid:16)x
2(cid:17) K1/3(cid:16)x
2(cid:17) − 0.3
x3
2 hK 2
4/3(cid:16) x
2(cid:17) − K 2
1/3(cid:16)x
2(cid:17)i .
(3.2)
Upon integrating the equation (3.1) for the electron distribution (2.6) and expanding the
function R(xeq) for xeq ≡ ν/c1 γ 2
eq , one can obtain simple analytic
expression for the synchrotron emissivity. For xeq ≫ 1, the function (3.2) decreases expo-
nentially, R(xeq) ∝ exp(−xeq), and thus we neglect emission at frequencies ν ≫ c1γ 2
eq. For
parameters a−7 ≡ a/(10−7 cm−3) and b−14 ≡ b/(10−14 cm−3), the synchrotron emissivity is
(3.3)
eq ≪ 1, R(xeq) ≈ 1.8 · x1/3
jsyn(ν) = (cid:2)f1 ν −1/2 + f2 ν 1/3(cid:3) Θ(νsyn,eq − ν),
where
f1 = 3 · 10−36 B3/2
µG a−7
eq B2/3
νsyn,eq = 1.3 BµG γ 2
eq
f2 = 2 · 10−43 γ −2/3
µG b−14
[cgs],
[cgs],
[Hz].
-- 11 --
For illustration, let us compare the synchrotron radiation spectra of the particle distributions
in the acceleration model of Ostrowski (2000) in successive times from the switching on of
the acceleration process (Fig. 1). From the beginning of injection, the accelerated electrons
form a power-law spectrum with an energy cut-off which grows with time. Then the resulting
emission has (for σ = 2) a power-law form ∝ ν −1/2 with a growing cut-off frequency. When
the maximum electron energy approaches γeq, a particle density peak starts to form and
grow at γ ≈ γeq. Then, an additional radiation component ∝ ν 1/3 appears and eventually
starts to dominate the spectrum at highest synchrotron frequencies, reaching the maximum
at νsyn,eq (∼ 1017 Hz for γeq ≈ 108 and BµG ≈ 10).
3.2.
Synchrotron Self-Compton radiation
The inverse-Compton photon emissivity in the source frame, nic(ǫ, Ω), is given by
nic(ǫ, Ω) = cZ dǫiI dΩiZ dγI dΩe (1 − β cos ψ) σ ni(ǫi, Ωi) ne(γ, Ωe),
(3.4)
where ǫ is the photon energy in the electron rest mass units, ǫ = hν/mc2, ni(ǫi, Ωi) is the
seed photon number density (we put the indices i = syn, ic = ssc for SSC emission, and
i = cmb, ic = ec for EC emission), ne(ǫe, Ωe) is the electron energy distribution, and ψ is
an angle between the electron and the incident photon directions (e.g., Dermer 1995). All
quantities are given in the source rest frame, which in our case, is a part of the boundary
region - a cylindrical layer moving with a constant Lorentz factor Γ (see section 3.4). The
rest-frame emissivity in cgs units can be found next as jic(ν, Ω) = h ǫ nic(ǫ, Ω).
Computations of the SSC emission are performed in the Thomson regime with a scat-
tering cross section independent of the seed photon energy, and with the scattered pho-
tons beamed along the electron direction. The synchrotron radiation and the electron
distribution are assumed to be isotropic in the source frame, ne(γ, Ωe) = ne(γ)/4π and
nsyn(ǫsyn, Ωsyn) = nsyn(ǫsyn)/4π. Thus, the average energy of the comptonized synchrotron
photon is ǫ = 4
obtain
3 γ 2ǫsyn, and in equation (3.4) one can put σ = σT δ(Ω − Ωe) δ(cid:0)ǫ − 4
3γ 2ǫsyn(cid:1) to
nssc(ǫ, Ω) =
cσT
4π Z dǫsynZ dγ nsyn(ǫsyn) ne(γ) δ(cid:18)ǫ −
4
3
γ 2ǫsyn(cid:19) .
(3.5)
For the electron energy distribution given by the equation (2.6) and the optically thin syn-
chrotron intensity, Isyn(νsyn) ≡ jsyn(νsyn) l = ch
4π ǫsyn nsyn(ǫsyn) (where l is an emitting region
linear size), putting ln (νmax/νabs) ∼ 10, where νmax = min(νsyn,eq , mc2/hγeq) = mc2/hγeq,
and neglecting absorption due to photon-photon pair production, one can obtain the isotropic
-- 12 --
SSC emissivity in the Thomson regime
where
f3 = 10−44 B3/2
jssc(ν) = (cid:2)f3 ν −1/2 + f4 ν 1/3(cid:3) Θ(νssc,eq − ν),
−7 + 10−7 γeq a−7 b−14(cid:1)
µG γ −4/3
µG lkpc(cid:0)3 a2
f4 = 5 · 10−58 B2/3
νmax γ 2
νssc,eq =
[cgs],
[Hz].
4
3
lkpc b2
−14
eq
eq = 1.6 · 1020 γeq
(3.6)
[cgs],
Note, that the high energy electron bump contributes to both components of the SSC emis-
sion (i.e., also to the one ∝ ν −1/2).
3.3. External Compton radiation
Cosmic microwave background (CMB) intensity, Icmb(ν), has a blackbody form with a
temperature T ∗ = 2.73 K. In a source frame moving with the relativistic Lorentz factor Γ,
the CMB radiation is strongly anizotropic, providing an external radiation flux mainly from
the direction opposite to the jet direction. Hence, for ultrarelativistic electrons, one can put
(1−β cos ψ) = (1 + µe), where Ωe = (cos−1 µe, φe). Then, the scattered photon energy is ǫ =
γ 2 ǫcmb (1 + µe). Therefore, in the Thomson regime σ = σT δ(Ω − Ωe) δ [ǫ − γ 2ǫcmb (1 + µe)],
and for the isotropic electron distribution, the equation (3.4) reads as
nec(ǫ, Ω) =
cσT
4π
(1 + µ)Z dǫcmbZ dγ ncmb(ǫcmb) ne(γ) δ(cid:2)ǫ − γ 2ǫcmb(1 + µ)(cid:3) .
(3.7)
Considering the power-law component of the electron energy distribution, it is convenient to
model the CMB field as a monochromatic radiation with photon energies ǫcmb = Γ < ǫ∗
cmb >≡
cmb > mc2). For the monoenergetic
2.7·Γ k T ∗/mc2, and a number density ncmb = Γ u∗
electron peak component, one should use the exact blackbody spectrum transformed to the
source frame. Hence, the emissivity of the comptonized CMB radiation, jec(ν, Ω), is
cmb/(< ǫ∗
jec(ν, Ω) = (cid:20)f5(µ) ν −1/2 + f6(µ)
ν 3
exp[f7(µ) ν] − 1(cid:21) Θ(νec,eq − ν),
(3.8)
where
f5(µ) = 8 · 10−41 a−7 [Γ (1 + µ)]3/2
eq b−14 (1 + µ)−2
f6(µ) = 10−85 γ −6
f7(µ) = 3 · 10−11 Γ (1 + µ)−1 γ −2
eq
[cgs],
[cgs],
[Hz−1],
-- 13 --
νec,eq(µ) = 2 · 1011 Γ (1 + µ) γ 2
eq
[Hz].
The high energy bump radiation with frequencies ν < f −1
7 (µ) (i.e., anizotropic Rayleigh-
Jeans part of the blackbody CMB spectrum comptonized by the monoenergetic electrons)
has a power-law form ∝ ν 2.
3.4. Kinematic effects
Let us briefly discuss the main implications of the kinematic effects on the jet radiative
output. First of all, the shear layer consists of sub-layers moving with different velocities and
thus, generating different beaming patterns. In order to specify the jet velocity structure,
we assume a uniform flow Lorentz factor Γj inside the jet spine, i.e. for 0 < r < Rj where r
is a distance from the jet axis, and a radial profile Γ = Γ(r) within the cylindrical transition
region Rj < r < Rc, where the flow velocity decreases and becomes zero (≡ the ambient
medium velocity) at the radius Rc. Let us consider the simplest case of the linear dependence
Γ(r) = 1 + (Γj − 1)
Rc − r
Rc − Rj
.
(3.9)
for Rj ≤ r ≤ Rc. In each considered sub-layer, in the 'local source frame' the synchrotron,
SSC and EC emissivities are approximately given by formulas (3.3), (3.6), and (3.8), respec-
tively. Assuming that the radiating electrons are distributed uniformly (as seen in the local
rest frames) inside the whole boundary region, one can find the observed emissivity of the
sub-layer at the distance r from the jet axis as
j ∗
i (ν ∗, θ∗, r) = δ2(r, θ∗) ji(cid:18)ν =
ν ∗
δ(r, θ∗)
, µ =
µ∗ − β(r)
1 − β(r) µ∗(cid:19) ,
(3.10)
where the index i = syn, ssc or ec, and δ(r, θ∗) is a Doppler factor for a sub-layer moving
−1/2 at an angle θ∗ ≡ cos−1 µ∗ with respect to
with the Lorentz factor Γ(r) = [1 − β 2(r)]
the line of sight, δ(r, θ∗) = 1 / Γ(r) [1 − β(r) µ∗]. Next, the observed flux density from the
boundary layer volume V ∗ can be found as S ∗(ν ∗, θ∗) = d−2R dV ∗ j ∗(ν ∗, θ∗, r), where d is
the distance to the observer. Noting that Γ(r)[1 + µ] ≡ δ[1 + µ∗]/[1 + β(r)], and neglecting
the slowly varying factor [1 + µ∗]/[1 + β(r)] (c.f. Dermer 1995) from the above equations one
obtains the flow modified beaming patterns for the synchrotron, SSC, and EC radiation, as
S ∗
syn(ν ∗, θ∗) ∝ (ν ∗)−αZ Rc
Rj
dr r δ2+α(r, θ∗),
(3.11a)
S ∗
ssc(ν ∗, θ∗) ∝ (ν ∗)−αZ Rc
Rj
dr r δ2+α(r, θ∗) / sin θ∗,
(3.11b)
-- 14 --
S ∗
ec(ν ∗, θ∗) ∝ (ν ∗)−αZ Rc
Rj
dr r δ3+2α(r, θ∗),
(3.11c)
where α is the spectral index of the respective power-law emission, and α = −2 for the
Rayleigh-Jeans part of the comptonized CMB spectrum. In the case of SSC emissivity, we
assumed isotropic distribution of the seed photons in the source rest frame, and we put the
effective emitting region linear size scaling roughly like l ∝ D / sin θ∗. Integration in (3.11)
can be performed numerically for the given values of Γj, Rj, Rc, θ∗, and for the assumed
radial velocity profile.
3.5. Composite spectra
Our model of high energy radiation generated within the jet boundary region has a few
free parameters. They include normalization of the power-law electron distribution, a, a
height of the monoenergetic electron peak, b, the magnetic field induction, B, the Alfv´en
velocity in the shear layer, VA (or the mass density of the plasma within the boundary region,
ρ), and the injection electron energy, γ0. In general, the electron power-law spectral index,
σ, can also be treated as a free parameter represented in this paper by the model value
σ = 2. Here, VA (or ρ) and B limit a value of the maximum electron energy, γeq. Relative
normalizations of radiation spectral components -- synchrotron, SSC and EC emission --
depend on the above parameters and hence the observed composite boundary layer spectral
energy distribution varies with the selected values of a, b, B, γeq, γ0 together with the
geometrical factors Γj, Rj, Rc and θ∗.
For a 'typical' tens-of-kpc scale jet the magnetic field and the Alfv´en speed are, say,
B ≈ 10−5 G and VA ∼ 108 cm/s, and the maximum electron energy can be as high as
γeq ∼ 108 (Eq. 2.5). We assume, that in the boundary region γ0 ≪ γeq and that the
electrons are uniformly distributed within the whole radiating volume (as seen in the local
layer rest frame). The magnetic field energy density can provide estimates for normalizations
of both electron spectral components, if we require the equipartition: a ≈ 10−7 cm−3 and
b ≈ 10−14 cm−3 (Eq. 2.7). The thickness of the large scale jet boundary layer and the jet
radius are assumed to be of the order of a kiloparsec, D ≈ Rj ≈ 1 kpc, and hence Rc = 2 Rj.
Below, for illustration, we consider a part of the jet with the observed length ∆z ≈ 1 kpc.
We also assume, that the jet spine is highly relativistic even at large scales, with that the
bulk Lorentz factor Γj ≈ 10 (c.f. Ghisellini and Celotti 2001). The jet with the described
parameters is used to illustrate possible role of kinematic effects at the observed brightness
distribution along the jet.
The observed spectral energy distribution (SED) of the emission generated at the con-
-- 15 --
sidered part of the boundary region, ν ∗S ∗(ν ∗, θ∗), is shown on figure 2 for four different
viewing angles. We put the source distance from the observer d ≈ 104 kpc.
In order to
estimate the observed luminosity of the boundary layer emission, one has to integrate over
the intrinsic luminosities of the sub-layers with volumes dV ∗ ≡ 2π ∆z r dr and the Doppler
factors δ(r, θ∗). For example, the synchrotron luminosity of the boundary layer electrons is
L∗
syn(θ∗) = Z δ3(r, θ∗) dLsyn ≡ ZZ δ3(r, θ∗) γloss mc2 ne(γ) dγ dV ∗,
(3.12)
where γloss = γ/Tloss denotes the synchrotron energy losses.
In our illustrative example,
synchrotron luminosity of the high energy bump electrons observed at θ∗ = 450 (i.e. the
observed flux multiplied by the 4πd2) can be as high as 1042 erg/s.
4. Multiwavelength large scale jet boundary emission
In this section we compare the model of a radiative boundary layer with the large scale
jet observations. We do not consider the possible contribution from the separate population
of relativistic electrons accelerated within the knots due to the first order shock acceleration.
Our considerations are intended to describe only the main spectral characteristics of the
relativistic electrons accelerated within the turbulent boundary region.
4.1. Radio-to-optical continuum
For the anticipated large scale jet parameters, assuming equipartition between the mag-
netic field and the relativistic electrons accelerated within the shear layer, one will find the
power-law electron component dominating the radio-to-optical jet boundary emission, as il-
lustrated on figure 2. As a result of a continuous acceleration taking place within the whole
boundary layer, the radio-to-optical spectral index is constant along the jet. Of course,
changes of the physical conditions with the distance from the nucleus (e.g., changes of the
magnetic field, turbulent conditions, etc.) can result in smooth changes of the slope of the
synchrotron continuum, like in the case of M 87 or 3C 273 (Meisenheimer et al. 1996; Jester
et al. 2001, respectively).
In our model the acceleration process creates the high energy
electron spectral bump and the optical emission joins smoothly the X-ray bump emission.
Such spectral character is different from the case usually considered in the literature of the
power-law continuum ended by the cut-off at optical or X-ray frequencies, followed by the
inverse-Compton component.
The important manifestation of the boundary layer radio-to-optical emission is a de-
-- 16 --
crease of the jet-counterjet brightness asymmetry as compared to the uniform jet, suggested
previously by Komissarov (1990). This effect arise from the fact, that in a presence of the
velocity shear, for a given viewing angle θ∗ > 1/Γj the observed synchrotron emissivity is
maximized for Γ(r) = 1/ sin θ∗ (i.e., for the maximum value of the Doppler factor δ(r, θ∗)),
while the uniform jet radiation, if present, is Doppler-hidden. Figure 3 illustrates the bound-
ary layer jet-counterjet brightness asymmetry,
S ∗
syn(θ∗)
S ∗
syn(π − θ∗)
Rj
= R Rc
R Rc
Rj
dr r δ2+α(r, θ∗)
dr r δ2+α(r, π − θ∗)
,
(4.1)
for the spectral index α = 0.5 corresponding to the radio emission of our power-law electron
component and the Lorentz factor profile (3.9) with Rc = 2 Rj and Γj = 10 or 3. For
comparison, respective flux asymmetries for the uniform jet models are also plotted. Due to
the considered effect, for Γj = 10 and the moderate inclinations, the observed jet asymmetry
can be reduced by more than one order of magnitude in comparison to the uniform jets.
Hence, the middly-relativistic velocities inferred from the observed jet asymmetries of the
tens-of-kpc scale jets can correspond to the slower boundary layer and not necessarily to the
fast spine, which can still be highly relativistic at the observed distances.
4.2. X-ray bump emission
In the model presented here, the X-ray radiation of large scale jets is dominated by the
high energy electron bump spectral component. As illustrated on figure 2, the observed
synchrotron X-ray flux can be above the extrapolated radio-to-optical continuum. The
spectrum inclination at X-ray frequencies can be different from the power-law observed at
lower frequencies, and the difference depends also on a viewing angle. Figure 4 illustrates this
dependence for the effective X-ray spectral index αx,ef f (θ∗) computed between h ν ∗
1 = 1 keV
and h ν ∗
2 = 5 keV as
log [S ∗
X(ν ∗
X(ν ∗
1 , θ∗) / S ∗
2 / ν ∗
log [ν ∗
1]
2, θ∗)]
.
(4.2)
αX,ef f (θ∗) =
X(ν ∗, θ∗) ∝ R r dr δ2(r, θ∗) R(x∗
eq), a function R(x∗
where S ∗
eq) is given by equation 3.2, and
eq = ν ∗/c1γ 2
eqδ(r, θ∗). The difference between αX,ef f for different jet Lorentz factors and
x∗
different viewing angles results from the spectral character of the high energy electron bump
emission, which is strongly concentrated around the maximum synchrotron frequency νsyc,eq.
Transformation to the observer's rest frame, which depends on the flow Lorentz factor and on
the inclination θ∗, results in the fact, that for the fixed observed frequency ν ∗ in this spectral
range (ν ∗ ∼ νsyn,eq )different parts of the X-ray continuum can contribute to the radiative
-- 17 --
output. For the higher Doppler beaming factors (i.e., higher Γj or smaller θ∗) the effective
X-ray spectral index of the bump electrons with γeq ∼ 108 approach the asymptotic value
−1/3. In the particular case considered by us the counterjet has a steeper X-ray continuum,
as compared to the jet spectrum. Moreover, for large viewing angles, X-ray emission of the
jet softens significantly, opposite to the counterjet emission. Note, that such a behavior is
opposite to what we expect in the case of IC models involving the low-energy tail of the
electron distribution, in which the X-ray spectral index of the jet and the counterjet should
be the same, and approximately constant with the distance from the nucleus and with the
viewing angle. The boundary X-ray emission at the large scales can therefore vary among
different types of the jets, reflecting different physical conditions at their edges and different
kinematic effects involved. Thus, it is interesting to note that the large scale X-ray emission
observed from several radio loud AGNs exhibit the different spectral characteristics.
Chandra X-ray Observatory has detected significant emission from the large scale jets
in quasars (e.g. PKS 0637-752: Chartas et al. 2000), FR II sources (e.g. Pictor A: Wilson
et al. 2001), and FR I radio galaxies (e.g. M 87: Marshall et al. 2001b). In the majority
of cases, thermal bramstrahlung can be excluded, and the non-thermal models seem to
provide the only possible explanation for the X-ray emission (Harris 2001). However, the
spectral index of the X-ray non-thermal continuum is different for different sources. Also
the ratio of the X-ray flux to the optical flux changes significantly in the Chandra jets
sample. This suggest, that different radiative processes can be at work. The models of
the non-thermal large scale jet X-ray emission usually involve the synchrotron emission of
high energy electrons, with Lorentz factor γ ∼ 107, or the inverse Compton scattering of
CMB photons by much less energetic electrons (γ ∼ 102) in a relativistic jet. Both classes
of models are subject to difficulties in explaining the variety of Chandra observations. The
synchrotron emission from one population of the electrons accelerated in a relativistic shock
is excluded if the X-ray flux is above the extrapolated radio-to-optical continuum. Also,
the continuous along the jet character of the large scale X-ray emission is inconsistent with
short lifetimes of the electrons radiating X-rays in the equipartition magnetic fields, unless we
consider continuous acceleration within the whole jet body. On the other hand, the 'beamed'
external Compton (EC(beam)) models require highly relativistic jet flows on the tens-of-kpc
scales, and extrapolation of the single power-law electron energy distribution to low energies,
while, in most of the cases, there is no direct evidence supporting these assumptions. As
discussed in detail by Harris and Krawczynski (2001) such models can be excluded if the
X-ray spectrum is much steeper than the one observed at radio frequencies. One should also
note, that the long life-time of the electrons scattering CMB photons is in contrast with the
knotty morphology of the X-ray jets, as well as with the off-sets between the radio and the
X-ray peaks reported in several brightest knots.
-- 18 --
The EC(beam) model with significant Doppler beaming seems to work well in the case
of the quasar PKS 0637-752 (Tavecchio et al. 2000), where VLBI observations suggest the
parsec-scale flow Lorentz factor Γj > 18 and the inclination angle θ∗ < 6.40. For other
quasars the situation is more complicated. For instance, observations of the jet in 3C 273
(Marshall et al. 2001a; Sambruna et al. 2001) are insufficient to choose between the syn-
chrotron or the EC origin of the X-ray knots' emission, and the existence of an additional
electron population radiating in this range is not excluded. One should also note recently
published HST observations (Jester et al. 2002) showing smooth transition from optical
through UV to X-ray frequencies, as expected in our model. The X-ray luminosities of the
known quasars' jets are in the range 1043 − 1045 erg/s, and their X-ray spectra are very flat,
i.e. αX ∼ 0.23 for 3C 207, (Brunetti et al. 2001), ∼ 0.5 for PKS 1127, (Siemiginowska et al.
2002), or ∼ 0.8 for 3C 273 and PKS 0637. In our model, the observed luminosity of the X-ray
emission generated at the boundary layer can reach the observed values for reasonable jet
parameters (c.f. discussion in section 3.5), and the low effective spectral index is consistent
with small jet inclinations (c.f. Fig. 4). The significant X-ray emission is also observed from
the large scale jets in radio galaxies. The lower X-ray luminosities of such jets, lying in the
range 1039 − 1042 erg/s, and their relatively steep X-ray spectra, αX ∼ 1.0 − 1.5 (Kraft et
al. 2000 for Cen A, Hardcastle et al. 2001 for 3C 66B, Worrall et al. 2001 for B2 0206 and
B2 0755), correspond to the large inclination angles on figure 4 and/or lower jet Lorentz
factors. In most of jets in the radio galaxies, the EC(beam) models are excluded, as they
require too high beaming factors, the magnetic fields much below the equipartition values
and too steep low-energy electron spectra. On the other hand, the synchrotron radiation of
the shock accelerated electrons cannot explain the strong diffusive emission extending few
kpc from the nucleus in, e.g., Cen A and 3C 66B.
We expect, that the detailed investigation of the knots emission can allow verification
of our two-component electron spectrum model, if the knots form by the shock compression
of the energetic electrons previously accelerated at the boundary layer upstream of the shock
(c.f. Begelman and Kirk 1990). More directly, the boundary layer X-ray radiation should
manifests at the inter-knot sections of the jet. Due to the acceleration taking place con-
tinuously within the whole radiating volume, such an X-ray flux should be approximately
constant (or change slowly) along the jet. Unfortunately, there are no observations in large
scale jets of the counterjet X-ray spectra.
In some cases, the lower limits for the X-ray
brightness asymmetry can be estimated, like for the 3C 66B (> 25, Hardcastle et al. 2001),
Pictor A (> 15, Wilson et al. 2001), or PKS 1127 (> 5, Siemiginowska et al. 2002). There
is evidence of a non-zero X-ray flux from the counterjet in the last object. Note, that in a
framework of radiative boundary layer model, the jet-counterjet X-ray brightness asymmetry
is smaller than the one in EC models with strong beaming.
-- 19 --
4.3. Very high energy γ-ray emission
Very high energy (VHE) γ-rays are directly observed only from a few nearby AGNs
during their flaring state (c.f. a review by Kifune 2001). Very short time scales of the
spectral variations at TeV photon energies suggest small sub-parsec sizes of the emitting
regions, where the violent particle shock acceleration take place. In our model, the electrons
accelerated within the boundary layer of the large scale jets are also able to produce such
energetic radiation, mainly due to the IC scattering of the CMB photons, as illustrated
on figure 2. One should note, that figure 2 corresponds to the most optimistic scenario,
with the highly relativistic jet spine and very efficient acceleration creating the electrons
with large Lorentz factors up to γeq ∼ 108. Smaller values of Γj and γeq would result in
decreasing the observed VHE flux and shifting the spectral cut-off to the lower frequencies.
Alternatively, for the lower jet velocities or smaller γeq at the few-of-kpc scales the blazar
emission illuminating the jet from behind can provide seed photons to produce TeV radiation.
Yet another possibility considered in the literature is producing the VHE γ-rays by the
relativistic protons accelerated in the large scale relativistic jets (Aharonian 2002).
VHE γ-rays generated at the distant jets lose their energies due to absorption on the
CIB radiation, CMB photons, and other photon fields surrounding the source. Aharonian
et al. (1994) suggested, that the electromagnetic cascades initiated by such an absorption
create the extended pair halo around the 10-100 TeV extragalactic sources. VHE radiation
generated at the jet boundary layer should naturally contribute to this process.
4.4. Discussion of the free parameters of the model
In order to illustrate the main physical features of the considered model we selected
its parameters in a way to exhibit clearly its main characteristics. In particular we neglect
possible radiation components from the jet spine and the ambient medium next to the
boundary layer, as well as from the shocks possibly formed in the flow.
In real jets the
respective components add to the observed object radiation. For the considered radiating
electrons we assume energy equipartition with the background magnetic field and the equal
division of energy between the power-law and the final bump components. One should note
that application of this simple two-component distribution allows for analytic study of the
formed spectrum, it enables a clear separation of signatures of these both components in
it, but to model the observations a more elaborated distributions can be required. Such
distributions will involve an initial power-law segment in the spectrum, which flattens at
larger energies and forms a final (not sharp) bump before the exponential cut-off (work in
preparation; cf. Petrosian and Donaghy 1999). We also take a particular set of background
-- 20 --
conditions for the jet and its boundary layer (the magnetic field and the matter content, the
turbulence and the large scale kinematic structure) and the ambient radiation fields for this
model. Let us consider how the main predictions of the presented model depend on certain
values of the free parameters assumed in the text, and in particular if the maximum energy
of the boundary layer electrons can substantially differ among different jet models.
For the efficient particle acceleration within the shear boundary layer we assume its tur-
bulent character with high amplitudes MHD or plasma waves scattering energetic electrons.
The assumed configuration of the mean magnetic field within the shear region with dominat-
ing components parallel to the jet axis was already discussed and seems to be confirmed by
the observations. For these conditions the acceleration process acts as described in Section
2, with the time scales for particle escape and radiative losses (Eq-s 2.3 and 2.4) satisfying
the inequality Tesc > Tloss, equivalent to Dkpc > 0.04 B−3/2
µG . One may note that for a range
of magnetic fields in large scale jets, 10 < BµG < 103, the above condition can be easily
satisfied for the transition layer thickness Dkpc ∼ 1 cited in the literature.
−4
−4 and Lj,47 ∼ 1.3 R2
j,kpc Γ2
j c Γ2
The estimate of the maximum electron energy in Eq. 2.5 gives γeq ∼ 8.7 · 107 B1/2
µG n−1/2
.
The parameters in this expression, the magnetic field and the jet particle density, cannot be
varied arbitrary as they also determine the jet energetics. With the anticipated model of
the relativistic jet dynamically dominated by cold protons, the jet (spine) kinetic power is
j mp c2 n and the comoving magnetic field energy density should be
given by Lj ∼ π R2
less than cold protons energy density, uB < mp c2 n. Let us rewrite the above expressions as
BµG < 2 · 103 n1/2
j n−4, where Lj,47 = Lj/1047 erg/s. Estimates often
give for powerful jets Lj,47 ∼ 1 (Sikora 2001). Hence, the number density n−4 ∼ 0.01 − 1 is
estimated for the cases of, respectively, highly relativistic and nonrelativistic large scale jets
with Rj,kpc ∼ 1 and the maximum magnetic field should be in the range 10−4 − 10−3 G. In
Section 2 we adopted the high value of the relativistic jet density, n−4 ∼ 1, as the scale of
1047 erg/s should be regarded as the lower limit for the total kinetic energy of the powerful jets
(cf. discussion in Sikora 2001). Observations of knots in the large scale jets indicate that their
magnetic fields amplified by the shock compression are ∼ 10−5 − 10−4 G. The magnetic field
within the boundary layer, possibly amplified by the velocity shear, is likely to be not greater
than these values. On the other hand, it is not expected to be smaller than the magnetic
field ∼ 10−6− 10−5 G inside radio lobes. Therefore, we conclude that the most likely value of
the magnetic induction within the boundary layer is close to the assumed above BµG ∼ 10.
However, even if one will adopt the less optimistic scenario of a heavy jet (n−4 ∼ 1) with the
weak boundary layer magnetic field (BµG ∼ 1) the maximum electron energy is still large,
γeq ∼ 108, and the maximum synchrotron frequency νsyn,eq ∼ BµG γ 2
eq > 1016 Hz. For more
realistic parameters the synchrotron emission peaks at high X-ray frequencies.
-- 21 --
As a result of weak dependence of γeq on the boundary layer parameters the expected
slowly varying along the jet radio-to-optical spectral index, the discussed decrease of the jet-
counterjet radio brightness asymmetry and distributed along the jet synchrotron emission
peaking at X-ray frequencies, are expected to hold quite generally. Presence of the flat high
energy electron component (pile-up bump) also seems to appear in a natural way, without a
particular tuning of the jet parameters. The main uncertainties of the presented results are
connected with the hardly known velocity profiles of the large scale jets. The bulk Lorentz
factor of the flow, Γ, determines the relative importance of the external photon fields cooling
electrons via inverse-Compton scattering. As discussed in Section 3, the energy density of
the AGN core emission exceeds the magnetic field energy density at distances zkpc < 10/Γ
from the galactic center, while the CMB energy density is higher than uB for BµG < 3 Γ
in our local universe. For the considered values of 1 < Γ < 10 at the tens-of-kpc jet
scales, we expect negligible contribution to the radiative output from comptonization of the
AGN core emission, and a rough equipartition between the magnetic field and the CMB
radiation. It results in approximately equal fluxes of the synchrotron and the IC boundary
layer radiation, if one neglects possible propagation effects to the observer. The form of the
Γ(r) radial profile and the related spatial variation of acceleration efficiency can significantly
influence the beaming pattern and intensity of the boundary layer emission. Therefore the
large scale jet spectra at γ-ray frequencies, in particular relative normalization of the SSC
and EC components presented at figure 2, illustrate their qualitative features only. The
isotropic distribution of the seed photons assumed in deriving the SSC emission may result
in its overestimate. However, even though comptonization of the CMB photons exceeds
the self-Compton emission (for anticipated uB ∼ ucmb ∼ ue), and therefore more detailed
treatment of the SSC radiation, involving anizotropic nature of its seed photons, does not
result in changing of the observed composite high energy spectra.
Finally, let us shortly comment on the jet-counterjet X-ray spectral index asymmetry
and its angular dependence, αX,ef f (θ∗) (Eq. 4.2). Although the hard electron component
is always expected to dominate the boundary layer X-ray emission, the observed spectral
index at critical frequencies can substantially vary for different jet viewing angles, the jet bulk
Lorentz factors and the exact shape of the electron spectrum at highest energies. Therefore,
figure 4 illustrates only a possible behavior of the X-ray jets demonstrating different spectral
indices of the jet and the counterjet.
-- 22 --
5. Conclusions
In the present paper, we study radiation of the ultrarelativistic electrons accelerated
at boundary shear layers of large scale relativistic jets. The acceleration process acting at
such layer is substantially different from the often considered first order shock acceleration.
It can generate, in a natural way, a characteristic two component particle spectrum with a
low energy power-law distribution extending in energy up to the harder pile-up component
forming at higher energies due to radiation losses.
In order to study a possible role of
the mentioned process we constructed a simple, in its several aspects qualitative model of
particle acceleration and radiation from the jet boundary. We propose, that the emission
of such electrons can substantially contribute to the jet radiative output and we investigate
the resulting observational effects. Modifications of the standard jet model arise from the
kinematic effects involved (c.f. Komissarov 1990; Laing et al. 1999; Chiaberge et al. 2000;
Celotti et al. 2001) and the mentioned nature of the acceleration process (Ostrowski 2000):
• With the considered above continuous in situ acceleration model, the radio-to-optical
range of the boundary layer emission is characterized by weak variation of the spectral
index along the jet, assuming the boundary layer physical parameters change smoothly
with the distance from the nucleus. The considered radial velocity profile of the rel-
ativistic jet defines the beaming pattern of the boundary layer emission. In general,
the velocity shear within the boundary region results in decreasing the observed jet-
counterjet brightness asymmetry, affecting evaluations of the jet bulk Lorentz factors.
• If the considered cosmic ray population involves the high energy bump electrons with
the energy density comparable to the magnetic field energy density, one should observe
a high frequency (X-ray) excess over the power-law radiation extrapolated from the
lower frequency measurements. In a framework of our model, the spectral properties
of the X-ray bump emission (the effective spectral index, jet-counterjet brightness
asymmetry, etc) can substantially differ from the one observed at lower frequencies.
We speculate, that such features can be seen at least in some Chandra jets.
• The considered model can explain the observed in some cases edge-brightened jets in
a natural way.
One should note, that our conclusions depend weakly on the unknown detailed physical con-
ditions within the transition layer. The model presented here, far from being complete, is
intended to describe the main radiative properties of the boundary region under simple as-
sumptions about turbulent character of the shear layer, relativistic jet velocities on the large
scales and energy equipartition between the magnetic field and the accelerated electrons. We
-- 23 --
expect, that careful analysis of observational data should reveal the effects discussed above
and thus provide restrictions both on physical parameters within such boundary layers and
the proposed model of electron acceleration.
We are grateful to Marek Sikora for his help and discussions. MO acknowledges a useful
discussions with Vah´e Petrosian. Remarks of the anonymous referee enabled to substantially
improve the final version of the paper. The present work was supported by Komitet Bada´n
Naukowych through the grant BP 258/P03/99/17.
REFERENCES
Aharonian, F.A., Coppi, P.S., Volk, H.J. 1994, ApJ, 423, L5
Aharonian, F.A. 2002, MNRAS, in press
Aloy, M.A., Ibanez, J.M., Marti, J.M., Gomez, and J.L., Muller, E. 1999, ApJ, 523, L125
Aloy, M.A., Gomez, J.L., Ibanez, J.M., Marti, J.M., and Muller, E. 2000, ApJ, 528, L85
Arav, N., and Begelman, M.C. 1992, ApJ, 401, 125
Attridge, J.M., Roberts, D.H., and Wardle, J.F.C. 1999, ApJ, 518, L87
Baan, W.A. 1980, ApJ, 239, 433
Bahcall, J.N., Kirhakos, S., Schneider, D.P., Davis, R.J., Muxlow, T.W.B., Garrington, S.T.,
Conway, R.G., and Unwin, S.C. 1995, ApJ, 452, L91
Begelman, M.C., and Kirk, J.G. 1990, ApJ, 353, 66
Benford, G., Ferrari, A., and Trussoni, E. 1980, ApJ, 241, 98
Berezhko, E.G. 1990, Preprint Frictional Acceleration of Cosmic Rays, The Yakut Scientific
Centre, Yakutsk
Biretta, J.A., Sparks, W.B., and Macchetto, F. 1999, ApJ, 520, 621
Birkinshaw, M. 1984, MNRAS, 208, 887
Birkinshaw, M. 1991, MNRAS, 252, 505
Brunetti, G., Bondi, M., Comastri, A., Setti, G. 2001, A&A, 381, 795
-- 24 --
Cawthorne, T.V., Wardle, J.F.C., Roberts, D.H., and Gabuzda, D.C. 1993, ApJ, 416, 519
Celotti, A., Ghisellini, G., and Chiaberge, M. 2001, MNRAS, 321, L1
Chartas, G., Worrall, D.M., Birkinshaw, M., Creistello-Dittmar, M., Cui, W., Ghosh, K.K.,
Harris, D.E., Hooper, D.J., Jauncey, D.L., Kim, D.W., Lovell, J., Marshall., H.L.,
Mathur, S., Schwartz, D.A., Tingay, S.J., Virani, S.N., and Wilkes, B.J. 2000, ApJ,
542, 655
Chiaberge, M., Celotti, A., Capetti, A., and Ghisellini, G. 2000, A&A, 358, 104
Crusius, A., and Schlickeiser, R. 1986, A&A, 164, L16
Dermer, C.D. 1995, ApJ, 446, L63
Earl, J.A., Jokipii, J.R., and Morfill, G. 1988, ApJ, 331, L91
Edwards, P.G., Giovannini, G., Cotton, W.D., Feretti, L., Fujisawa, K., Hirabayashi, H.,
Lara, L., and Venturi, T. 2000, PASJ, 52, 1015
Eilek, J.A. 1982, ApJ, 254, 472
Ferrari, A., Trussoni, E., and Zaninetti, L. 1978, A&A, 64, 43
Ferrari, A., Trussoni, E., and Zaninetti, L. 1979, A&A, 79, 190
Ferrari, A., Trussoni, E., and Zaninetti, L. 1980, MNRAS, 193, 469
Ferrari, A., Trussoni, E., and Zaninetti, L. 1981, MNRAS, 196, 1051
Ferrari, A., Trussoni, E., and Zaninetti, L. 1982, MNRAS, 198, 1065
Ghisellini, G., Celotti, A., Fossati, G., Maraschi, L., and Comastri, A. 1998, MNRAS, 301,
451
Ghisellini, G., Celotti, A. 2001, MNRAS, 327, 739
Hanasz, M., and Sol, H. 1996, A&A, 315, 355
Hanasz, M., and Sol, H. 1998, A&A, 339, 629
Hardcastle, M.J., Alexander, P., Pooley, G.G., and Riley, J.M. 1997, MNRAS, 288, L1
Hardcastle, M.J., Birkinshaw, M., and Worrall, D.M. 2001, MNRAS, 326, 1499
Hardee, P.E. 1979, ApJ, 234, 47
-- 25 --
Hardee, P.E. 1983, ApJ, 269, 94
Harris, D.E. 2001, in Proc. Oxford Workshop 'Particles and Fields in Radio Galaxies', eds.
R.A. Laing & K.M. Blundell, Oxford.
Harris, D.E., and Krawczynski, H. 2001, ApJ, 565, 244
Jester, S., Roser, H.J., Meisenheimer, K., Perley, R., and Conway, R. 2001, A&A, 373, 447
Jester, S., Roser, H.-J., Meisenheimer, K., Perley, R. A&A, accepted
Jokipii, J.R., Kota, J., and Morfill, G. 1989, ApJ, 345, L67
Kahn, F.D. 1983, MNRAS, 202, 553
Kifune, T. 2001, PASA, 19, in press
Komissarov, S.S. 1990, SvAL, 16, 284
Kraft, R.P., Forman, W., Jones, C., Kenter, A.T., Murray, S.S., Aldcroft, T.L., Elvis, M.S.,
Evans, I.N., Fabbiano, G., Isobe, T., Jerius, D., Karovska, M., Kim, D.-W., Prestwich,
A.H., Primini, F.A., Schwartz, D.A. 2000, ApJ, 531, L9
Laing, R.A. 1996, in 'Energy Transport in Radio Galaxies and Quasars', ASP Conference
Series vol. 100, eds. P.E. Hardee, A.H. Bridle & J.A. Zensus, San Francisco
Laing, R.A., Parma, P., de Ruiter, H.R., and Fanti, R. 1999, MNRAS, 306, 513
Marshall, H.L., Harris, D.E., Grimes, J.P., Drake, J.J., Fruscione, A., Juda, M., Kraft, R.P.,
Mathur, S., Murray, S.S., Ogle, P.M., Pease, D.O., Schwartz, D.A., Siemiginowska,
A.L., Vrtilek, S.D., and Wargelin, B.J. 2001a, ApJ, 549, L167
Marshall, H.L., Miller, B.P., Davis, D.S., Perlman, E.S., Wise, M., Canizares, C.R., and
Harris, D.E. 2001b, ApJ, 564, 683
Meisenheimer, K., Roser, H.J., Schlotelburg, M., 1996, A&A, 307, 61
Ostrowski, M. 1990, A&A, 238, 435
Ostrowski, M. 1998, A&A, 335, 134
Ostrowski, M. 2000, MNRAS, 312, 579
Ostrowski, M., and Sikora, M. 2001, in Proc. 20th Texas Symp. on Relativistic Astrophysics,
eds. J.C. Wheeler & H. Martel (Austin), AIP Conf. Ser. 586, 865
-- 26 --
Owen, F.N., Hardee, P.E., and Cornwell, T.J. 1989, ApJ, 340, 698
Perlman, E.S., Biretta, J.A., Fang, Z., Sparks, W.B., and Macchetto, F.D. 1999, AJ, 117,
2185
Petrosian, V., and Donaghy, T.Q. 1999, ApJ, 527, 945
Renault, C., Barrau, A., Lagache, G., and Puget, J.-L. 2001, A&A, 371, 771
Sambruna, R.M., Urry, C.M., Tavecchio, F., Maraschi, L., Scarpa, R., Chartas, G., and
Muxlow, T. 2001, ApJ, 549, L161
Scarpa, R., Urry, C.M., Falomo, R., and Treves, A. 1999, ApJ, 526, 643
Schlickeiser, R. 1984, A&A, 136, 227
Siemiginowska, A., Bechtold, J., Aldcroft, T.L., Elvis, M., Harris, D.A., Dobrzycki, A. 2002,
ApJ, accepted
Sikora, M., and Madejski, G. 2000, ApJ, 534, 109
Sikora, M. 2001, in 'Blazar Demographics and Physics', eds. P. Padovani & C.M. Urry (San
Francisco), ASP Conf. Ser. 227, 95
Stawarz, L., and Ostrowski, M. 2001, PASA, 19, in press
Swain, M.R., Bridle, A.H., and Baum, S.A. 1998, ApJ, 507, L29
Tavecchio, F., Maraschi, L., Sambruna, R.M., and Urry, C.M. 2000, ApJ, 544, L23
Wilson, A.S., Young, A.J., Shopbell, P.L. 2001, ApJ, 547, 740
Worrall, D.M., Birkinshaw, M., and Hardcastle, M.J. 2001, MNRAS, 326, L7
This preprint was prepared with the AAS LATEX macros v5.0.
-- 27 --
Fig. 1. -- A simple model spectrum N(γ) ≡ d ne(γ)/d log γ of particles accelerated within
the turbulent shear layer. The acceleration process creates power-law energy distribution
ended by the cut-off (the dashed line) or by the high energy pile-up bump at the maximum
electron energy γeq (the soild line) (cf. Ostrowski 2000).
-- 28 --
22
27
12
17
* = 300
ssc1
ec2
ssc2
22
27
12
17
* = 900
12
17
-5
* = 50
-10
syn2
syn1
ec1
12
17
* = 600
-15
-5
-10
-15
]
2
-
m
c
1
-
s
g
r
e
[
)
*
,
*
(
*
S
*
g
o
l
12
17
22
27
12
17
log n
* [Hz]
22
27
22
27
22
27
-5
-10
-15
-5
-10
-15
Fig. 2. -- The observed spectral energy distributions of the radiation generated within the
boundary shear layer for γeq = 108 and θ∗ = 50 , 300 , 600 , 900, at the respective panels. The
presented spectra are integrated over the flow Lorentz factor profile (3.9) and correspond to
the parameters discussed in the text. The index '1' refers to the emission ∝ ν −1/2 of the
power-law part of the electron spectrum. The index '2' refer to the emission of the high
energy electron bump component. In the case of the SSC process, monoenergetic electrons
contribute also to the emission ∝ ν −1/2. Absorption of VHE γ-rays during the propagation
to the observer is neglected.
q
q
q
q
n
n
q
-- 29 --
boundary layer
uniform jet
30
* [deg]
60
90
]
)
*
-
p
(
*
S
/
)
*
(
*
S
[
g
o
l
8
6
4
2
0
0
10
3
10
3
Fig. 3. -- A ratio of the observed radio fluxes from the jet and the counterjet, S ∗(θ∗)/S ∗(π −
θ∗), as a function of the viewing angle θ∗. The dotted lines correspond to the model with the
uniform flow Lorentz factor Γj = 10 or 3 within the whole jet. The solid lines correspond to
the radiation from a boundary shear layer with the radial profile (3.9) and the same Γj = 10
or 3. The Lorentz factors are provided near the respective curves.
q
q
q
-- 30 --
cj
)
*
(
f
f
,
e
X
1,8
1,2
0,6
0,0
j
G
G
j = 10
j = 3
0
30
q * [deg]
60
90
Fig. 4. -- An illustration of the possible effective X-ray spectral index of a jet ('j') and a
counterjet ('cj') boundary layer emission, αX,ef f , as a function of the viewing angle θ∗. The
solid lines correspond to Γj = 10 and the dashed ones to Γj = 3.
a
q
|
astro-ph/0308003 | 2 | 0308 | 2003-11-03T12:09:25 | Pixel correlation searches for OVI in the Lyman alpha forest and the volume filling factor of metals in the Intergalactic Medium at z ~ 2-3.5 | [
"astro-ph"
] | Artificial absorption spectra are used to test a variety of instrumental and physical effects on the pixel correlation technique for the detection of weak OVI absorption. At HI optical depths <= 0.3-1, the apparent OVI detections are spurious coincidences due to HI absorption at other redshifts. In this range, the apparent OVI optical depth is independent of HI optical depth. At larger HI optical depths, apparent OVI optical depth and HI optical depth are correlated. Detailed modelling is required in order to interpret the significance of this relation. High-resolution spectra of four QSOs together with a large suit of synthetic spectra are used to show that the detection of OVI in individual spectra is only statistically significant for overdensities >= 5. These overdensities are larger than would be naively inferred from the onset of the correlation and a tight optical depth-density relation. The lower limit for the volume filling factor of regions which are enriched by OVI is 4% at 95% confidence. This is no larger than the observed volume filling factor of the winds from Lyman break galaxies. Previous claims that the observed OVI absorption extends to underdense regions and requires a universal metal enrichment with large volume filling factor, as may be expected from population III star formation at very high redshift, appear not to be warranted. | astro-ph | astro-ph | Mon. Not. R. Astron. Soc. 000, 1 -- 11 (2003)
Printed 29 October 2018
(MN LATEX style file v2.2)
Pixel correlation searches for O VI in the Lyman α forest
and the volume filling factor of metals in the Intergalactic
Medium at z ∼ 2 − 3.5
Matthew M. Pieri1,3⋆ and Martin G. Haehnelt2
1 Imperial College London, Blackett Laboratory, Prince Consort Road, London SW7 2BW, UK
2 Institute of Astronomy, Madingley Road, Cambridge CB3 0HA, UK
3 Max-Planck-Institut fur Astrophysik, Karl-Schwarzschild-Str. 1, D-85740 Garching, Germany
Accepted 2003 October 9. Received 2003 11 August
ABSTRACT
Artificial absorption spectra are used to test a variety of instrumental and physical
effects on the the pixel correlation technique for the detection of weak O vi absorption.
At H i optical depths <
∼ 0.3 − 1, the apparent O vi detections are spurious coincidences
due to H i absorption at other redshifts. In this range, the apparent O vi optical depth
is independent of H i optical depth. At larger H i optical depths, apparent O vi optical
depth and H i optical depth are correlated. Detailed modelling is required in order
to interpret the significance of this relation. High-resolution spectra of four QSOs
together with a large suit of synthetic spectra are used to show that the detection of
O vi in individual spectra is only statistically significant for overdensities >
∼ 5. These
overdensities are larger than would be naively inferred from the onset of the correlation
and a tight optical depth-density relation. The lower limit for the volume filling factor
of regions which are enriched by O vi is 4% at 95% confidence. This is no larger than
the observed volume filling factor of the winds from Lyman break galaxies. Previous
claims that the observed O vi absorption extends to underdense regions and requires
a universal metal enrichment with large volume filling factor, as may be expected from
population III star formation at very high redshift, appear not to be warranted.
Key words: intergalactic medium - quasars - galaxies:formation - metal enrichment.
3
0
0
2
v
o
N
3
2
v
3
0
0
8
0
3
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
1
INTRODUCTION
The Lyα forest in QSO absorption spectra is now widely
believed to be caused by fluctuating Gunn-Peterson absorp-
tion due to an undulating warm photoionised Intergalactic
Medium (IGM). With this interpretation for the origin of
the Lyα forest, it has become possible to relate the absorp-
tion optical depth to the density of absorption systems in
a meaningful manner. According to this scheme low column
density absorption systems arise from low density regions of
the Universe. A detection of associated metal absorption in
weak Lyα absorption systems would suggest that the metal
enrichment of the IGM is widespread.
Models put
forward to explain this metal
en-
richment fall
into two categories: they either propose
that early and widespread Population III star forma-
tion in pre-galactic structures at z >
∼ 10 is responsible
(e.g. Nath & Trentham 1997; Ferrara, Pettini & Shchekinov
2000; Barkana & Loeb 2001; Madau, Ferrara & Rees 2001)
or suggest a later episode of metal enrichment by winds
from starbursting galaxies at z 6 5 (e.g. Aguirre et al.
2001, Theuns et al. 2001, Adelberger et al. 2003). Trans-
porting metals from the galaxies or protogalaxies where
they are produced into the (low density) IGM is not,
however, a trivial matter. A variety of mechanisms
have been proposed. Some have argued for supernova-
driven (Couchman & Rees 1986, Dekel & Silk 1986) or
'miniquasar'-driven (Haiman, Madau & Loeb 1999) winds.
Others have argued for ejection via galactic mergers
(Gnedin & Ostriker 1997 and Gnedin 1998) or photoevapo-
ration during reionisation (Barkana & Loeb 1999). With an
accurate determination of the fraction of the universe (or
'volume filling factor') enriched by metals it may be pos-
sible to test the veracity of such models. A large volume
filling factor would favour early enrichment by stars form-
⋆ Email: [email protected]
c(cid:13) 2003 RAS
2 Matthew M. Pieri and Martin G. Haehnelt
ing in shallow potential wells, out of which metals are more
easily transported into low density regions.
Observationally, the most important tracers of met-
als in the photoionised IGM are C iv and O vi . Associated
C iv absorption has been detected down to HI column densi-
ties of order 1014 cm−2 (Tytler et al. 1995, Cowie et al. 1995,
Ellison et al. 2000, Pettini et al. 2003) corresponding to an
optical depth of 1 − 10. To investigate even lower density
systems a pixel correlation technique has been employed
where, instead of fitting an associated C iv absorption line,
a pixel-by-pixel search for excess absorption at the corre-
sponding C iv wavelength is performed (Cowie & Songaila
1998). This method was hoped to be more robust than
the use of line 'stacking' (Tytler et al. 1995; Ellison et al.
2000), but there is some ambiguity in the interpretation of
the results. At the relevant redshifts, the O vi species is the
most sensitive tracer of metals in low density regions of the
IGM, provided that the spectrum of the UV background
is sufficiently hard (Rauch, Haehnelt & Steinmetz 1997 and
Hellsten et al. 1998). However, the relevant O vi absorption
lines (λOvia = 1032A, λOvib = 1038A) occur in the same
wavelength range as the Lyman α forest and higher order
lines of the Lyman series, which makes the detection of weak
O vi absorption challenging.
Schaye et al. (2000) claim to have detected O vi at
H i optical depths significantly smaller than unity in some
QSO absorption spectra and used the optical depth-density
relation of numerical hydrodynamic simulations to argue
that this indicates a detection of metals in underdense re-
gions of the IGM. In this work, we apply the pixel-by-
pixel search to synthetic spectra and thereby investigate in-
strumental and physical effects on this technique (see also
Aguirre, Schaye & Theuns 2002). We then assess the sta-
tistical significance of the O vi detection and the implied
density threshold above which the presence of metals is con-
fidently detected.
The spectra (both simulated and observed) are de-
scribed in Section 2. In Section 3 we outline the pixel correla-
tion method used to detect the presence of O vi absorption.
The results of the search for O vi are shown and discussed
in Section 4. In Section 5 we explore the consequences for
the metal enrichment of the low density IGM.
2 SIMULATED AND OBSERVED SPECTRA
2.1 Calculating synthetic spectra
2.1.1 Basic Assumptions
We use the method developed by Bi (Bi, Boerner & Chu
1992; Bi & Davidsen 1997) to analytically calculate syn-
thetic spectra in the fluctuating Gunn-Peterson approxima-
tion for a warm photoionised IGM with a density distribu-
tion that is expected in a ΛCDM structure formation model.
The basic assumptions are:
(i) the baryonic matter traces the dark matter on scales
larger than a filtering scale, which is related to the Jeans
scale;
(ii) the distribution of mildly non-linear densities has a
lognormal probability distribution function (PDF);
(iii) the absorbing gas obeys a simple temperature-
density relation and is not shock-heated.
Table 1. Model parameters used for all simulations
Parameter Chosen Value
Ωm
h
Ωλ
σ8
log[ ¯Teff ]
µ
Γ
0.3
0.65
0.7
0.9
4.6
0.641
1.333
Table 1 shows the model parameters. A brief summary
of how we calculate the synthetic spectra follows.
2.1.2 The Spatial Distribution of Baryons
According to standard structure formation models the mat-
ter distribution was initially a Gaussian random field with
small density fluctuations that evolve under gravity into the
pattern of sheets and filaments characteristic of CDM mod-
els. Many features of such a distribution can be reproduced
by a rank-ordered mapping of a linear Gaussian random field
onto a lognormal PDF of the density. This idea is at the
heart of the method used to calculate synthetic absorption
spectra (Bi et al. 1992). This method is an efficient way of
producing large numbers of realistic synthetic spectra. Note
that we are interested in producing synthetic spectra with
a realistic density PDF and realistic instrumental proper-
ties. Reproducing the detailed clustering properties of the
density field is less important.
We begin by creating a linear Gaussian random
field which is fully determined by its power spectrum,
PDM (k),
for which we have used the form given by
Efstathiou, Bond & White (1992),
PDM(k) ∝
(1 + [ak + (bk)3/2 + (ck)2]ν)2/ν ,
k
(1)
where a = 6.4/Γ, b = 3.0/Γ, c = 1.7/Γ, Γ = Ωmh (Ωm
& h have their normal definitions) and ν = 1.13. The power
spectrum was normalised by the rms fluctuation amplitude
on a 8h−1Mpc scale, σ8, as given in Table 1. In order to take
into account pressure effects that suppress fluctuations on
scales smaller than a certain filtering scale (Gnedin & Hui
1998) the power spectrum of the baryon density in the linear
regime is assumed to have the form,
PB(k) =
PDM(k)
(1 + x2
bk2)2 ,
xb =(cid:20)
2γk ¯Teff
3µmpΩ(1 + z)(cid:21)
1
2
,
(2)
where xb is a comoving scale related to the Jeans length
(Peebles 1980; Bi & Davidsen 1997), γ is the ratio of specific
heats, ¯Teff is the effective mean temperature and µ is the
molecular weight of the IGM. An appropriate value for ¯Teff
has been determined from comparison with hydrodynamic
simulations by Bi & Davidsen (1997). This is the baryonic
matter power spectrum at z = 0 and linear theory is used
to evolve it backward to high redshift.
Equations (1) and (2) provide the 3D power spectrum.
A 1D power spectrum needed for the description of line-of-
sight (LOS) fluctuations is obtained by an integration of the
c(cid:13) 2003 RAS, MNRAS 000, 1 -- 11
Pixel correlation searches for O VI in the Lyman α forest
3
Table 2. Observed Data Sample
QSO
Q1122-165
Q1442+293
Q1107+485
Q1422+231
zem
2.40
2.67
3.00
3.62
zrange
¯Dovi
¯Dhi
2.02-2.34
2.51-2.63
2.71-2.95
3.22-3.53
0.157
0.194
0.254
0.418
0.165
0.193
0.253
0.386
3D power spectrum (Kaiser & Peacock 1991). The matter
density and the peculiar velocity field are related and can
be described as coupled Gaussian random fields (see Bi et al.
1992 for more details). Random realisations of the density
and peculiar velocity in real space (co-moving coordinates)
are obtained using a Fourier transform routine. The non-
linear evolution is modelled by a local rank-ordered mapping
to a lognormal density probability distribution,
n/¯n = exp(cid:20)δ(x) −
hδ2i
2 (cid:21) .
(3)
where δ = (n − ¯n)/¯n and n/¯n is the overdensity fac-
tor. At the relevant redshift the baryonic density field is
still in the mildly non-linear regime and the use of a log-
normal PDF is a reasonable assumption (Bi et al. 1992;
Nusser & Haehnelt 2000).
2.1.3 From the density to the absorption spectrum
The H i optical depth depends not only on the density
but also on the temperature of the gas, due to the tem-
perature dependence of the recombination rate. For den-
sities n/¯n <
∼ 10, numerical simulations show that most of
the gas is not shocked and that a simple power law re-
lation between density and temperature is established by
the balance of photoionisation heating and adiabatic cool-
ing (Hui & Gnedin 1997),
T = T0(n/¯n)α,
(4)
∼ α <
∼ 0.6 and T0 is a constant.
where 0.3 <
For gas at temperatures of a few times 104K in pho-
toionisation equilibrium the neutral hydrogen density is pro-
portional to n2T −0.7/ΓH i, where ΓH i
is the photoionisa-
tion rate. As a result, the optical depth to Lyα absorption
at redshift z can be written as
component of the Hubble parameter and σ0 is the cross-
section for resonant Lyα scattering.
The effect of peculiar velocity and Doppler broadening
are taken into account to obtain the optical depth in redshift
space. We have thereby approximated the Voigt profile by a
Gaussian distribution, which is a reasonable approximation
for absorption lines with equivalent width < 0.7A.
The optical depth for the O vi doublet is calculated as-
suming a fixed ratio of O vi number density to H i number
density (nOVI/nHI). The optical depth of the higher order
Lyman lines are also calculated.
2.1.4 Realistic Synthetic Spectra
To facilitate a more accurate comparison with the observed
spectra we model the optical depth distribution for the same
wavelength range as the observed spectra. There are two re-
gions of the spectra of particular interest; the 'O vi region',
where O vi absorption is searched for, and the 'H i region',
where the corresponding Lyα absorption occurs. The spec-
tra are obtained by co-adding the optical depth of Lyα,
O vi and higher order Lyman series lines. We then use the
estimated statistical error on the flux for the observed spec-
trum to add random noise and perform a Gaussian smooth-
ing to mimic the effect of instrumental broadening.
The optical depth is scaled such that the mean flux
decrement of the observed QSO is reproduced. As we will
discuss later the results of the pixel-by-pixel search depend
sensitively on the mean flux level. We have thus rescaled
the optical depth for the H i and the O vi regions indepen-
dently to reproduce the observed mean flux level in both
regions ( ¯Dhi and ¯Dovi respectively). ¯Dhi is used to set ¯τLyα
in the H i region as well as ¯τOvi (with fixed nOVI/nHI) and
the optical depth of the higher order Lyman lines in the
O vi region. ¯DOvi is then used to set ¯τLyα in the O vi region.
Note that this means that we allow a difference in the value
of Ω2
bar/ΓHi in the two regions of the spectrum. It seems
plausible to assume that the errors we obtain in this way
for a range of Monte Carlo realisations of the LOS density
distribution are similar to those we would have found if we
had only picked random realisations that can reproduce the
mean flux decrement in the H i and the O vi regions simulta-
neously. The latter approach is computationally prohibitive.
In this way we have obtained ensembles of simulated
spectra with the same wavelength range, mean flux decre-
ment and noise properties as the observed spectra, but dif-
fering in their specific random realisation of LOS matter
distribution and noise.
τLyα(z) = σ0 ¯nHII(z),
(5)
2.2 Observations
where
I(z) =Z z2
z1
cdz ′
1 + z
c
H0E(z ′) (cid:18) n(z ′)
¯n (cid:19)1.7
1
b
exp −(cid:18) c(z − z ′)
(1 + z)
+ vpec(z ′)(cid:19)2,b2! ,
(6)
z1 and z2 are the lower and upper limits of the redshift
region of interest, b is the Doppler parameter, nHI is the
neutral hydrogen density, vpec is the peculiar velocity and
a value α = 0.4 is assumed. E(z) is the redshift dependent
c(cid:13) 2003 RAS, MNRAS 000, 1 -- 11
We have used four observed QSOs which are part of the
sample used by Schaye et al. (2000) for a detailed compar-
ison with synthetic QSOs spectra: Q1122-165, Q1442+293,
Q1107+485 and Q1422+231. The data for Q1122-165 was
taken during Commissioning I and Science Verification ob-
servations for the UVES instrument on the VLT (Kueyen)
and released by ESO for public use. The reduction method
can be found in Kim, Cristiani & D'Odorico (2001). The
other spectra were taken with the HIRES instrument
(Vogt S. S. et al. 1994) on Keck I and reduced using pro-
cedures described in Barlow & Sargent (1997). These two
4 Matthew M. Pieri and Martin G. Haehnelt
n OVI
HIn
<D>OVI
(n /<n>)
cut
2. A sketch showing
Figure
τOVI,app − τLyα relation on the principal parameters
nOVI/nHI and (n/¯n)cut.
the dependence
of
the
¯DOvi,
Figure 1. An example of the results from the pixel-by-pixel
search method for the absorption spectrum of Q1107+485.
instruments provide spectra with resolutions Rmax ≈ 80000
and 110000 respectively and have typical S/N of 50. Table
2 gives a list of the redshift range and flux decrement for
the O vi and H i regions. Regions of strong absorption from
other known metal lines were removed.
3 PIXEL-BY-PIXEL SEARCH
We use a technique for the detection of O vi in the Lyman
α forest initially developed by Cowie & Songaila (1998) for
the detection of weak C iv absorption. Schaye et al. (2000)
adapted this method to search for the O vi doublet (see also
Aguirre et al. 2002). We briefly summarise the procedure
here. The redshift range explored in each of the four QSOs
is shown in Table 2. This is the same as used by Schaye et al.
(2000). Due to the increase in noise we did not find it worth-
while to extend the search to lower wavelength.
The continuum-fitted absorption spectrum is re-binned
to pixels of equal size in redshift space for the Lyman se-
ries and both members of the O vi doublet. A width of
(c∆λ/λLyn)kms−1 is taken (where Lyn is the highest or-
der Lyman line used). In order to take into account possible
velocity shifts between O vi and H i absorption, we have also
tried larger pixels widths. This did not improve the sensi-
tivity for O vi detection.
A search for correlated absorption at the wavelengths
of the Lyman series and the O vi doublet is then performed.
Each pixel of the Lyα absorption region is considered in
turn. If the flux level is within σnoi/2 of the continuum level
(where σnoi is the estimated statistical error on the flux for
the observed spectrum) or in a region where the associated
O vi absorption is not covered the pixel is discarded. If the
pixel is within σnoi/2 of saturation we attempt to use the
higher order Lyman lines to estimate τLyα. We thereby use
the higher order line with the lowest equivalent Lyα optical
depth τLyα = min(τLynfLyα/fLynλLyn) that is not within
σnoi/2 of saturation or the continuum (where fLyn is the
oscillator strength of the Lyn line).
Once we have established a value for τLyα, we consider
the pixels corresponding to absorption by the O vi doublet
at the same redshift. The limiting factor in the use of
this method is coincident absorption from H i, either in
the form of higher order Lyman lines at similar redshift or
by Lyα lines at lower redshift. Obviously we want to min-
imise this contamination. The doublet nature of the O vi is
utilised in this regard. The smaller equivalent absorption,
τOvi = min(cid:18)τOvia,
fOviaλOviaτOvib
fOvibλOvib (cid:19) ,
(7)
will be the least contaminated. This gives us an up-
per limit to the O vi optical depth. We will thus call
the O vi optical depth obtained in this way the apparent
O vi optical depth.
As pointed out by Aguirre et al. (2002) the contam-
ination by higher order Lyman lines can be estimated
from the corresponding Lyα absorption and corrected. Such
contamination is subtracted from the O vi optical depth
where possible. If this is not possible because of saturated
Lyα absorption we also discard the pixel.
We then bin the ensemble of optical depth pairs in
Lyα optical depth and plot the relation between the median
optical depth of O vi and Lyα . We use the median instead
of the mean as the distribution of O vi optical depth in each
bin is skewed toward high optical depths due to the contam-
ination by H i absorption.
Fig. 1 shows an example of the relation of the apparent
O vi and Lyα optical depth (τOVI,app − τLyα relation) ob-
tained in this way. At small Lyα optical depth, the median
O vi optical depth is constant. As we will demonstrate in
detail in the next section (see also Aguirre et al. (2002)),
the apparent O vi is here almost exclusively due to contam-
ination of coincident H i absorption and this region of the
diagram cannot be used to infer a detection of O vi . At
larger Lyα optical depth, the apparent O vi optical depth
rises and is mainly due to the presence of O vi . The level
of metal enrichment is, however, not straightforward and
requires detailed modelling.
4 RESULTS FROM SIMULATED SPECTRA
4.1 Principle parameters
To obtain an indication of how the pixel-by-pixel search de-
pends upon instrumental effects and physical parameters
and what errors should be expected due to variation of
the LOS density distributions, we investigated a large suit
c(cid:13) 2003 RAS, MNRAS 000, 1 -- 11
Pixel correlation searches for O VI in the Lyman α forest
5
Figure 3. The τOVI,app − τLyα relation for an ensemble of forty synthetic spectra. Left: ¯DOVI increases from bottom to top. Right: ¯DHI
increases from bottom to top
of synthetic spectra. We have varied the mean flux decre-
ment in the O vi and H i regions ( ¯DOvi and ¯DHi), the ratio
of O vi to H i density (nOvi/nHi) and introduced a varying
cut-off in the density below which we set the O vi density
to zero ((n/¯n)cut ). Fig. 2 sketches how the relation of ap-
parent O vi optical depth and Lyα optical depth depends
on the three key parameters ¯DOvi, (n/¯n)cut and nOvi/nHi.
A detailed discussion of the effect of these and the other pa-
rameters using synthetic spectra of Q1122 will follow below.
The parameters were varied around fiducial values denoted
by an 'F' in figures 3-5. The error bars in these figures are the
1σ spread of the optical depths for ensembles of 40 spectra.
4.2 Varying the mean optical depth
Fig. 3 shows how changing the mean flux decrement af-
fects the τOVI,app − τLyα relation. In the left panels we con-
sider changes in the O vi region, ¯DOvi. With increasing
flux decrement the level of spurious coincidences due to
Lyα absorption from gas at a lower redshift rises and mim-
ics the detection of O vi . Since this Lyα absorption is un-
correlated with absorption in the H i region, this leads to
a floor of constant apparent O vi optical depth which rises
with increasing ¯DOvi. In the right panels of Fig. 3 we vary
the mean flux decrement in the H i region ¯DHi. This time
the shape of the relation between H i optical depth and ap-
c(cid:13) 2003 RAS, MNRAS 000, 1 -- 11
parent O vi optical depth does not change but the error bars
increase substantially with decreasing mean flux decrement.
This is because, for small mean flux decrement, large optical
depth Lyα pixels are poorly sampled.
4.3 Varying the O vi distribution
In Fig. 4 we vary the O vi distribution for our synthetic
spectra. The left panels shows how varying the ratio of
O vi to H i density affects the τOVI,app − τLyα relation. With
increasing ratio a correlation between apparent O vi optical
depth and H i optical depth develops for large H i optical
depth but the errors are large. As discussed in the introduc-
tion the main aim of the search for O vi at small H i optical
depth is to establish the volume filling factor of metals in the
Universe. Because of the possible spurious detections due to
random coincidences this is not straightforward.
There is another important effect which can lead to spu-
rious detection of O vi . Low Lyα optical depth is not neces-
sarily associated with low density gas but can also be due to
the wing of a strong Lyα absorption feature which is caused
by a high density region. In order to test the density range
for which the presence of O vi is required to reproduce the
observed τOVI,app − τLyα relation, we introduce a cut-off in
6 Matthew M. Pieri and Martin G. Haehnelt
Figure 4. As Figure 3. Left: nOvi/nH i increases from top to bottom. Right: (n/¯n)cut increases from top to bottom.
the density, (n/¯n)cut below which we set the O vi density to
zero, i.e.
4.4 Varying S/N and continuum level
nOV I
nHI
= 0,
if
n
¯n
<(cid:16) n
¯n(cid:17)cut
,
nOV I
nHI
= const,
if
n
¯n
> (cid:16) n
¯n(cid:17)cut
.
(8)
(9)
In the right panel of Fig. 4 we vary (n/¯n)cut. The onset
of the correlation moves to larger H i optical depth. Unfor-
tunately, this dependence is weak.
The left panels of Fig. 5 show the effect of changing the sim-
ulated noise. As expected, the errors increase significantly
with increasing noise level.
The effect of an error in the continuum fit is shown in
the right panels of Fig. 5, where we have raised and lowered
the continuum as a whole by 1% in each synthetic spectrum.
When we raise the continuum level the apparent O vi optical
depth increases because the optical depth of spuriously coin-
cident H i absorption increases. Note that if the continuum is
placed too low the correlation between H i and O vi optical
depth misleadingly appears to extend to smaller H i optical
depth.
c(cid:13) 2003 RAS, MNRAS 000, 1 -- 11
Pixel correlation searches for O VI in the Lyman α forest
7
Figure 5. As Figure 3. Left: noise increases from top to bottom. Right: the continuum level increases from top to bottom.
5 SEARCH FOR O vi IN THE LOW DENSITY
IGM
5.1 Comparison of real and simulated spectra
Here we perform a detailed comparison between four ob-
served QSO spectra and sets of synthetic spectra. We use
the same S/N, ¯DOVI, ¯DHI and wavelength range as in the
observed spectra with one exception. In the case of Q1122
we decrease ¯DOVI by 35% in order to reproduce the level
of apparent τOvi at low τHi. The need for this is most likely
due to a deficit of saturated regions in our synthetic spec-
tra. The contribution of saturated regions to the mean flux
decrement is high at low redshift and the failure to ade-
quately reproduce the high incidence of these systems leads
to a bias in the mean flux decrement (Viel et al. 2003).
The triangles in the left panel of Fig. 6 show the
τOVI,app − τLyα relation for the four observed spectra. For
the synthetic spectra we assume as before that
the
O vi distribution takes the form described in equations (8)
and (9). We have varied the O vi density and the overden-
sity threshold for addition of O vi over the range nOVI/nHI =
0 − 0.3 and (n/¯n)cut = 0 − 100. We have produced samples
of 40 synthetic spectra for each set of values. We have then
calculated the τOVI,app − τLyα relation for each sample and
assessed the agreement with the observed relation by cal-
culating χ2. Note, however, that there may be additional
c(cid:13) 2003 RAS, MNRAS 000, 1 -- 11
sources of error such inhomogeneities in the metal distribu-
tion which our Monte Carlo technique does not take into
account. The crosses and squares in the left panels of Fig. 6
show the τOVI,app − τLyα relation with no O vi and the best
fitting values for nOVI/nHI ((n/¯n)cut = 0), respectively. The
best fitting value has a reduced χ2
r close to or somewhat
smaller than one. There is thus good agreement between
the τOVI,app − τLyα relation for our best fitting simulations
and the real data. In the right panels of Fig. 6 we use the
τOVI,app − τLyα relation of synthetic and observed spectra
for Q1107 to provide an example of the difficulty in con-
straining (n/¯n)cut .
The top panel of Fig. 7 shows the reduced χ2
r as a func-
tion of nOVI/nHI for the four observed QSOs. In Q1442 there
is a marginal detection of O vi, in Q1107 O vi is detected
with a poorly constrained O vi density (nOVI/nHI ≈ 0.08)
while in Q1122 O vi is clearly detected with nOVI/nHI ≈
0.06. In Q1422 no O vi is detected. Indeed, we find that the
spectrum of Q1422 is inconsistent with O vi absorption at
the same level as detected in Q1122 and Q1107 with a con-
fidence of greater than 99%.
In the bottom panel of Fig. 7 we show how the reduced
χ2
r varies as a function of (n/¯n)cut.We have assumed that
nOVI/nHI and (n/¯n)cut are independent in the calculation of
the confidence level. (n/¯n)cut of greater than 4 for Q1122,
7 for Q1107 and 4 for Q1442 are ruled out with 95% confi-
8 Matthew M. Pieri and Martin G. Haehnelt
Figure 6. Right: the τOVI,app − τLyα relation for observed and simulated synthetic spectra for four QSOs; this is shown for synthetic
spectra with both the best best fitting constant ratio of nOvi/nHi and with no O vi . Left: the effect of varying (n/¯n)cut on synthetic
spectra of Q1107 compared to the τOVI,app − τLyα relation of the observed spectrum.
c(cid:13) 2003 RAS, MNRAS 000, 1 -- 11
Pixel correlation searches for O VI in the Lyman α forest
9
Figure 7. Results of a χ2
ulated and observed τOvi. The top panel shows χ2
nOvi/nHi. The bottom panel shows χ2
χ2
r test of the agreement between sim-
r for varying
r for varying (n/¯n)cut . The
r values for no O vi and a constant nOvi/nHi are also shown.
dence. There is thus no significant detection of O vi at over-
densities <
∼ 5.
5.2 Other searches for O vi absorption in the
observed sample
Schaye et al. (2000) have claimed that the correlation of
apparent O vi optical depth and H i extends to τHi ∼ 0.1
in the case of Q1122, Q1107 and Q1442. They have in-
terpreted this as a detection of O vi in underdense re-
gions with n/¯n ∼ 0.3 − 0.5. We cannot confirm this
with our more detailed analysis. Dav´e et al. (1998) found
that Q1422 is consistent with no O vi in a search for
weak O vi absorption lines consistent with our result.
Note that they detected O vi in a O vi-C iv pixel corre-
lation search which probed large Lyα optical depth. In
the spectrum of Q1442, Simcoe, Sargent & Rauch (2002)
found two strong O vi systems associated with high col-
umn density H i absorption systems but no weak ab-
sorption systems consistent with our marginal detection.
Carswell, Schaye & Kim (2002) found a large number of
O vi absorption systems associated with low and high col-
c(cid:13) 2003 RAS, MNRAS 000, 1 -- 11
Figure 8. The volume filling factor (top panel) and the mass
fraction (bottom panel) as a function of gas overdensity for a
lognormal density PDF at the redshift of the four observed QSOs.
umn density H i absorption systems in the spectrum of
Q1122 again consistent with our results.
5.3 Implications for the metal enrichment of the
IGM
The lack of a significant detection of O vi at low densities
cannot be used as evidence against the presence of these
metals. As discussed extensively in the previous section,
O vi at these densities may be masked by H i absorption.
Also, at high redshift the spectrum of the UV background
may not be hard enough to ionise oxygen up to O vi. The
decreasing contamination by H i absorption (due to an ex-
pected lower ¯τLyα) is likely to be the main reason why the
lowest redshift QSO in the sample is the only one for which
we clearly detect O vi at moderate overdensities.
Fig. 8 shows how the volume filling factor depends
on overdensity for the lognormal density PDF in the red-
shift ranges of the observed QSO absorption. For this
PDF the 95% confidence limits for the density threshold,
(n/¯n)cut, translate into lower limits of 4%, 1.5% and 4% for
the volume filling factor of O vi for Q1122 (z = 2.0 − 2.3),
Q1107 (z = 2.7 − 3.0) and Q1442 (z = 2.5 − 2.6), respec-
tively. There appears to be no evidence for a large volume
filling factor of metals from the observed oxygen absorp-
tion. A lower limit of 1.5 − 4% for the volume filling factor is
no larger than that inferred for winds from Lyman break
10 Matthew M. Pieri and Martin G. Haehnelt
galaxies (Adelberger et al. 2003). A picture where metal
enrichment is due to winds from rather large galaxies at
z ∼ 2 − 5 is consistent with the observed O vi absorption
in QSO spectra. Haehnelt (1998) and Pettini et al. (2003)
find that the same is true for the observed C iv absorption,
although Schaye et al. (2003) may find new evidence of rel-
evance.
Fig. 8 also shows the relation between mass fraction and
overdensity for the lognormal density distribution. Despite
the rather low inferred volume filling factor for which metals
have been detected, about 20-40 % of baryons must already
be enriched with metals to explain the observed absorption.
6 CONCLUSIONS
We have used a large sample of synthetic spectra which
mimic the observational properties of four observed QSO
spectra to interpret the results of a search for O vi in the
low density IGM. Our results can be summarised as follows.
(1) At low Lyα optical depth the corresponding appar-
ent O vi optical depth obtained by using the pixel corre-
lation technique developed by Cowie & Songaila (1998) and
Schaye et al. (2000) depends mainly on the mean flux decre-
ment in the O vi region of the spectrum.
(2) In a sample of four QSO in the redshift range 2-3.5
we detect O vi significantly in two QSOs and marginally in
the third. The significance of the detection increases with
decreasing redshift.
(3) The position of the bend in the relation of apparent
O vi and Lyα optical depth depends only weakly on the low-
est density at which OVI is present in the IGM.
(4) We obtain upper limits of 4, 7 and 4 for the min-
imum density for which O vi has been detected with 95%
confidence. For the lognormal model density distribution
this translates into lower limits of 4%, 1.5% and 4% for
the volume filling factor of metals. We thus do not confirm
previous claims of a detection of O vi in underdense regions
and a corresponding large volume filling factor of metals.
(5) The O vi absorption in QSO absorption spectra as de-
tected by the pixel correlation technique provides no ev-
idence for (or against) a widespread metal enrichment at
very high redshift (z ∼ 10 − 20). The lower limit for the vol-
ume filling factor of metals is equally consistent with metal
enrichment by winds from Lyman break galaxies.
ACKNOWLEDGEMENTS
We would like to thank Michael Rauch, Len Cowie and ESO
for providing the observed spectra and the referee Anthony
Aguiree for a helpful report. The authors would further
like to thank Steve Warren for his useful suggestions and
Alex King and Thomas Babbedge for helpful comments on
the manuscript. This work was supported by the European
Community Research and Training Network "The Physics
of the Intergalactic Medium"
REFERENCES
Adelberger K. L., Steidel C. C., Shapley A. E., Pettini M.,
2003, ApJ, 584, 45
Aguirre A., Hernquist L., Schaye J., Weinberg D. H., Katz
N., Gardner J., 2001, ApJ, 560, 599
Aguirre A., Schaye J., Theuns T., 2002, ApJ, 576, 1
Barkana R., Loeb A., 1999, ApJ, 523, 54
Barkana R., Loeb A., 2001, Phys. Rep., 349, 125
Barlow T. A., Sargent W. L. W., 1997, AJ, 113, 136
Bi H., Davidsen A. F., 1997, ApJ, 479, 523
Bi H. G., Boerner G., Chu Y., 1992, A&A, 266, 1
Carswell B., Schaye J., Kim T.-S., 2002, ApJ, 578, 43
Couchman H. M. P., Rees M. J., 1986, MNRAS, 221, 53
Cowie L. L., Songaila A., 1998, Nat, 394, 44
Cowie L. L., Songaila A., Kim T.-S., Hu E. M., 1995, AJ,
109, 1522
Dav´e R., Hellsten U., Hernquist L., Katz N., Weinberg
D. H., 1998, ApJ, 509, 661
Dekel A., Silk J., 1986, ApJ, 303, 39
Efstathiou G., Bond J. R., White S. D. M., 1992, MNRAS,
258, 1P
Ellison S. L., Songaila A., Schaye J., Pettini M., 2000, AJ,
120, 1175
Ferrara A., Pettini M., Shchekinov Y., 2000, MNRAS, 319,
539
Gnedin N. Y., 1998, MNRAS, 294, 407
Gnedin N. Y., Hui L., 1998, MNRAS, 296, 44
Gnedin N. Y., Ostriker J. P., 1997, ApJ, 486, 581
Haehnelt M. G., 1998, in ASP Conf. Ser. 146: The Young
Universe: Galaxy Formation and Evolution at Intermedi-
ate and High Redshift Probing beyond the epoch of galaxy
formation. p. 249
Haiman Z., Madau P., Loeb A., 1999, ApJ, 514, 535
Hellsten U., Hernquist L., Katz N., Weinberg D. H., 1998,
ApJ, 499, 172
Hui L., Gnedin N. Y., 1997, MNRAS, 292, 27
Kaiser N., Peacock J. A., 1991, ApJ, 379, 482
Kim T.-S., Cristiani S., D'Odorico S., 2001, A&A, 373, 757
Madau P., Ferrara A., Rees M. J., 2001, ApJ, 555, 92
Nath B. B., Trentham N., 1997, MNRAS, 291, 505
Nusser A., Haehnelt M., 2000, MNRAS, 313, 364
Peebles P. J. E., 1980, The Large-Scale Structure of the
Universe. Princeton Univ. Press, Princeton, NJ
Pettini M., Madau P., Bolte M., Prochaska J. X., Ellison
S. L., Fan X., 2003, preprint, (astro-ph/0305413)
Rauch M., Haehnelt M. G., Steinmetz M., 1997, ApJ, 481,
601
Schaye J., Aguirre A., Kim T.-S., Theuns T., Rauch M.,
Sargent W. L. W., 2003, preprint, (astro-ph/0306469)
Schaye J., Rauch M., Sargent W. L. W., Kim T.-S., 2000,
ApJ, 541, L1
Simcoe R. A., Sargent W. L. W., Rauch M., 2002, ApJ,
578, 737
Theuns T., Mo H. J., Schaye J., 2001, MNRAS, 321, 450
Tytler D., Fan X.-M., Burles S., Cottrell L., Davis C., Kirk-
man D., Zuo L., 1995, in QSO Absorption Lines, Pro-
ceedings of the ESO Workshop Held at Garching, Ger-
many, 21 - 24 November 1994, edited by Georges Meylan.
Springer-Verlag Berlin Heidelberg New York. Also ESO
Astrophysics Symposia, 1995., p.289 Ionization and Abun-
dances of Intergalactic Gas. p. 289
c(cid:13) 2003 RAS, MNRAS 000, 1 -- 11
Pixel correlation searches for O VI in the Lyman α forest
11
Viel M., Haehnelt M. G., Carswell R. F., Kim T. S., 2003,
preprint, (astro-ph/03068078)
Vogt S. S. et al. 1994, in Proc. SPIE Instrumentation in
Astronomy VIII Vol. 2198, HIRES: the high-resolution
echelle spectrometer on the Keck 10-m Telescope. p. 362
c(cid:13) 2003 RAS, MNRAS 000, 1 -- 11
|
astro-ph/9508062 | 1 | 9508 | 1995-08-15T20:33:35 | Occlusion Effects and the Distribution of Interstellar Cloud Sizes and Masses | [
"astro-ph"
] | The frequency distributions of sizes of ``clouds" and ``clumps" within clouds are significantly flatter for extinction surveys than for CO spectral line surveys, even for comparable size ranges. A possible explanation is the blocking of extinction clouds by larger foreground clouds (occlusion), which should not affect spectral line surveys much because clouds are resolved in velocity space along a given line of sight. We present a simple derivation of the relation between the true and occluded size distributions, assuming clouds are uniformly distributed in space or the distance to a cloud comples is much greater than the size of the complex. Because the occlusion is dominated by the largest clouds, we find that occlusion does not affect the measured size distribution except for sizes comparable to the largest size, implying that occlusion is not responsible for the discrepancy if the range in sizes of the samples is large. However, we find that the range in sizes for many of the published observed samples is actually quite small, which suggests that occlusion does affect the extinction sample and/or that the discrepancy could arise from the different operational definitions and selection effects involved in the two samples. Size and mass spectra from an IRAS survey (Wood \etal\ 1994) suggest that selection effects play a major role in all the surveys. We conclude that a reliable determination of the ``true" size and mass spectra of clouds will require spectral line surveys with very high signal-to-noise and sufficient resolution and sampling to cover a larger range of linear sizes, as well as careful attention to selection effects. | astro-ph | astro-ph | OCCLUSION EFFECTS AND THE DISTRIBUTION OF
INTERSTELLAR CLOUD SIZES AND MASSES
John Scalo and A. Lazarian
Astronomy Department, University of Texas, Austin, TX 78712-1083
Received
;
accepted
Submitted to the Astrophysical Journal
5
9
9
1
g
u
A
5
1
1
v
2
6
0
8
0
5
9
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
– 2 –
ABSTRACT
The frequency distributions of sizes of “clouds” and “clumps” within clouds
are significantly flatter for extinction surveys than for CO spectral line surveys,
even for comparable size ranges. A possible explanation is the blocking of
extinction clouds by larger foreground clouds (occlusion), which should not
affect spectral line surveys much because clouds are resolved in velocity space
along a given line of sight. We present a simple derivation of the relation
between the true and occluded size distributions, assuming clouds are uniformly
distributed in space or the distance to a cloud comples is much greater than the
size of the complex. Because the occlusion is dominated by the largest clouds,
we find that occlusion does not affect the measured size distribution except for
sizes comparable to the largest size, implying that occlusion is not responsible
for the discrepancy if the range in sizes of the samples is large. However, we find
that the range in sizes for many of the published observed samples is actually
quite small, which suggests that occlusion does affect the extinction sample
and/or that the discrepancy could arise from the different operational definitions
and selection effects involved in the two samples. Size and mass spectra from
an IRAS survey (Wood et al. 1994) suggest that selection effects play a ma jor
role in all the surveys. We conclude that a reliable determination of the “true”
size and mass spectra of clouds will require spectral line surveys with very high
signal-to-noise and sufficient resolution and sampling to cover a larger range of
linear sizes, as well as careful attention to selection effects.
Subject headings: molecular clouds, interstellar medium
– 3 –
1.
Introduction
The mass spectrum of density fluctuations, defined in various operational ways as
“clouds”, is an important function that must be related to the processes by which clouds
form and evolve and to the mass spectrum of stars that form within these clouds. A fairly
large and growing number of studies of (mostly) molecular clouds yield a differential mass
spectrum which, if fit by a power law, has a form f (m) ∼ m−γ , with γ ∼ −1.5 ± 0.2. These
studies are primarily based on masses derived from column densities inferred from 13CO
spectral line observations and linear size. Some of the results are for surveys that cover a
significant area of the Galactic disk, e.g. the second quadrant survey of Casoli et al. (1984),
who find γ = −1.4 to −1.6 in both the Perseus and Orion arms, the 12CO first quadrant
survey by Sanders et al. (1985) who find γ ∼ −1.6 using virial masses, and the recent
comparison of 204 inner- and outer- galaxy molecular clouds by Brand and Wouterloot
(1995), who find γ = −1.6 for outer Galaxy clouds and γ = −1.8 for all 204 clouds. These
surveys together cover a mass range from about 100 M⊙ to over 106 M⊙ , although each
individual study generally covers a much smaller mass range over which a power law is
an adequate fit. Other work has concentrated on the mass spectrum of “clumps” within
individual clouds complexes, and find similar mass spectra in regions as different in star
formation properties as the Maddalena-Thaddeus cloud (Williams, deGeus and Blitz 1994),
which shows no evidence for star formation, the ρ-Oph core region (Loren 1989 as revised
by Blitz 1993), which is forming low- to intermediate- mass stars, the Rosette Molecular
Cloud (Williams et al. 1994, Williams and Blitz 1995), the M17SW cloud (Stutzki and
Gusten 1990), and the Orion region (Lada, Bally, and Stark 1991, Tatematsu et al. 1993),
all of which are actively forming stars up to large masses, and even lower-mass clumps
in MBM 12, a molecular cloud that is not gravitationally bound (Pound 1994). All these
studies give γ ∼ −1.5 ± 0.3 (the flattest being the Williams and Blitz result for the Rosette
cloud, with γ ∼ −1.3).
However there is a probable discrepancy when these results are compared with studies
of the mass spectra of clouds derived using extinction surveys, which are also based on
masses from sizes and column densities. If the distribution of sizes is given by f (r) ∼ rα
and the cloud internal density n is related to size by n ∝ rp , then the mass spectrum is
f (m) ∼ mγ with γ = (α − p − 2)/(3 + p). Estimates of p are uncertain and vary from (at
least) p ∼ −1.2(see Scalo 1985, 1987) to p = 0 (sizes or masses uncorrelated with density,
e.g. Casoli et al. 1984, Williams and Blitz 1995), or that the correlation is at least in part
an artifact due to selection effects (Scalo 1990). The spectral line studies mentioned earlier
give values of α around -2 to -2.5, based either on the published size data when available or
on the above transformation between mass and size spectra.
– 4 –
Scalo (1985) presented the frequency distribution of angular surface areas of dark clouds
from the catalogues of Lynds (1962) and Khavtassi (1960). The resulting size spectrum, if
fit by a power law, has α ∼ −1.4 ± 0.2, much flatter than the size distributions inferred
from the spectral line surveys. The implied mass spectra (γ ∼ −1.2 ± 0.3) seem significantly
flatter than the spectral line mass spectrum, but, because of the above relationship between
size and mass spectra (which gives ∆α = (2 − 3)∆γ for p = −1 to 0), and because size is
a directly measured quantity in both types of studies, the discrepancy is more clearly seen
in the size spectrum. Feitzinger and Stuwe (1986) studied the statistics of the combined
sample of Lynds clouds and their own Southern dark cloud survey, and found a distribution
of areas proportional to (area)−1 . The corresponding size spectrum has α = −1. This
gives a mass spectrum index of -1 for any p. Thus the discrepancy with the molecular line
survey size or mass spectra is even larger. Other published studies of mass spectra based
on extinction are not so clear, but point in the same direction, especially for lower-mass
clumps (e.g. Bhatt et al. 1984) for Lynds clouds in Orion, ρ Oph, and Taurus. Drapatz and
Zinnecker (1981) give size and mass spectra for several samples based on both extinction
and CO.
In the present paper we examine the possibility that this flatter size spectrum seen
in extinction is due to the effects of occlusion (smaller clouds being hidden behind large
clouds) on the extinction studies; this effect would not affect the spectral line studies nearly
as much because in that case two clouds along the same line of sight can be distinguished
in velocity space. (Of course occlusion in velocity space can also occur; we discuss this
briefly in § 3 below.) We derive an expression for the real size distribution of clouds in
terms of the measured distribution that is affected by binary occlusion and derive the range
of parameters over which the difference in size spectra between the two approaches can be
reconciled.
A relation between the distribution of physical sizes of clouds and their angular sizes is
established in § 2, while a relation between the actual distribution of angular sizes and the
distribution measured in the presence of occlusion is presented in § 3.
2. “Apparent” sizes of clouds
Consider that N1 (l) is the ‘real’ size distribution and N2 (θ) is the “apparent” angular
size distribution, without accounting for occlusion. In this seciton we define the relation
between N1(l) and N2 (θ). This problem is similar to the one discussed in Feitzinger and
Stuwe (1986). Due to the geometry of diverging lines of sight, clouds with the same
– 5 –
physical size but at different distances from the observer will fall into different ranges of
apparent angular sizes. As a first approximation, assume the “true” properties of clouds
to be independent of the distance from the observer. This is probably reasonable for
observations in the galactic plane and of nearby individual cloud complexes (e.g. Taurus,
Oph, Chameleon, Orion,. . .)
In our model, the distribution of clouds at distance r is given by the product (D)N1 (l),
where (D) is the total density of clouds at distance D, and we take N1(ℓ) normalized to
unity. Then the number of clouds within the distance interval D , D + dD is (D)ωD2dD ,
where ω is the solid angle. Within this volume, the clouds with sizes from Dθ to (θ + dθ)D ,
where θ is the angular size of clouds, will contribute to the apparent angular cloud
distribution N2(θ). The total number of “pro jections” with angular sizes (θ, θ + dθ) within
the solid angle ω can be found by integrating (D)ωD2N1(Dθ)DdθdD over the line of sight.
Therefore,
N2 (θ)ωdθ = ω Z Dmax
Dmin
A change of variables Dθ = x results in
1
x3 (cid:18) x
θ4 Z θDmax
θ (cid:19) N1 (x)dx
θDmin
Assuming (x) = constant, differentiation gives
1
Dmax (cid:16)N2 (θ)θ4(cid:17)′
= θ3D3
maxN1 (θDmax)
D3(D)N1 (Dθ)dDdθ
N2 (θ) = −
(1)
(2)
(3)
− θ3D3
minN1 (θDmin )
ℓmin (cid:17)−γ
For power law N1 (ℓ) = N1(ℓmin ) (cid:16) ℓ
, the first term is the most important if γ < 3,
whereas if γ > 3, the second term dominates. The cases of greatest interest here have γ < 3.
Whenever the second term is negligible and N1(ℓ) is a power-law distribution, N2 (θ) is also
power-law with the same index.
Similarly for Dmin = 0,
N1 (θDmax) =
1
θ3D4
max
(N2(θ)θ4 )′
(4)
and for any other Dmin , the power-law distribution N1(l) entails a power-law distribution
N2 (θ) with equal slope. Therefore the index of the size distribution is not affected by the
differing distances of the clouds in the sample, and the index of the angular size distribution
is the same as the index of the linear size distribution. An exception occurs for a delta
function linear size distribution, i.e. when all clouds have the same size. In that case the
– 6 –
apparent angular size distribution varies as θ−4 (see Bhatt et al. 1984). In what follows, we
therefore identify N2(θ) with the “real” distribution of sizes, with the understanding that
clustering of clouds and gradients in the number density of clouds with distance could alter
this identification. Obviously, if the distance to a cloud complex is much greater than the
extention of the complex, statistics of the “real” size distribution and the “apparent”angular
size distribution coinside.
3. Occlusion effect
N2 (θ) is the pro jected apparent angular size distribution of clouds when occlusion
is ignored; i.e.
it is the angular size distribution corresponding to the “real” linear size
distribution. If occlusion is “switched on,” some of smaller clouds are hidden behind (or
in front) of bigger ones. Let N3 (θ) be the distribution of pro jections in the presence of
occlusion. Then
πN3(θ)θ2ωdθ
(5)
is the angular area covered by cloud pro jections with sizes within the range θ, θ + dθ. The
part of the sky not covered by cloud pro jections with angular sizes greater than θ is
π
A Z θu
θ
where A is the angular area covered by the survey and θu is the upper size limit for the
sample. Therefore the number of cloud pro jections of angular size θ that are not occluded
by larger clouds is
N3 (x)x2dx
1 −
(6)
N2 (θ)ωdθ 1 −
N3(x)x2dx!
π
A Z θu
θ
Since this is the number of clouds that is seen, equating this to N3(θ)ωdθ gives
N3(x)x2dx!
N3 (θ) = N2 (θ) 1 −
π
A Z θu
θ
The real size distribution N2(θ) can therefore be derived from the apparent (occluded size
distribution N3(θ) from
(7)
(8)
N3(θ)
A R θu
1 − π
θ N3 (x)x2dx
The second term in the denominator is just the fraction of the survey area A covered by
clouds with sizes greater that θ. The largest value this fraction can have occurs at θ = θℓ ,
N2(θ) =
(9)
– 7 –
the minimum size detected in the survey, for which the second term is the total area filling
factor of clouds detected in the survey (< 1).
It is also possible to derive the observed distribution N3 (θ) that would result from a
given real distribution N2 (θ), as shown in the Appendix. However, that formulation is not
as useful for the purposes of the present paper because the solution involves the unknown
properties of the real distribution.
To illustrate the properties of the N2 − N3 relation, assume that the observed occluded
distribution is a power law, N3(θ) = c3θ−γ3 . Then
N2(θ) =
c3θ−γ3
c3
(3−γ3 ) (θ−r3+3
u
1 − π
A
− θ−γ3+3)
(10)
f3,tot =
(11)
N3(θ)πθ2dθ =
The total areal filling fraction is
1
πc3
A Z θu
(cid:17) .
A(3 − γ3) (cid:16)θ−γ3+3
u
θℓ
The second term is negligible for θℓ ≪ θu and γ3 < 3. So, from eqs. (10) and (11), we see
that for θ significantly smaller than θu , N2 (θ) = N3(θ)/(1 − f3,tot ); i.e. for small clouds the
real number of clouds is larger than the observed number by a factor (1 − f3,tot ), but the
power law index is unaffected. The probability of a small cloud to be hidden by a large
cloud is independent of its size if its size is much smaller than θu because the areal filling is
dominated by the largest clouds (if γ3 < 3).
− θ−γ3+3
ℓ
To see this more clearly, consider the local logarithmic slope of the real distribution at
size θ (i.e. the exponent of a local power law fit at that size). From eq. (10) we obtain
γ2(θ) =
dℓnN2(θ)
dℓnθ
πc3
A
= γ3 +
θ−γ3+3
h1 − πc3
A(3−γ3 ) (cid:16)θ−γ3+3
u
≡ γ3 + ∆γ (θ)
− θ−γ3+3(cid:17)i
(12)
The maximum value of the change in exponent ∆γ (θ) occurs for θ near θu , at which size
∆γ (θ) = πc3 θ−γ3+3
/A ≈ (3 − γ3)f3,tot (for θℓ ≪ θu and γ3 < 3). If f3,tot ≈ 0.5, as is
u
typical for dark cloud surveys (not selected according to opacity class or size), then the
dark cloud power law γ3 ∼ 1.4 gives ∆γ ≈ 1.6f3,tot ∼ 0.8. While this is about the value
needed to reconcile the extinction size distribution with the spectral line size distribution,
it only occurs very close to θu . At smaller θ, say xθu (x < 1), ∆γ is reduced by a factor of
x−γ3+3 ∼ x1.6 for the parameters chosen. So even for clouds half or a third of the size of
the largest clouds, ∆γ is too small to account for the discrepancy, and for x = 0.1, ∆γ is
essentially negligible.
– 8 –
The same considerations hold even if the observed distribution N3 (θ) is not a power
law, as long as it is not locally too steep (γ > 3): the real size distribution tracks the
observed distribution (although at larger amplitude) except for sizes close to θu , at which
sizes the real distribution is steeper than the observed distribution.
We would be tempted to conclude that occlusion cannot account for the discrepancy,
except for the fact that the range in sizes in the observed surveys is actually quite small. For
both types of surveys, the cloud masses are proportional to the square of some characteristic
size times a column density, so the range in sizes, which is a directly observed datum, only
corresponds to the square root of a given range in the masses (which is what is usually
displayed). Since, for the published spectral line surveys of clumps within cloud complexes,
power laws are only good fits over a limited mass range (limited by small numbers at the
largest masses and resolution incompleteness and other effects at small masses), usually a
factor of 10–100, the range in sizes is not very large. The range in sizes for a few early
surveys is listed in Drapatz, and Zinnecker (1984). The range in sizes for the line surveys
of Stutzki and Gusten (1990), Lada et al. (1991), Tatematsu et al. (1993), Williams et al.
(1984) and Williams and Blitz (1995) is less than a factor of 10, although the range in mass
used to derive the mass spectra is larger in some of the surveys. This suggests that the
mass distributions derived from spectral line surveys will be very sensitive to the definition
of, and systematic uncertainties in the measurement of, cloud sizes. For the extinction
sample the range of sizes over which the power law size spectrum is applicable is less than a
factor of about 10 in all cases, even for the full sample of the Lynds and Khavtassi surveys,
and various selection effects come into play at smaller and larger masses (see Scalo 1985, §
III.B.2. for a discussion).
Thus we conclude that the discrepancy between size distributions derived from
extinction surveys and spectral line surveys may be due to occlusion effects in the extinction
surveys because the minimum size is not much smaller than the maximum size in both
types of surveys, or because of different operational definitions of size in the two types
of surveys. Actually these two possibilities are not independent because the size range is
related to how clouds are defined. It is worth pointing out that in some of the spectral line
surveys the noise level is so large that the surveys are really only observing the “tips of the
mountain range” if the column density map is thought of as a 2-dimensional surface with
height equal to column density. For example, in the Rosette data (Blitz and Stark 1986),
the rms noise is only about a factor of 2 smaller than the average peak line temperature, so
the cloud sizes may be severely affected. For the dark clouds, identification of the cloud
boundary is usually much less affected by “noise” (in this case fluctuation in star densities),
except for the lowest-opacity clouds. Thus even though the line surveys have the advantage
of separating clouds in velocity space it is not clear that they give more realistic size and
– 9 –
mass spectra compared to extinction surveys.
Evidence that the empirical cloud mass spectra are sensitive to selection effects comes
from the following two examples.
Clemens and Barvainis (1988) compiled a catalogue of isolated small dark cloud
(“globules”) identified on POSS plates and compiled properties based on their CO
observations. For clouds with mean size larger than 3.5 arcmin (smaller size clouds are
probably affected by incompleteness), we can fit the frequency distribution of angular sizes,
and hence linear sizes if the clouds are uniformly distributed in distance, by f (r) ∼ r−2 ,
which gives a power law mass spectrum with γ = −1.5 to -1.7 for p = −1 or 0. These
clouds were selected to be small and isolated, so occlusion should not be important. Since
this result agrees with the molecular line surveys, it suggests that the flatter size and mass
spectra derived from general extinction surveys are products of occlusion effects, if selection
effects are unimportant in the estimation of properties from CO.
However the survey of 255 IRAS cloud cores by Wood, Myers, and Daugherty (1994),
which derives sizes and masses based on IRAS 100 µm optical depth for clouds with AV >∼ 4
mag, yields a frequency distribution of areas f (A) ∼ A−0.54 , or f (r) ∼ r−0.08 , which is
extremely flat compared to not only the molecular line surveys, but even extinction surveys.
For constant column density, as Wood et al. find, p = −1, so f (m) ∼ m−0.54 , consistent
with their directly determined (from individual areas and column densities) f (m) ∼ m−0.49 .
The sizes and masses for the fits have ranges of well over 1000. Since all the cores are
optically thin at 100 µm, occlusion cannot be a factor; a small core behind a larger core
would be seen as a column density enhancement of about a factor of two, because all the
cores in the sample apparently have about the same column density. This result suggests
that all the surveys, whether based on extinction, molecular line, or IRAS, are affected by
selection effects.
4. Velocity occlusion
The same argument used above for purely spatial occlusion can be somewhat extended
to include the effects of blending in velocity space for spectral line surveys.
In this
illustrative example we assume that each identified “cloud” or “clump” (for convenience
we use the latter term in what follows) has an internal velocity dispersion ∆v (θ) which is
strictly correlated with the size of the clump, as found in several surveys, at least for clumps
in which self-gravity is important. In that case the fraction of the total survey volume of
– 10 –
the data cube AV (A = area of the survey in the plane of the sky, V = radial velocity extent
of the survey) occupied by occluded clumps of size θ is
1
AV Z θu
θ
where Nv (θ) is the size distribution found in the (blended) survey. The real size distribution
is then
πθ2Nv (θ)∆v (θ)dθ
fv (θ) =
N2(θ) =
Nv (θ)
1 − fv (θ)
The maximum value of fv (θ) occurs at θℓ and is the total volume filling factor of observed
clumps in the data cube. Since this number is small for the surveys we are aware of (see
Fig. 7 in Williams and Blitz 1995), the effect of this type of occlusion (due to finite internal
velocity dispersion of the clumps) on the derived size distribution must be negligible, at least
for velocity resolutions much smaller than the minimum ∆v . However, this analysis does
not account for the fact that clumps with similar centroid velocities may lie along the same
line of sight. Taking this effect which probably dominates the blending in velocity space,
into account would involve calculting the probability that, for a prescribed centroid velocity
distribution, two clouds along a given line of sight have a centroid velocity difference smaller
than the sum of the linewidths of the two clouds (which is a function of θ), a calculation
which we postpone to a later publication.
5. Conclusions
Our study has examined the effect of occlusion on extinction surveys. The predicted
change in the shape of the frequency distribution of cloud sizes for extinction surveys
compared to spectral line surveys is small, except very near the maximum cloud size. Thus
the discrepancy between the empirical results for the two types of surveys probably cannot
be attributed to occlusion in the extinction survey, if the size range of both types of survey
is large. Howver an examination of the literature shows that many of the observed surveys
employ a very limited range of sizes. In these cases the discrepancy might still be due to
occlusion. On the other hand, some of the spectral line surveys do include clouds with a
fairly large range of sizes (e.g. Brand and Wouterloot 1995), and these surveys do find
size and mass spectra much steeper than the dark cloud results. Furthermore, the IRAS
cloud-core survey of Wood et al. (1995) gives size and mass spectra which are much flatter
than both the extinciton and line survey results. This suggests that the inferred shapes of
the size and mass spectra of clouds are affected by the manner in which clouds are defined
and by the selection and noise effects inherent in both types of surveys.
– 11 –
This work was supported by NASA grant NAG52773.
– 12 –
APPENDIX
Rather than solve for the real distribution function in terms of the observed (occluded)
distribution) it is possible to derive the observed distribution N3(θ) that would result from
a given real distribution N2 (θ). Differentiating eq. 9 with respect to θ gives
3(θ) = N3(θ) " πθ2
2 (θ) # N2(θ)
N ′
2(θ)
N ′
N 2
A
Integrating this equation, with a lower integration limit θℓ , gives
exp ( π
N2 (x)x2dx)
A Z θ
θℓ
We can obtain N3 (θℓ )/N2(θℓ ) by imposing the condition that the largest cloud in the sample
cannot suffer any occlusion, i.e. by substituting N3(θu ) = N2(θu ) at θ = θu in eq. A2. This
condition results in
N2 (θ)
N2 (θℓ )
N3 (θ)
N3 (θℓ )
(2)
(1)
+
=
(3)
N3(θℓ )
N2(θℓ )
= exp ( π
N2 (x)x2dx) = exp(−Atot /A),
A Z θu
θℓ
where Atot is now the total area covered by all clouds in the unoccluded (real) distribution,
and may be greater than the survey area A. Dividing the integral from θℓ to θu into parts
from θℓ to θ and from θ to θu and substituting into eq. A2 gives
N2(x)x2dx) = N2 (θ) exp[−A(> θ)/A],
N3(θ) = N2(θ) exp (−
π
A Z θu
θ
where A(> θ) is the area covered by clouds with sizes greater than θ in the unoccluded
distribution and may xxx by greater than A. For a power law N2(θ) = c2 θ−γ2 we find
N3(θ) = N2 (θ) exp (
− θ−γ2+3)) .
The local logarithmic slope of the predicted occluded distribution is then (assuming θℓ ≪ θu
and γ2 < 3)
θu !−γ2+3
A θ
Atot
dℓnN3 (θ)
dℓnθ
Once again we see that although the change in local logarithm slope may be large near θu ,
the effect becomes increasingly negligible for θ ≪ θu .
πc2
(3 − γ2)A
(θ−γ2+3
u
γ3(θ) =
= γ2 −
(6)
(4)
(5)
.
However this formulation is not as useful as that given in the main text (which expressed
N2 (θ) in terms of N3(θ)) because the total covering fraction of the real distribution is
unknown, although it can be evaluated for a model which specifies the total number of
clouds in the distribution (again unknown from observations).
– 13 –
REFERENCES
Bhatt, H. C., Rowse, D. P., & Williams, J. P. 1984, MNRAS, 209, 69.
Blitz, L. 1993, in Protostars & Planets III, ed. E. H. Levy & J. L. Lunine (Tucson: Univ.
Arizona Press), p. 125
Blitz, L. & Stark, A. A. 1986, ApJL, 300, L89
Brand, P. & Wouterloot 1995, AstrAp, in press
Casoli, F., Combes, F., & Gerin, M. 1984, AstrAp, 133, 99
Clemens, D.P. & Barvainis, R. 1988, Ap. J. Suppl, 68, 257
Drapatz, S. & Zinnecker, H. 1984, MNRAS, 210, 11P
Feitzinger, J. V. & Stuwe, J. A. 1986, ApJ, 305, 534
Khavtassi, D. Sh. 1960, Atlas of Galactic Dark Nebulae (Abastumani, USSR: Abastumani:
Astrophys. Obs.)
Lada, E. A., Bally, J., and Stark, A. A. 1991, Ap. J., 368, 432
Loren, R. B. 1989, ApJ, 338, 902
Lynds, B. T. 1962, ApJS, 7, 1
Pound, M. 1994, Ph.D. dissertation, Univ. of Maryland
Sanders, D. B., Scoville, N. Z., & Solomon, P. M. 1985, ApJ, 289, 373
Scalo, J. M. 1985, in Protostars & Planets II, ed. D. C. Black & M. S. Mathews (Univ.
Ariz. Press), p. 201
———. 1987, in Interstellar Processes, ed. D. J. Hollenbach & H. A. Thronson (Dordrecht-
Reidel), p. 349
———. 1990, in Physical Processes in Fragmentation & Star Formation, ed. R. Capuzzo-
Dolcetta et al. (Doredrecht: Kluwer), p. 151
Stutzki, J. and Gusten, R. 1990, ApJ, 356, 513
Tatematsu, K., Umemoto, T., et al. 1993, Ap. J., 404, 643
– 14 –
Williams, J. P. & Blitz, L. 1995, ApJ, in press
Williams, J. P., de Geus, E. J., & Blitz, L. 1994, ApJ, 428, 693
Wood, D.O.S., Myers, P. C., and Daugherty, D. A. 1994, ApJS, 95, 457
This manuscript was prepared with the AAS LATEX macros v3.0.
|
astro-ph/0211490 | 1 | 0211 | 2002-11-21T22:15:36 | Ground level muons in coincidence with the solar flare of April 15, 2001 | [
"astro-ph"
] | The counting rate of single muon tracks from the Project GRAND proportional wire chamber array is examined during the Ground Level Event (GLE) of April 15, 2001. The GLE was seen by neutron monitor stations shortly after the time of the solar X-ray flare. GRAND's single muon data are presented and compared with neutron monitor data from Climax, Newark, and Oulu. The single muon data have mean primary hadron energies higher than those of these neutron monitor stations and so contain information about higher energy hadrons. For the single muon data for Project GRAND, the GLE is detected at a statistical significance of 6.1-sigma. | astro-ph | astro-ph | Ground level muons in coincidence with the solar flare of April 15, 2001.
J. Poirier and C. D'Andrea
Center for Astrophysics, Department of Physics,
University of Notre Dame, Notre Dame, Indiana 46556 USA.
The counting rate of single muon tracks from the Project GRAND proportional wire chamber
array is examined during the Ground Level Event (GLE) of April 15, 2001. The GLE was seen
by neutron monitor stations shortly after the time of the solar X-ray flare. GRAND's single muon
data are presented and compared with neutron monitor data from Climax, Newark, and Oulu. The
single muon data have mean primary hadron energies higher than those of these neutron monitor
stations and so contain information about higher energy hadrons. For the single muon data for
Project GRAND, the GLE is detected at a statistical significance of 6.1σ.
2
0
0
2
v
o
N
1
2
1
v
0
9
4
1
1
2
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
I.
INTRODUCTION
On April 15, 2001 the IPS Radio and Space Services website reported an X14 solar flare with a maximum X-ray
intensity at 13.83 UT (8.83 hours EST or local time) [IPS Space Services, 2001]. The flare originated on the sun
at coordinates S20 W85 [NOAA Space Environment Center Website, 2001] and had an X-ray intensity of 14 × 10−4
W/m2 at one AU from the sun. Flares of this high intensity are rare. During the period from 1976 - 1998, only
sixteen flares of magnitude X10 or greater were reported. Observations on some of these high energy flares are given
in [Bieber et al., 2002], [Falcone, 1999] and [Swinson and Shea, 1990].
Data from the Project GRAND proportional wire chamber array are examined for a signal during the time of the
Ground Level Event (GLE) from this flare. During the peak time of the flare, the sun was at an elevation angle of
35◦ above GRAND's local horizontal plane and an azimuth of 110◦ from north. Even though the sun is near the
minimum detectable elevation angle for GRAND, the Interplanetary Magnetic Field (IMF) and the magnetic field of
the earth alter the direction of the protons from the sun. Charged particles from the sun follow a spiral path about
the IMF (which itself curves in an Archimedes spiral about the sun) and are additionally deflected on entering the
earth's magnetic field.
Therefore the GLE associated with a particular X-ray flare should occur at a slightly later time than the X-ray
event (due to the longer path length) and at a different angle than the sun, thus providing better detector acceptance
even though a direct line to the sun is at a low elevation angle (see also [Munakata et al., 2001]).
GRAND detects ground level muons originating from hadronic primaries with energies above the characteristic
energy of neutron monitors. The value of this characteristic energy depends upon the spectral index of the proton
flux from the sun during this GLE. The energy response function of Project GRAND is discussed in more detail in
Section 5. Typically, surface muon telescopes are sensitive to proton primary energies above about 4 GeV [Shea and
Smart, 2002].
II. EXPERIMENTAL ARRAY
Project GRAND is an array of 64 proportional wire stations located at 41.7◦ N and 86.2◦ W at an altitude of 220 m
above sea level. Each station contains four pairs of orthogonal proportional wire chamber planes; each plane of the
pair has 80 wire cells 14 mm wide by 19 mm high. Each pair of planes is positioned 200 mm vertically above its
neighbor. This geometrical arrangement allows for the direction of a muon track to be measured to within 0.26◦, on
average, in each of two projected angles: Θxz and Θyz. The angle Θxz is measured in the xz-plane from the zenith
(z) toward the east (x), Θyz is measured in the yz-plane from the zenith toward the north (y). A 50 mm thick steel
plate is located above the bottom pair of PWC planes. The steel plate, in conjunction with the PWC pair of planes
underneath, enables secondary muon tracks to be distinguished from those of electrons or hadrons. Each station has
a total active area of 1.29 m2; the array currently collects data at a rate of ∼ 2400 muons per second.
These data are read from the 64 stations in parallel to a central data acquisition system in 70 microseconds. Eight
computer nodes are used, each receiving an event (data from all 64 stations) in sequence. The node's task is to search
the 80 wires in each of 512 proportional wire planes and record the wire numbers in the eight planes of a station
which had one and only one hit in each plane of that station. The eight computer nodes operating in sequence are
able to conduct this search with little overall deadtime. Once a particular node has accumulated data on 900 muons
in its buffer memory, the entire buffer is written to magnetic tape as a single record. Upon later off-line analysis,
96% of this recorded data turn out to fit straight tracks. A radio signal from WWVB in Boulder, Colorado provides
time information with millisecond precision. This signal is used to determine the start time for the first and the last
2
muon event in each record. These times are then recorded along with the muon data for that record on magnetic
tape. Additional details on the Project GRAND detector are available on GRAND's webpage [Project GRAND, 2001].
Other muon detectors are described, for example, in [Manakata et al., 2001] and [Duldig, 2000].
III. DATA ANALYSIS
Project GRAND recorded a continuous data file on magnetic tape containing information from April 14, 2001
at 8:40 UT through April 16 at 21:30 UT, providing background information before and after the flare time. The
counting rate from 14.5 hours to 24.5 hours UT was examined in 0.1 hour bins. The data file is segmented into
records, each containing information on 900 muons. In order to find the time of a particular muon, the end time of
the current record is compared to the end time of the previous record from that computer node; the times for each of
the 900 muons is interpolated evenly within that interval. This interpolation of time should be quite adequate since
the average length of a record was 3.6 seconds and thus the 0.1 hour bin size for the data is 90,000 times longer than
the average time between muons.
A station can turn on or off, either deliberately for miscellaneous repairs (which were not always recorded), or due
to problems like, for example, high humidity coupled with a malfunctioning dehumidifier in a station. Deviations
(r.m.s.) from average counting rate were measured individually for each station of the array. In order to prevent
spurious individual station variations from giving a possible signal in the sum-of-all-stations, one-third of the stations
with the highest individual r.m.s. deviations were eliminated from consideration leaving 39 stations (an extremely
severe cut to ensure with great confidence that erratic station behavior could not cause a false signal). GRAND's
counting rate above background is shown as a function of time in Figure 1 (lower points with error bars) with the
data binned in 0.1 hour intervals.
The six bins of time from 14.0 to 14.6 hours UT were used as a signal time and background bins from 9.5 to 19.5
hours UT (minus the signal time) were used. Since it is known that the muon rate depends slightly on the time of day
[Poirier and D'Andrea, 2001a] (due to small solar effects in air pressure, temperature, and in the IMF), the counting
rate for the background time period was fit to a curve of the form A(1 + Bt + Ct2) with A = 496000, B = 0.00101
and C = 0.000241 where t is time measured from 14.3 hours UT. The smallness of the B and C coefficients compared
to one is a measure of the flatness of the muon counting rate during this time interval. The data as percentage above
background are shown in Figure 1 with error bars.
For comparison, the data from neutron monitors at Newark for this GLE [ J. Clem and R. Pyle, 2001] and Oulu
[Oulu Neutron Monitor, 2001] are also shown. The Climax neutron monitor data [C. Lopate, 2001] are almost
indistinguishable from that of Newark for this GLE and, for clarity, are not shown.
GRAND's muon data rate, although sensitive to the atmospheric pressure, has not been corrected for changes in
pressure. However, the pressure was recorded every half hour and found to be constant for the entire interval of time
analyzed to within ±0.15%. Thus, the pressure is unlikely to be a factor in the analysis. Indeed, during the time
of the flare the pressure actually rose slightly which would, if anything, depress the counting rate slightly during the
time of the GLE. Thus, pressure variations could not be a cause of the excess counting rate observed from 14.0 to
14.6 hours UT.
The muon counting rate for April 14 and April 16 (the preceding and the next day from the signal) are shown in
Figure 2 as well as the data containing the signal on April 15. The background curve fit to the April 15 data is also
included. It is seen that the background from the day before and the day after are similar to the background curve for
April 15; also, the average of four years of data show a similar shape. The zero of muon counting rate is suppressed
in order to better show these small deviations from an almost constant rate.
The amount of signal above background was determined by the excess of the six signal bins above the fitted
background curve. The signal in these six bins yields 10,708 counts above background with an error of 1762; this
error includes the error on the background fit (A) and uses the r.m.s. deviation of the background about the fitted
curve as an estimator for the rms error of the data in the signal region. This yields a Ground Level Event signal with
a significance of 6.1σ (where σ is the error on the signal).
Each of the 39 stations has an active area of 1.29 m2; correcting for those stations which were not 100% efficient
yields a total effective area of 42 m2 for this analysis. The excess counting rate was 255 muons/m2 or 0.36% of the
background muon counting rate during the 0.6-hour signal interval.
A. Can GRAND see gamma ray or neutron primaries from this GLE?
To determine if GRAND detects secondary muons from gamma ray [Poirier et al., 2001] or neutron primaries which
originate from the sun (rather than the protons studied above), the muon rate was analyzed inside a 12◦ by 12◦ square
e
g
a
t
n
e
c
r
e
P
60
50
40
30
20
10
0
-10
3
3
2.5
2
1.5
1
0.5
0
-0.5
9.5
10.5 11.5 12.5 13.5 14.5 15.5 16.5 17.5 18.5 19.5
Time (UT)
FIG. 1: Percentage counting rate increase above background versus time on April 15. Bottom points (circles) with error bars
are from GRAND with %-scale on the right; background is the curve in Fig. 2. Middle points (squares) are from Newark
Neutron Monitor (left scale). Top points (triangles) are from Oulu Neutron Monitor (left scale). The arrow denotes the onset
of the X-ray signal on the sun (at 13.83 hours UT [IPS Space Services, 2001]).
506000
504000
502000
500000
s
t
n
u
o
C
498000
496000
494000
492000
490000
488000
9.5
10.5 11.5 12.5 13.5 14.5 15.5 16.5 17.5 18.5 19.5
Time (UT)
FIG. 2: Single muon counting rates in 0.1 hour bins versus time with suppressed zero to emphasize the small anisotropies.
Bottom points (squares) are April 14. Middle points (circles) are April 15 with background curve fit and correspond to the
numbers on the left. Top points (triangles) are April 16. The April 14 and 16 data are offset for clarity. Note that the small
hourly background variations in time (±0.6%) are consistent over this three-day period.
angular window centered on the sun. During the signal time from 14.0 to 14.6 hours UT, 368 counts were measured
compared to an average of 376 background counts for the 0.6 hour interval of time as obtained during 9.5 to 19.5
hours UT (minus the signal time). The low counting rate in this angular window is caused by the sun's low elevation
angle at that time. A negative signal of -- 8 counts from background is observed inside this 12◦ by 12◦ window, or
-- 0.4 σ if compared to the error on the signal. Thus there is no evidence for direct gamma rays or neutrons from the
sun; as well, they cannot be the cause of the 6.1 σ excess obtained in the data with no angular cuts.
IV. ANGULAR DISTRIBUTION
4
The sky was divided into sixteen angular regions each with approximately equal counting rates in order to obtain
some information about the angular characteristics of the signal [Poirier and D'Andrea, 2001b]. Each individual region
of the sky was examined for the significance of a possible GLE signal. The significance of the GLE signal (in terms
of sigma) for each of the 16 windows is shown in Table 1. For these 16 calculations, the background was individually
fit to the same quadratic form as before with coefficients A, B, and C; however, due to the reduced amount of data
in a region (∼1/16th of the entire angular region), the coefficients B and C were fixed at the value obtained for the
whole sky and only the A parameter was varied in the fit to estimate the background under the signal region for each
of the 16 angular regions. The angular information presented in Table 1 could be used in later analyses in an overall
fit of this GLE using data from other muon detectors and neutron monitor stations around the world to obtain the
mean arrival direction, pitch angle distribution, and spectral index.
TABLE I: The significance (in terms of sigma) of the GLE signal for each region of the sky from 14.0 to 14.6 UT above the
background measured from 9.5 to 19.5 UT. Θxz is measured from the zenith toward the east; Θyz is measured from the zenith
toward the north. Local zenith is at the center of the table; to the right is east, to the top is north.
Θyz
Θxz
≤ -- 13◦ -- 13 to 0 0 to 13 ≥13◦
0.5
1.7
0.9
-- 1.0
1.1
1.4
1.4
-- 0.8
0.4
2.5
2.3
-- 0.1
0.6
3.7
3.3
-- 0.1
≥13◦
0 to 13
-- 13 to 0
≤ -- 13◦
An overall fit for each hemisphere (in a local horizon coordinate system) gives a signal strength of 4.8 σ for the
northern and 1.9 σ for the southern hemisphere, while a signal strength of 3.7 σ is obtained for the eastern and 3.4 σ
for the western hemisphere. These results show a definite preference for the northern hemisphere compared to the
southern and only a slight preference for the eastern hemishpere over the western. The mean location of the sun at
GRAND during the time of the GLE was θ
xz = +53◦ (eastward) and θ
yz = -- 26◦ (southward).
V. PRIMARY ENERGY SENSITIVITY
The response function of GRAND as a function of energy for primary protons yielding muons which reach detection
level is calculated by using the Monte Carlo program FLUKA [Fass´o et al., 2000]. The atmosphere in these simulations
is approximated by 50 layers of air beginning 80 km above sea level. The calculation is similar to that reported in
[Poirier et al., 2001] except that the primary particles were changed from the gamma rays of that reference to primary
protons for this calculation. Each primary proton is incident on the top of the atmosphere at zero degrees from zenith
angle (as well as three non-zero angles for the 3, 10, and 30 GeV energies) and all non-hadronic secondaries are
followed until they are below the threshold energy to produce pions or the charged tracks reach ground level. The
calculation is done for a series of fixed proton primary energies for energies from 1 GeV to 1 TeV. A summary of the
results of these calculations is contained in Table 2 for proton primaries incident at the top of the atmosphere at 0◦
from zenith. Figure 3 is a plot of this response function. Protons incident at 0◦ from zenith are shown by a dashed
line. Incident angles of 17◦, 34◦, and 51◦ from zenith (solid lines) show GRAND has progressively smaller responses
to angles which deviate from normal incidence. The deviation is fractionally larger at the lowest energy (3 GeV)
plotted.
L. Miroshnichenko [Miroshnichenko, 2002 preliminary] has calculated a preliminary fit to monitor stations at Ap-
atity, Moscow (IZMIRAN), Rome, and Athens to obtain the integral flux of protons from this GLE versus the energy
of the protons from 1 to 10 GeV. These values are plotted in Figure 4 as square points. The circular points below 0.1
GeV are proton fluxes obtained by the GOES satellite [GOES Website, 2001]..
To obtain the overall differential response expected of GRAND for this GLE, the response function for GRAND (at
0◦ in Table 2) is plotted in Figure 4 as a dashed line (the numbers on the ordinate are the same with units of muons
at detection level per proton of that primary energy). This response function is then folded with the flux of primary
protons in each energy interval to obtain the differential energy response of GRAND for this GLE (shown as diamond
points below 10 GeV). At energies ≥ 10 GeV, the proton intensity (shown by solid curves) was approximated using a
5
)
n
o
t
o
r
p
(
N
/
)
n
o
u
m
N
(
100
10
1
0.1
0.01
0.001
0.0001
0.00001
0.000001
0.0000001
1
10
100
1000
10000
Energy (GeV)
FIG. 3: GRAND's response as a function of energy. The dashed line is the number of muons at detection level per incident
proton primary. The highest solid line is for protons incident at 17◦ from zenith, the middle line is for protons at 34◦, and the
bottom line is protons at 51◦.
TABLE II: The results of the Monte Carlo calculation with the FLUKA code for the number of primary protons (N
energy E
different discrete primary proton kinetic energies.
p incident normally at the top of the atmosphere yielding the number of muons (N
p) of kinetic
µ) which reach ground level for
E
p (GeV)
N
p
N
µ
1
3
10
30
100
300
1000
36.0 × 106
27
2.00 × 106
664
0.75 × 106 38511
0.40 × 106 208928
0.10 × 106 256847
0.10 × 106 738561
0.040 × 106 825910
p
µ/N
R = N
7.5 × 10−7
3.32 × 10−4
0.0513
0.522
2.57
7.39
20.6
power law spectrum with differential spectral indices of -- 8 (upper curve) and -- 10 (lower curve) which are normalized
to the point at 9 GeV.
.
A. Asymptotic Direction of Observation versus Rigidity
For rigidities (momentum per charge) in the GV range (kinetic energies in the GeV range), the primaries suffer
deflections in the earth's magnetic field. In order to place GRAND's data for this GLE in the perspective of other
muon stations and neutron monitors, it is of interest to know the direction of a proton primary before entering the
earth's magnetic field (asymptotic direction) in order to arrive at vertical incidence 20 km above Project GRAND.
This asymptotic direction depends on the rigidity of the primary as well as the location of the detector on the surface
of the earth. We are indebted to Margaret Shea and Don Smart for performing this calculation [Shea and Smart,
2002], summarized in Figure 5, which shows the asymptotic direction of an antiproton starting at 20 km above sea level
at the geographical location of GRAND (41.7◦ N and 86.2◦ W) and integrating the trajectories of various rigidities
outward through a model of the earth's magnetic field to find their direction after exiting this field. The coordinate
system for the asymptotic directions can be visualized as the intersection of this antiproton's trajectory on a huge
sphere of the earth's coordinates with a radius much larger than the earth's magnetosphere. This is then the direction
of a proton traveling toward the earth in the geomagnetic field which would reach an altitude of 20 km height above
GRAND at 0◦ relative to its local zenith.
The model of the earth's magnetic field which was used is the International Geomagnetic Reference Field, Epoch
1995 (I95). This is a quiescent field model and does not take into account magnetospheric perturbations (the K
index during the GLE was 4). The individual asymptotic directions change with time because of the magnetospheric
p
6
1000
100
10
1
0.1
0.01
0.001
0.0001
1E-05
1E-06
1E-07
1E-08
1E-09
1E-10
0.01
0.1
10
Primary Proton Kinetic Energy (GeV)
1
100
FIG. 4: The circular points represent integral proton intensity data from the GOES satellite [GOES Website, 2001] in units of
particles cm−2s−1sr−1, while the square points are integral proton intensities from neutron monitor data compiled by Leonty
Miroshnichenko [Miroshnichenko, 2002 preliminary] in the same units with an estimated factor of two error. The dashed line
shows the R-values from Table 2 versus the primary proton kinetic energy calculated with FLUKA (in units of muons per
primary proton). The diamond points from 1.5 to 9 GeV represent the fold of proton intensity with R-values to obtain the
expected muon differential intensity at ground level for each interval (in units of particles cm−2s−1sr−1GeV −1. At energies
greater than 10 GeV, the spectrum has been approximated with differential spectral indices of -- 8 (upper solid curve) and -- 10
(lower solid curve) normalized to the 9 GeV point.
configuration. However, for rigidity values above about 2 to 2.5 GV, the changes are slight. Of greater importance
would be changes related to the secular variation of the geomagnetic field which, for some areas of the world, can greatly
alter the asymptotic directions close to cutoff values. Also during major geomagnetic disturbances the asymptotic
directions can be changed considerably since the cutoff rigidity can be considerably lowered. However, for the Easter
Sunday GLE and for GRAND, the values calculated should be adequate for almost any analysis of muon response
[Shea and Smart, 2002].
GRAND
20 GV
)
s
e
e
r
g
e
d
(
e
d
u
t
i
t
a
L
50
40
30
20
10
0
-10
-20
-30
-40
-50
6 GV
3 GV
5 GV
4GV
3.5 GV
-90
-75
-60
-30
-45
-15
Longitude (degrees)
0
15
30
FIG. 5: Asymptotic directions as a function of particle rigidities for GRAND [Shea and Smart, 2002].
Data on the Interplanetary Magnetic Field (IMF) is shown in figure 6 [GOES website, 2002]. The magnitude of
the IMF is shown as circular points. The components in GSE coordinates are by squares (x, from the earth to the
Sun along the ecliptic), triangles (y, perpendicular to x in the ecliptic plane), and diamonds (z, in the direction of the
north ecliptic pole). While the magnitude of the IMF remains fairly constant, the direction of the magnetic field is
changing over the period of this GLE.
7
)
T
n
(
t
h
g
n
e
r
t
S
d
e
F
l
i
4
3
2
1
0
-1
-2
-3
14
14.1
14.2
14.3
14.4
14.5
14.6
Time (UT)
FIG. 6:
Interplanetary Magnetic Field during the April 15, 2001 GLE [GOES Website, 2002]. The circular points represent
the magnitude of the magnetic field. The squares, triangles, and diamonds represent the x,y, and z components of the magnetic
field in the GSE coordinate system.
VI. OTHER OBSERVATIONS
Data from the Climax, Newark, and Oulu Neutron Monitor Stations are compared for this time interval on April
15 [Lopate, Clem and Pyle, and Oulu Neutron Monitor Website, 2001]. In Figure 1, the data for Climax are very close
to that of Newark and for clarity are not shown. The energy thresholds for the Climax, Newark, and Oulu neutron
monitor stations are influenced by their vertical geomagnetic cutoff rigidity at 3.0, 2.1, and 0.8 GV respectively
[University of Chicago Neutron Monitor Website, 2001] and overburden of air; their corresponding heights above sea
level are 3400, 50, and 15 m. The corresponding numbers for GRAND are a geomagnetic cutoff rigidity of 1.9 GV and
a height of 220 m above sea level. However, the primary energy for GRAND is influenced by the different mechanisms
which produce muons (rather than neutrons) at detection level. While looking at single muon tracks at ground level,
GRAND's primary proton energy is approximately 4 GeV (see Figure 4 and [Shea and Smart, 2002]) for the spectral
shape of this GLE.
When binned in six-minute intervals, the Newark signal has an onset time at 14.21 hours, the Climax signal at
14.16 hours, and Oulu at 14.27 hours (where the "onset" is the time of the half-height of the signal as measured from
its rise above background to the highest measured data point of the GLE). A similar calculation for the onset time
for GRAND is 14.09 hours. The width (full width at half-of-maximum above background) of the Climax, Newark,
and Oulu peaks are 1.27, 1.24, and 1.61 hours in contrast to GRAND's value of 0.48 hour. Thus the GRAND peak
is: a) shorter in duration (typically, higher energies have a shorter duration than lower energies as is seen in satellite
data for protons at lower energies [GOES Website, 2002]), b) the onset is ∼0.1 hour earlier (the earlier time might
be expected on the basis of higher primary energies compared to the neutron monitors [Clem and Dorman, 2000]
with faster protons traveling at smaller pitch angles yielding shorter path lengths, and c) smaller in amplitude due to
GRAND's sensitivity to higher primary energies.
Project GRAND sees a ground level signal with a significance of 6.1 σ when examining the secondary muon counting
rate at ground level between 14.0 and 14.6 hours UT on April 15, 2001. This signal is obtained with no restrictions on
VII. CONCLUSIONS
the angle of the muons. This is consistent with protons originating at the surface of the sun and accelerated during
the time of the X14 X-ray flare.
8
A slightly negative signal ( -- 0.4 σ) is obtained if a small region of the sky (±5◦ in θ
yz) near the angle
of the sun's position is examined. Thus, neither gamma ray nor neutron primaries are the cause of the 6.1σ signal
found in the data with no angular cuts.
xz and in θ
GRAND's GLE detection occurs slightly earlier than Climax, Newark, or Oulu neutron monitor signals which might
be expected from the slightly higher primary energies detected by GRAND (above about 4 GeV) compared to the
neutron monitors (the vertical geomagnetic cutoff energies for the Climax, Newark, and Oulu neutron monitor stations
are 3, 2, and 0.8 GeV respectively). Earlier times would be consistent with smaller pitch angles relative to the IMF
(and thus a shorter flight path) for the higher energies as well as the increased velocity. The fact that GRAND's GLE
signal is not as pronounced as those from the neutron monitors indicates that there are fewer particles at GRAND's
higher energies. The mean energy of primary protons which produces the detected muons depends somewhat upon
the spectral index of this GLE in the 4 GeV region. In the future, combined analyses of world-wide neutron monitor
stations and muon telescopes are anticipated to yield more detailed information on this GLE.
Acknowledgments
The authors wish to thank Monica Dunford, Jon Vermedahl, and Mari´e L´opez del Puerto for their participation
in the data analyses; Kent Doggett (and the ACE science team); the GOES science team; and Stefan Roesler and
Alberto Fasso for their assistance with the Monte Carlo simulations. Thanks to Leonty Miroshnichenko for preliminary
intensity calculations. We also thank Margaret Shea and Don Smart for calculations of GRAND's asymptotic direction
as a function of rigidity; John Clem and Roger Pyle (and Bartol), Cliff Lopate (and Climax), and the Oulu Neutron
Monitor for the use of their data; Ted Bowen for suggestions on improving the text; and Jule Poirier for proofreading.
The Newark neutron monitor is operated by the Bartol Research Institute Neutron Monitor Program and is funded
by National Science Foundation Grant ATM-0000315. The Climax Neutron Monitor is operated by The University
of Chicago and is funded through National Science Foundation Grant ATM-9912341. Project GRAND is funded
through the University of Notre Dame and private grants.
[1] Advanced Composition Explorer (ACE) Homepage, World Wide Web: http://www.srl.caltech.edu/ACE.
[2] J. Bieber et al., Energetic Particle Observations During the 2000 July 14 Solar Event, Astrophysical Journal, 622, March
1, 2002.
[3] M. Duldig, Muon Observations, Space Science Reviews, p. 207-226, January, 2000.
[4] J. Clem and L. Dorman, Neutron Monitor Response Functions (and references therein), Space Science Reviews, 93, 335,
2000.
[5] J. Clem and R. Pyle, Private Communication, 2001. World Wide Web: http://www.bartol.udel.edu/∼neutronm.
[6] A. Falcone, J. Ryan, Milagro as a Solar Observatory, Astroparticle Physics, 283, 1999.
[7] A. Fass`o, A. Ferrari, P.R. Sala, Electron-photon transport in FLUKA: Status, Proceedings of the Monte Carlo 2000
Conference, Lisbon, 159, 2000.
[8] A. Fass`o et al., FLUKA: Status and Prospective for Hadronic Applications, ibid, 955, 2000.
[9] GOES Website. World Wide Web: http://rsd.gsfc.nasa.gov/goes
[10] GRAND's Homepage. World Wide Web: http://www.nd.edu/∼grand
[11] IPS Radio and Space Services Website, 2001. World Wide Web: http://www.ips.gov.au.
[12] C. Lopate, Private Communication, 2001.
[13] L. Miroshnichenko, Preliminary Results from Pivate Communication, 2002.
[14] L. Miroshnichenko, Solar Cosmic Rays, Kluwer Academic Publishers, 2001.
[15] K. Munakata, A Prototype Muon Detector Network Covering a Full Range of Cosmic Ray Pitch An-
the 27th International Cosmic Ray Conference, Hamburg, 3494, 2001. World Wide Web:
gles, Proceedings of
http://www.copernicus.org/icrc/programme.htm.
[16] NOAA Space Environment Center Website, 2001. World Wide Web: http://www.sec.noaa.gov.
[17] Oulu Neutron Monitor Website, 2001. World Wide Web: http://spaceweb.oulu.fi/projects/crs.
[18] J. Poirier and C. D'Andrea, Variation of Muon Counts versus Solar Time, Proceedings of the 27th International Cosmic
Ray Conference, Hamburg, 3934, 2001a. World Wide Web: http://www.copernicus.org/icrc/programme.htm.
[19] J. Poirier and C. D'Andrea, A Measurement of Secondary Muon Angular Distributions, Proceedings of the 27th Interna-
tional Cosmic Ray Conference, Hamburg,
3923, 2001b. World Wide Web: http://www.copernicus.org/icrc/programme.htm.
[20] J. Poirier, S. Roesler, and A. Fass`o, Distributions of Secondary Muons at Sea Level from Cosmic Gamma Rays Below 10
TeV, astro-ph/0103030, 2001.
[21] J. M. Ryan and the MILAGRO Collaboration, High Energy Solar Particles in the 6 November 1997 Ground Level
the 27th International Cosmic Ray Conference, Hamburg, 3367, 2001. World Wide Web:
Event, Proceedings of
http://www.copernicus.org/icrc/programme.htm.
[22] M. Shea and D. Smart. Private Communication, 2002.
[23] D.B. Swinson and M.A. Shea, The September 29, 1989 Ground Level Event Observed at High Rigidity, Geophysical
Research Letters, 8, 1073, 1990.
[24] University of Chicago Neutron Monitor Website, 2001. World Wide Web: http://ulysses.uchicago.edu/NeutronMonitor/neutron mon.html.
[25] E. V. Vashenyuk, The Ground Level Enhancement of 14 July 2000: Explaining the difference between near-by neutron
monitors at Apatity and Oulu, Proceedings of the 27th International Cosmic Ray Conference, Hamburg, 3383, 2001.
9
|
astro-ph/0702693 | 1 | 0702 | 2007-02-26T21:19:36 | The present day mass function in the central region of the Arches cluster | [
"astro-ph"
] | We study the evolution of the mass function in young and dense star clusters by means of direct N-body simulations. Our main aim is to explain the recent observations of the relatively flat mass function observed near the centre of the Arches star cluster. In this region, the power law index of the mass function for stars more massive than about 5-6 solar mass, is larger than the Salpeter value by about unity; whereas further out, and for the lower mass stars, the mass function resembles the Salpeter distribution. We show that the peculiarities in the Arches mass function can be explained satisfactorily without primordial mass segregation. We draw two conclusions from our simulations: 1) The Arches initial mass function is consistent with a Salpeter slope down to ~1 solar mass, 2) The cluster is about half way towards core collapse. The cores of other star clusters with characteristics similar to those of the Arches are expected to show similar flattening in the mass functions for the high mass (>5 solar mass) stars. | astro-ph | astro-ph |
Mon. Not. R. Astron. Soc. 000, 1 -- 6 (2007)
Printed 14 May 2018
(MN LATEX style file v2.2)
The present day mass function in the central region of the
Arches cluster
Simon Portegies Zwart,1,2⋆ Evghenii Gaburov,1,2⋆ Hui-Chen Chen1,2,3⋆ and
M. Atakan Gurkan1⋆
1 Astronomical Institute 'Anton Pannekoek' University of Amsterdam, the Netherlands
2 Section Computational Science, University of Amsterdam, the Netherlands
3 Graduate Institute of Astronomy, National Central University, No 300 Jhongda Rd. Jhongli City, Taiwan
14 May 2018
ABSTRACT
We study the evolution of the mass function in young and dense star clusters by
means of direct N -body simulations. Our main aim is to explain the recent observations
of the relatively flat mass function observed near the centre of the Arches star cluster.
In this region, the power law index of the mass function for stars more massive than
about 5 -- 6 M⊙ is larger than the Salpeter value by about unity; whereas further out,
and for the lower mass stars, the mass function resembles the Salpeter distribution. We
show that the peculiarities in the Arches mass function can be explained satisfactorily
without primordial mass segregation. We draw two conclusions from our simulations:
1) The Arches initial mass function is consistent with a Salpeter slope down to ∼
1 M⊙, 2) The cluster is about half way towards core collapse. The cores of other star
clusters with characteristics similar to those of the Arches are expected to show similar
flattening in the mass functions for the high mass ( >
∼ 5 M⊙) stars.
Key words: methods: N-body simulations -- open clusters and associations: individ-
ual: Arches -- galaxies: star clusters
1
INTRODUCTION
The mass function of a star cluster changes because of
both stellar evolution and stellar dynamics. Stellar evolu-
tion causes the turn-off mass to decrease as the most mas-
sive stars evolve away from the main sequence, ascend the
giant branch to ultimately shed their envelopes to turn into
compact objects. Stellar evolution therefore has a character-
istic effect on the mass function by truncating it at the high
mass end.
The dynamical evolution of a cluster has a more compli-
cated effect on changes in the mass function. The dominant
effect here is dynamical friction, which causes the most mas-
sive stars to sink to the cluster centre on a time scale that is
inversely proportional to the stellar mass, i.e. the most mas-
sive stars tend to sink more quickly than relatively lighter
stars. At the same time, stars less massive than the average
mass tend to leave the inner regions. As a result of this mass
segregation, the local stellar population becomes a function
of the distance to the cluster centre.
⋆ E-mail:
(EG);
(MAG)
[email protected]
[email protected]
(SPZ); [email protected]
[email protected]
(H-CC);
Mass segregation, though mostly noticeable in the clus-
ter's central regions, is a global phenomenon. A star cluster
that is born with the same mass function across its radial
coordinate will gradually grow a top-heavy mass function in
its centre and a top-depleted mass function in its outskirts.
Near the half mass radius, the mass function remains closest
to the initial mass function (Vesperini & Heggie 1997).
In this letter, we concentrate on the evolution of the
stellar mass function in the inner part of young and dense
star clusters, using N -body simulations. Our interest in this
topic was initiated by the recent accurate measurements
published by Stolte et al. (2005); Kim et al. (2006) in which
the mass function in the inner ∼ 10" from the centre of
the Arches star cluster was studied. These observations, es-
pecially the latter, revealed that the mass function of near
the centre of Arches cluster is a broken power law, with the
turning point mp ∼ 5−6M⊙. We were able to reproduce this
feature without invoking any special mechanism. Our sim-
ulations allow us to draw conclusions on the history of the
dynamical evolution of the Arches cluster.
L2
Simon F. Portegies Zwart et al.
2 DYNAMICAL EVOLUTION OF THE MASS
FUNCTION
2.1 Parameters for the simulations
As a cluster evolves, stars more massive than the mean mass
hmi tend to sink to the cluster centre whereas lighter stars
move outwards. For the most massive stars, the time scale
for dynamical friction is proportional to two-body relaxation
time, tr:
tdf ∝
hmi
m⋆
tr,
(1)
were m⋆ is the mass of the massive star, which segregates
inwards. The value of the relaxation time at the cluster's
half-mass radius, rh is given by (Spitzer 1987, eq. 2.63)
tr =
0.138N
h
ln Λ (cid:18) r3
GM(cid:19)1/2
.
(2)
Here G is Newton's constant of gravity, M and N are the to-
tal mass and the number of stars in the cluster and ln Λ is the
Coulomb logarithm, for which we adopt ln Λ = ln(0.01N )
(Giersz & Heggie 1996). For the central relaxation time, we
use
trc =
σ3
3D
4.88π G2 ln Λ nhmci2 ,
(3)
where σ3D, n and hmci are the three-dimensional velocity
dispersion, number density and average stellar mass at the
cluster centre (Spitzer 1987, eq. 3.37).
We follow the dynamical evolution of our mod-
els by means of direct N -body simulations, which
we carry out with the starlab software environment
(Portegies Zwart et al. 2001). The calculations are per-
formed on the GRAPE-6 special purpose
computer
(Makino et al. 1997; Makino 2001).
Our numerical experiments are performed with N =
12288 and 24576 stars. For each N , we perform simulations
starting with a full range of density profiles for which we
chose King (1966) models with the dimensionless parameter
W0 ranging from 3 to 12. The mass function in our simu-
lations is described by a power-law, dN/dm = mx, where
we adopt the Salpeter value for the index (x = −2.35), with
masses ranging from 1 M⊙ to 100 M⊙. To validate our re-
sults, we carried out additional simulations with N = 49152
as well as with a Salpeter mass function with 0.1 M⊙ as
the lower limit. It will turn out that the presence of a tidal
field has little effect on the results, but reducing the lower
limit to the initial mass function to 0.1 M⊙ has a profound
effect on the results, as we discuss below. For clarity we
mainly focus on the models with 12288 and 24576 stars.
With these parameters, the relaxation time at the virial ra-
dius for the 12k models is about 360 N -body time units,
whereas for the 24k models this is 625 N -body time units
(see Heggie & Mathieu (1986)1).
The close proximity of the Arches cluster to the Galac-
tic centre (Cotera et al. 1992, 1996) would seemingly require
the simulations to include tidal effects. And for understand-
ing the dynamics in the cluster outskirts or the evaporation
1 For the definition of an N-body unit, or the summary at
http://en.wikipedia.org/wiki/Natural units.
time scale the tidal field will prove crucial. For studying the
evolution of the central region on the short time scale re-
ported here, however, the tidal field has negligible effect.
We support this statement by carrying out additional sim-
ulations which include the tidal field, and those show no
discernible effect. We therefore focus on the results of sim-
ulations without a tidal field. This has the attractive side
effect that it allows us to scale our results with respect to
N . We also ignore the effects of stellar evolution. This ap-
proximation is rectified as on the short lifetime of the cluster
(2 ± 1 Myr) even the most massive stars remain on the main
sequence, though some effect of the stellar mass loss at the
top end of the mass function can be expected. For example,
a 60 M⊙ zero-age main sequence star with solar metalicity
loses about 3 M⊙ in its first ∼ 2.4 Myr (Lejeune & Schaerer
2001), which has a negligible effect on the slope of the mass
function.
2.2 Dynamical evolution towards core collapse
In our simulations we identify the moment of core collapse
as soon as a persistent binary forms with a binding energy of
at least 100 kT (where the energy scale kT is defined by the
condition that the total stellar kinetic energy of the system,
excluding internal binary motion, is 3
2 N kT). For a cluster
with a mass function that is consistent with the observed
mass function in young star clusters, core collapse occurs at
a more or less constant fraction of the initial central relax-
ation time tcc ∼ 0.2 ± 0.1trc (Portegies Zwart & McMillan
2002; Gurkan et al. 2004). In Fig. 1 we plot the moment of
core collapse as a function of the initial concentration of the
cluster. The slight dependence of tcc/trc on W0, as well as
the offset between our results with those of Gurkan et al.
(2004) is presumably mainly caused by their broader range
of stellar masses (0.2 < m/M⊙ < 120) in the initial mass
function, whereas here we adopt an initial mass function
with 1 < m/M⊙ < 100. An additional effect is expected
from the difference in the number of stars. The simulations
of Gurkan et al. (2004) were carried out with 106. The sys-
tematic difference between our results from simulations with
12k and 24k stars shows that this also affects the results sys-
tematically. At this moment, however, we cannot quantify
this effect.
Dynamical friction causes the massive stars to segre-
gate to the cluster centre making the mass function flat-
ter at the higher end, in this region, until the formation
of a hard binary. In Figure 2 we illustrate the evolution of
the mass function between rcore and 2rcore. We show the
mass function at birth (top curve, the Salpeter mass func-
tion), halfway to core collapse and at core collapse (bottom
curve). We denote the point around which the slope of the
mass function changes by mp, and the power-law indices in
higher and lower ends by xm<mp and xm>mp , respectively.
The apparent decrease in the number of stars in the mass
function presented in Figure 2 is the result of the cluster
becoming more concentrated which causes the adopted an-
nulus (rcore < r < 2rcore) to become narrower.
The effect of flattening of the mass function is less pro-
nounced further away from the cluster centre. This is illus-
trated in Figure 3, where we present the evolution of xm>mp
for the 24k simulations with W0 = 5 and for r = 0 to rcore
(top curve), for r = rcore to 2rcore and for r = 2rcore to 3rcore
The Arches Cluster Mass function
L3
)
c
r
t
/
c
c
t
(
g
o
l
0.6
0.4
0.2
0
-0.2
-0.4
-0.6
-0.8
-1
3
4
5
6
7
W0
8
9
10
11
p
m
>
m
x
-0.4
-0.6
-0.8
-1
-1.2
-1.4
-1.6
-1.8
-2
-2.2
-2.4
-2.6
0
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
1.8
t/tcc
Figure 1. The moment of core collapse (tcc) in units of the core-
relaxation time (trc) as a function of W0. The dashed curve gives
the results of our simulations with 12k stars and the solid curve
with 24k. Since we performed only one simulation per set of initial
conditions no error bars are presented. The bottom (dotted) line
denote the results of Gurkan et al. (2004).
Figure 3. The evolution of xm>mp for various radial bins in the
24k simulation with a W0 = 5 King model. The time is given in
units of core collapse time tcc. The radial bins are 0 < r/rcore < 1
(upper solid curve), 1 < r/rcore < 2 (dotted curve), and 2 <
r/rcore < 3 (lower solid curve). To guide the eye, we plotted
straight lines through the simulation data.
n
g
o
l
3
2.5
2
1.5
1
0.5
0
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
1.8
log m
Figure 2. The mass function of one of the simulations (N = 24k,
W0 = 5) for an annulus 1 < r/rcore < 2, around the cluster
centre. From top to bottom, the curves are at times t = 0, t ≃
0.61tcc, and t ≃ 1.05tcc. The solid lines are least squares fits to
the mass function with a broken power-law.
(bottom curve). The values of mp, xm<mp and xm>mp are
obtained by a three-point least squares fit to the mass func-
tion in a predetermined annulus of the simulated date. Note
that we relaxed the fitting procedure in the sense that the
mass function is not required to be continuous. The point of
stalling of the evolution of the mass function can be iden-
tified by the moment of core collapse, regardless of the ini-
tial concentration or the number of stars in the simulation.
Therefore, we normalize the time axis in Figure 3 to that
instant.
p
m
>
m
x
-1
-1.2
-1.4
-1.6
-1.8
-2
-2.2
-2.4
0
2
4
6
8
10
r/rcore
Figure 4. The value of xm>mp as a function of distance form
the cluster centre after core collapse. The solid curve gives the
average value of xm>mp over the various simulations with 24k
stars for W0 = 3, 4 up to W0 = 12, the dashed line gives the
data for the simulations with N = 12k. The squares, bullets
and triangles with error bars give the observed values taken from
Figer et al. (1999), Stolte et al. (2005) and Kim et al. (2006), re-
spectively (see Tab. 1). The lower thin dashed line gives the value
of xm>mp for simulations with 24k particles with W0 = 9 and a
lower limit to the initial mass function of 0.1 M⊙.
2.3 Post-collapse mass function
After the formation of a hard binary, the mass function
achieves a quasi steady state. The slope of the high-mass
end of the mass function varies throughout the cluster. In
in Figure 4 we show how the mass function for stars with
m > mp after the moment of core collapse is a function of
the distance to the cluster centre, being flatter closer in and
resembling the initial mass function further out. Overplotted
L4
Simon F. Portegies Zwart et al.
n
u
S
M
p
m
/
8
7
6
5
4
3
2
1
0
0
2
4
6
8
10
r/rcore
Figure 5. The value of mp after the point of core collapse, as
a function of distance to the cluster centre for various of simula-
tions. The circles connected with a thin solid line give the results
for the simulations with 24k stars and with W0 = 5 averaged
between the moment of core collapse and twice the core collapse
time. The error bars indicate the variation of the value of mp
over this time period. More concentrated initial models tend to
have a slightly lower value of mp, which we illustrate by plotting
the W0 = 9 simulation as the thick lower dotted line. The upper
dotted line gives the results of the simulation with N = 12k and
for W0 = 5. The thin dashed line at the bottom gives the value of
mp for simulations with 24k particles with W0 = 9 and a lower
limit to the initial mass function of 0.1 M⊙.
are the observed values of the mass function exponent (see
Tab. 1, see § 3). The results of our simulations with a tidal
field are statistically identical to those with a tidal field. The
simulations with a minimum mass to the initial mass func-
tion of mmin = 0.1 M⊙ is plotted as the thin dashed curve in
Fig. 4. For clarity we did not plot error bars for this figure,
but the results with a lower limit of 0.1 M⊙ are inconsistent
with the observed values.
In Figure 5, we show the value of mp as a function of
distance from the cluster centre for various simulations, past
the moment of core collapse. It turns out that more concen-
trated initial models tend to result in a slightly smaller value
of mp whereas simulations with a smaller number of stars
give rise to a higher value of mp. The behaviour of xm>mp is
rather insensitive to the initial concentration of the cluster.
It may be noted that the results of simulations with
0.1 M⊙as the lower limit of the mass function are not con-
sistent with the observed values of xm>mp and mp.
3 MASS FUNCTION OF THE ARCHES
CLUSTER
At a projected distance of about 25 pc from Sgr A*, the
Arches cluster (α = 17h45m50s, δ = −27◦49′28′′ in J2000),
discovered by Cotera et al. (1992, 1996), is peculiar. The
total cluster mass is about 2 · 104 M⊙ . The core radius of
the cluster (defined as the radial distance from the cluster
centre where the luminosity profile drops by a factor two)
is rcore= 5".0 (Stolte et al. 2005), and corresponds to about
0.2 pc if we assume that the distance to the Galactic centre
Table 1. Parameters for the observed mass function of the
Arches cluster. The first two columns give the range over which
the mass function is measured, in units of the cluster's core radius
(rcore ≃ 0.20 pc). The third and fourth columns give the range in
masses for which the exponent of the mass function (last column)
is fitted. Column 5 gives the reference for the mass function expo-
nents 1: Stolte et al. (2005), 2: Kim et al. (2006), 3: Figer et al.
(1999), and the last column the measured value of x between
mmin and mmax.
rmin
(rcore)
0
1
1
0
1
1
0.6
4
3
2.5
2
n
g
o
l
1.5
1
0.5
0
0
rmax mmin mmax
(M⊙)
(M⊙)
(rcore)
ref
x
1
2
2
2
1.8
1.8
1.8
8
12
6
16
6
6.3
1.3
6.3
2.8
60
16
60
60
50
50
125
32
1 −1.26 ± 0.07
1 −1.69 ± 0.08
1 −2.21 ± 0.09
1 −1.86 ± 0.02
2 −1.71 ± 0.15
2 −1.91 ± 0.08
3
3
∼ −1.65
O(−2.35)
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
1.8
log m
Figure 6. The present day mass function of the Arches star
cluster between 1 rcore and 1.8 rcore. The error bars are taken
directly from Fig. 5 of Kim et al. (2006). The three thin lines are
from Fig. 2 at zero age (top line), at t ≃ 0.61tcc (middle line) and
at t ≃ 1.05tcc (bottom line). The thick dashed line is identical to
the thin t ≃ 0.61tcc but then renormalized with -0.35 dex. This
line produces a satisfactory fit to the observed mass function.
is 8 kpc. The age of the cluster is 2 ± 1 Myr (Figer et al.
1999).
Recently, Kim et al. (2006) observed the Arches cluster
using Keck/NIRC2 laser guide star adaptive optics. Their
observations covered the inner parts of the cluster and some
control fields at a distance of about 2.4 pc (60") from the
cluster centre. They subsequently constructed the luminos-
ity and mass functions down to about 1.3 M⊙, in an annulus
of 5" (about 1.0 rcore) to 9" (about 1.8 rcore) from the cluster
centre. In Table 1 we give the various measurements of the
mass function and the distance from the cluster centre in
terms of the observed core radius.
The data show that the slope in the mass function for
stars more massive than ∼ 5 M⊙ flattens towards the cluster
centre (see fig. 4). For lower mass stars, as well as for further
out than ∼ 4 rcore it is closer to to the Salpeter mass func-
tion. The observed mass function in the Arches is not a sim-
ple power-law (Stolte et al. 2005; Kim et al. 2006). We argue
that in the inner parts of the cluster, r <
∼ 4rcore, the mass
function is best described by two power laws with the break
around mp = 5 -- 6 M⊙. The mass function below this break
(mp) resembles the initial mass function (xm<mp ≡ xIMF),
and above mp it becomes flatter (xm>mp > xIMF). Further
out than r ≃ 4 rcore the break disappears and the mass
function becomes gradually better represented with the ini-
tial mass function.
4 DISCUSSION AND CONCLUSIONS
We performed detailed simulations of the evolution of young
and dense star clusters using direct N -body simulations, in
order to constrain the observed mass function within about
one parsec from the centre of the Arches cluster. The initial
conditions of our simulations range over the full spectrum
of King model density profiles.
The mass function in the central region of the Arches
cluster is peculiar as it appears to be split in two power-
laws, one for the stars less massive than 5 -- 6 M⊙ and a much
shallower slope for the more massive stars. The two power-
laws fit the observed data between 5" and 9" is marginally
better (χ2 ≃ 0.5) than a single power-law (χ2 ≃ 0.93).
The simulations we perform to mimic the Arches cluster
are able to reproduce this observed broken power-law mass
function at the observed projected distance from the cluster
centre (r = 0 -- 4 rcore). The best comparison between obser-
vations and simulations is obtained if the cluster is about
half way core collapse (t = 0.4 -- 0.6 tcc).
Our simulations, however, are performed without stel-
lar evolution and without including the effects of an external
tidal potential. As a result, they are scale-free, and no spe-
cific choices for the scalings to mass, size and therefore to
time are obliged. However, the scale-free aspect of our sim-
ulations hinders the direct comparison to some extend as
the size scale (in parsec) and time scale (in Myr) are impor-
tant for an unbiased comparison with the observed Arches
cluster.
In the comparison with the observations we adopt the
same definition of the core radius by projecting the clus-
ter and assigning luminosities of all stars in our simulations
using zero-age main-sequence luminosities2. We ignore here
the fact that very massive stars may become brighter in the
2±1 Myr lifetime of the cluster, but this only affects the most
massive stars, whereas the measurements are dominated by
stars in the mid-range of masses.
The Arches cluster does not show any evidence for pri-
mordial mass segregation as our simulations (which were
initialized without primordial mass segregation) are able to
satisfactorily reproduce the observed mass function over the
2 For comparison, we adopt the observers' definition of core ra-
dius: the point where the surface brightness drops to half its
central value. This is similar to but different from definitions
used in theoretical works of (Spitzer 1987, eq. 1-34) and (Aarseth
2003, eq. 15-4). In starlab we adopt the method as discussed by
Heggie, Trenti, & Hut (2006). The latter definition of the core
radius systematically is about twice the observers' definition.
The Arches Cluster Mass function
L5
entire range of observed masses and distances from the clus-
ter centre. Note also that the presence of primordial gas
which failed to form stars does not seem to have affected
the early cluster evolution, as the observed cluster structure
at an age of 2 ± 1 Myr is satisfactory explained with the
simulations, which do not include gas dynamics. The initial
mass function of the Arches cluster is then consistent with a
Salpeter slope between 1 M⊙ and 100 M⊙ without the need
for a radial dependence. There seems to be no need for a
large population of stars less massive than ∼ 1 M⊙.
In fig. 4 we show the evolution of xm>mp for the an-
nuli and distances from the cluster centre reported from our
compilation from the literature in Tab. 1. The best match
between the simulations and the observations is acquired
for simulations between t = 0.4tcc and 0.6 tcc, i.e: we pre-
dict that the cluster is about half way towards core collapse.
We therefore conclude that the Arches cluster has not yet
experienced core collapse but is currently in a pre collapse
stage.
In Tab. 1 we have quantified the slope to the low mass
end of the mass function in 5" to 9" annulus of the Arches
cluster as consistent with Salpeter, whereas the naive mea-
surement in Fig. 5 of Kim et al. (2006) would results in
xm<mp = −3.67 ± 0.14 (with mp = 5 M⊙), which is unusu-
ally steep. If this slope would represent the intrinsic Arches
initial mass function and we adopt a minimum mass of 1 M⊙
the observed ∼ 2600 stars more massive than ∼ 5 M⊙ in the
Arches cluster would result in a total number of more than
2 × 105 stars, which is unrealistically high. From an obser-
vational point of view there are good arguments that the
low mass end of the mass function is over-estimated, as it
is plagued by selection effects. One of these effects is the
artificial correction of missing stars in a crowded field and
the selection of the three control fields to compensate for
the background population. In the Keck observations these
control fields are within about 2.4 pc from the cluster cen-
tre, which corresponds to ∼ 12 rcore. For a King model with
>
W0
∼ 5.2 the control fields would then be located near the
cluster tidal radius. And since the cluster is about half way
towards core collapse it is conceivable that the density pro-
>
file is described with a King model with W0
∼ 7, in which
case the control field are part of the cluster halo.
Due to mass segregation the cluster outskirts will be
depleted of high mass stars and low mass stars will be over-
represented (the opposite effect as we discussed for the core
population). Correcting the mass function in the cluster core
with a population taken from near the cluster halo will there-
fore result in an enormous over correction towards the low
mass stars, and consequentially result in a steepening of the
'corrected' mass function.
One of the control fields (field B of Kim et al, 2006) is
taken near the location where one expects the tidal tail of
the cluster in the potential of the Galaxy to pass though.
The tidal tail is, since it consists of the halo population, also
likely to be dominated by low mass stars.
Each of the effects discussed tend to steepen the lower-
mass end of the mass function, though it is not trivial to
quantize the effect without a much more detailed study. We
however, argue that the initial mass function of the Arches
cluster was probably consistent with Salpeter over the ob-
served mass range. The observed break in the mass function
around 5 -- 6 M⊙ and the consequential flattening of the mass
Kim S. S., Figer D. F., Kudritzki R. P., Najarro F., 2006,
ApJ, 653, L113
King, I. R. 1966, AJ, 71, 64
Lejeune T., Schaerer D., 2001, A&A, 366, 538
Makino, J. 2001,
in S. Deiters, B. Fuchs, A. Just, R.
Spurzem, R. Wielen (eds.), ASP Conf. Ser. 228: Dynamics
of Star Clusters and the Milky Way, p. 87
Makino, J., Taiji, M., Ebisuzaki, T., Sugimoto, D. 1997,
ApJ, 480, 432
Meaburn, J., Hebden, J. C., Morgan, B. L., Vine, H. 1982,
MNRAS, 200, 1P
Moffat, A. F. J., Poitras, V., Marchenko, S. V., Shara,
M. M., Zurek, D. R., Bergeron, E., Antokhina, E. A. 2004,
AJ, 128, 2854
Piatti, A. E., Bica, E., Claria, J. J. 1998, A&AS, 127, 423
Portegies Zwart, S. F., McMillan, S. L. W. 2002, ApJ, 576,
899
Portegies Zwart, S. F., McMillan, S. L. W., Hut, P.,
Makino, J. 2001, MNRAS, 321, 199
Stolte, A., Brandner, W., Grebel, E. K., Lenzen, R., La-
grange, A.-M. 2005, ApJ, 628, L113
Spitzer, L. 1987, Dynamical evolution of globular clusters
(Princeton, NJ, Princeton University Press)
Vesperini, E., Heggie, D. C. 1997, MNRAS, 289, 898
L6
Simon F. Portegies Zwart et al.
function for higher masses is then the result of the dynamical
evolution of the cluster.
The initial model which is most comparable to the ob-
served Arches cluster has a Salpeter initial mass function
between 1 M⊙ and 100 M⊙ and with a reasonably concen-
trated initial density profile (W0
>
∼ 4 and W0
<
∼ 8).
The break in the mass function in the inner parts of
the cluster (for r <
∼ 4rcore) appear at mp ≃ 2hmi, which for
our simulations is at about 5 M⊙. The break in the observed
mass function in the Arches cluster appears around the same
mass of mp ≃ 5 -- 6 M⊙. We performed additional simulations
with W0 = 9 using a Salpeter mass function down to 0.1 M⊙,
and in this case the break in the core mass function also
developed around mp ≃ 2hmi, which for the adopted mass
function is about 1.0 M⊙(Figure 5). Based on these findings
we argue that the initial mass function in the Arches cluster
has a lower limit of about 1.0 M⊙, as in our simulations that
reproduce the observations best.
We predict that other clusters with similar parameters
as the Arches cluster, like Westerlund 1 (Piatti et al. 1998),
NGC 3603 (Moffat et al. 2004), R 136 (Meaburn et al. 1982)
and Quintuplet (Cotera et al. 1992) will show similar char-
acteristics as Arches. The mass functions in their cores will
also be rather flat for stars more massive than 5 -- 6 M⊙. And
the mass functions further away from the cluster centre will
gradually be more like the initial mass function.
ACKNOWLEGMENTS
We are grateful to Peter Anders, Mark Gieles, Alessia
Gualandris, Douglas Heggie and Henny Lamers for many
discussions. This work was supported by NWO (grants
#635.000.303 and #643.200.503), NOVA, the LKBF, the
ISSI in Bern, Switzerland and the Taiwanese government
(grants NSC095-2917-I-008-006 and NSC95-2112-M-008-
006). MAG is supported by a Marie Curie Intra-European
Fellowship under the sixth framework programme. The cal-
culations for this work were done on the MoDeStA computer
in Amsterdam, which is hosted by the SARA supercomputer
centre.
REFERENCES
Aarseth S. J., 2003, Gravitational N-body simulations,
Cambridge University press
Cotera, A. S., Erickson, E. F., Colgan, S. W. J., Simpson,
J. P., Allen, D. A., Burton, M. G. 1996, ApJ, 461, 750
Cotera, A. S., Erickson, E. F., Simpson, J. P., Colgan,
S. W. J., Allen, D. A., Burton, M. G. 1992, in Bulletin of
the American Astronomical Society, p. 1262
Figer, D. F., Kim, S. S., Morris, M., Serabyn, E., Rich,
R. M., McLean, I. S. 1999, ApJ, 525, 750
Giersz, M., Heggie, D. C. 1996, MNRAS, 279, 1037
Gurkan, M. A., Freitag, M., Rasio, F. A. 2004, ApJ, 604,
632
Heggie D. C., Trenti M., Hut P., 2006, MNRAS, 368, 677
Heggie, D. C., Mathieu, R. D. 1986, LNP Vol. 267: The
Use of Supercomputers in Stellar Dynamics, in P. Hut, S.
McMillan (eds.), Lecture Not. Phys 267, Springer-Verlag,
Berlin
|
astro-ph/0611889 | 1 | 0611 | 2006-11-29T13:13:22 | Origin of Superluminal radio jets in GRS 1915+105 and the role of the Plateau state | [
"astro-ph"
] | We have studied the accretion disk during the radio plateau state and the following superluminal relativistic radio jets and have provided a tight correlation between accretion disk and superluminal jet parameters. We find that accretion rate during the plateaux is very high and suggest that the accretion disk during the radio plateaux is always associated with radiation-driven wind. The internal shock forms in the previously generated slowly moving wind (during plateau) with $\beta$ $\le$ 0.01 as the fast moving discrete jet (usually at the end of plateau) with $\beta$ $\sim$ 1 catches up and interacts with it. The power of superluminal jet is determined by the strength and speed of these two components; the slow moving wind and the fast moving jet which are related to the accretion disk during the plateau state. Finally, we discuss the implication of this work. | astro-ph | astro-ph | Origin of Superluminal radio jets in GRS 1915+105 and
the role of the Plateau state
Tata Institute of Fundamental Research, Homi Bhabha Road,
Mumbai-400005, India
J S Yadav
Abstract
We have studied the accretion disk during the radio plateau state and
the following superluminal relativistic radio jets and have provided a tight
correlation between accretion disk and superluminal jet parameters. We find
that accretion rate during the plateaux is very high and suggest that the
accretion disk during the radio plateaux is always associated with radiation-
driven wind. The internal shock forms in the previously generated slowly
moving wind (during plateau) with β ≤ 0.01 as the fast moving discrete jet
(usually at the end of plateau) with β ∼ 1 catches up and interacts with it.
The power of superluminal jet is determined by the strength and speed of
these two components; the slow moving wind and the fast moving jet which
are related to the accretion disk during the plateau state. Finally, we discuss
the implication of this work.
6
0
0
2
v
o
N
9
2
1
v
9
8
8
1
1
6
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
VI Microquasar Workshop: Microquasars and Beyond
September 18-22, 2006
Como, Italy
1
1
Introduction
In active galactic nuclei (AGN) & quasars, large superluminal outflows/jets are known to
exist for almost half century yet remain poorly understood. This is probably due to the
lower characteristic accretion disk temperature as well as the fact that the central part of
these systems is usually obscured by a large amount of dust. The long time scales associated
with these massive systems pause an additional observational problem. Discovery of the first
microquasar GRS 1915+105 in 1994 in our Galaxy revived great hope to study the disk-jet
connection in general and the origin of superluminal jets in particular [1]. Microquasars are
closer, smaller and show faster variability that is easily observable. GRS 1915+105 has shown
exceptionally high variability in X-rays as well as in radio. Since 1996, rich X-ray variability of
this source is observed by RXTE [2, 3] and by the Indian X-ray Astronomy Experiment, IXAE
[4]. Belloni et al. (2000) have classified the complex X-ray variability of GRS 1915+105 in 12
separate classes on the basis of their light curves and color-color diagrams and suggested three
basic states of this source, namely hard state and two softer states with different temperatures
of the accretion disk.
The radio emission from GRS 1915+105 can be broadly put into two classes; (1) radio
emission close to the compact object (¡200 AU), and (2) radio emission at large distances (≥
240 AU). The former class includes (a) steady radio jets (radio plateau state), (b) preplateau
flares, and (c) oscillations/baby jets (discrete jets) of 20-40 min duration in infrared (IR) &
radio while the latter class includes large superluminal radio jets. The steady radio jets of 20
-- 160 mJy flux density are associated with the canonical low hard X-ray state and observed
for extended durations [6, 7]. These are optically thick compact jets with velocity β of 0.1 -- 0.4
[8, 9]. The radio emission is correlated with the X-ray emission as Lradio ∝ L0.7
X for several
different sources [10]. Pooley & Fender (1997) observed radio oscillations with delayed emission
at lower frequency. Simultaneous X-ray, IR and radio multi-wavelength observations provided
first major step in our understanding of disk-jet interaction and suggested that the spike in
X-ray coincides with the beginning of IR flare and it has self absorbed synchrotron emission
associated with adiabatic expansion [12, 13]. These are also compact jets with velocity β ∼ 1
[8]. The relativistic superluminal jets with up to 1 Jy flux density have steep radio spectrum
and are observed at large distances few hundred AU to 5000 AU from the core [1, 14, 8]. These
radio jets are very energetic with luminosity close to the Eddington luminosity, LEdd and have
been observed in several sources [14, 15, 16, 17]. Progress in our understanding of these jets,
especially of their connection to the accretion disk has been slow. The physical connection
between X-ray emission and the superluminal flares has been the hardest to understand.
In this paper, we investigate the association of large superluminal jets with the radio
plateaux. We have analysed the available RXTE PCA/HEXTE X-ray data during radio
plateaux and the radio flare data from the Green Bank Interferometer (GBI). Yadav (2006)
has described the analysis procedure and the selection criteria in detail.
2
Superluminal Radio Flare Properties
Associated Preceding Plateau Properties
MJD/
Date
Peak
Flux
(mJy)
Decay Time
constant
(days)
50750/1997 Oct 30
50916/1998 Apr 13
50933/1998 Apr 30
50967/1998 Jun 03
51204/1999 Jan 30
51337/1999 Jun 08
51535/1999 Dec 23
52105/2001 Jul 16b
a: Integrated 3 -- 150 keV X-ray flux in 10−8 ergs cm−2 s−1
b: VLBA radio data (see text for details)
3.16+0.07
−0.09
4.02+0.36
−0.33
3.98+0.13
−0.13
2.82+0.48
−0.28
1.12+0.03
−0.05
2.67+0.22
−0.16
2.67+0.12
−0.14
1.77+0.20
−0.20
550
920
580
710
340
490
510
210
Rise
Time
(day)
<0.6
<0.8
<0.9
0.3
0.25
<0.7
<0.8
<0.7
Date of
GBI
Flux
RXTE
(mJy) Obser.
47.3
91.0
91.0
56.2
27.8
45.5
51.3
20.0
1997 Oct. 25
1998 Apr. 11
1998 Apr. 28
1998 May 31
1999 Jan. 24
1999 Jun. 03
1999 Dec. 21
2001 Jul. 11
QPO Total
Freq. X-ray
Fluxa
(Hz)
2.04
1.88
2.42
1.60
2.34
1.41
1.76
2.14
1.84
2.55
1.98
1.77
2.23
2.12
2.40
1.75
NH
(1022 cm−2)
13.27+0.66
−0.91
14.95+0.82
−0.65
13.55+0.78
−0.31
12.99+0.81
−0.32
11.19+0.92
−0.95
12.43+0.95
−1.15
13.61+1.17
−0.94
10.60+0.77
−0.40
Table 1: All selected superluminal radio flares and their properties (see text for details). X-ray
Properties from RXTE PCA/HEXTE data during preceding radio plateau.
2 Observations and analysis
The typical sequence of events for a superluminal radio flare is shown in Figure 1 for the 550
mJy radio flare on 1997 October 30 (gap in GBI data at peak) and for the 340 mJy radio
flare on 1999 January 30. The start of the radio flares is offset to zero. A superluminal event
starts with small preplateau flares followed by a steady long plateau, followed by superluminal
radio flares. The preplateau flares are discrete ejections of adiabatically expanding synchrotron
clouds with flat radio emission and are similar to the oscillations/discrete jets. The exponential
decay is an important characteristic of superluminal radio flares which differentiates them from
the other class of radio flares which occur close to the compact object [20]. It is also found
that the radio plateau is always associated with a superluminal radio flare [14, 21, 22]. We
searched 2.25 GHz GBI radio monitoring data during the period from 1996 December to 2000
April and selected radio plateaux and the following radio flares which decay exponentially. All
the selected superluminal flares are given in Table 1 along with associated X-ray properties.
One more radio flare on 2001 July 16 which was observed by the Very Large Baseline Array
(VLBA) and Ratan radio telescope is also added in Table 1. VLBA observations clearly
showed an ejecta well separated from the core [23].
We study X-ray properties during radio plateaux within the preceding week from the start
of superluminal flares to avoid changes in accretion disk over the long durations of plateaux
(from -7 to -2 in Figure 1). We also avoid the last day of plateaux as radio data suggest rapid
changes in the accretion disk. For the timing analysis, we used single-bit-mode RXTE/PCA
data in the energy ranges 3.6 -- 5.7 and 5.7 -- 14.8 keV. Normalised power density spectra
of 256 bin are generated and co-added for every 16 s. During the plateaux, it is in a very
high luminosity state (VHS) with power law index Γ > 2 . The spectrum is dominated by
the Compton scattered emission (≥ 85%) rather than the disk. A model consisting of disk +
power law + a comptonised component (CompTT) with a gaussian line at 6.4 keV is used for
the X-ray analysis. The integrated radio flux is calculated by integrating the fitted exponential
function over a duration three times the decay time constant. Further details of X-ray and
3
Figure 1: GBI 2.25 GHz radio data for the 1999 January 30 (start at MJD 51204.7) and 1997
October 30 (start at MJD 50750.6) large superluminal radio flares. Start of the flares is offset
to 0. The dotted line connects the data points of the 1999 January 30 superluminal radio flare
to guide the eyes.
radio data analysis are given in Yadav (2006) [20].
3 Results and discussion
It is increasingly believed that the coronal material (and not the disk material) is ejected
prior to radio flares [24, 22, 18, 25]. It is consistent with the observation that radio flares
are observed during the transition from the low hard state to the high soft state but never
observed during the opposite transition [21]. It also agrees with the suggestion that the spike
in X-ray during the β- class is associated with the change of the X-ray state due to a major
ejection episode [26]. Once this coronal material is ejected, it invariably decouples from the
disk. However, this decoupling may not produce any observable effect during the plateaux
as inflow and outflow in the accretion disk are in equilibrium. Both the accretion disk and
the radio flare are in steady state. On the other hand, decoupling is supposed to produce
maximum effect during the superluminal radio flares which are observed at large distances of
a few hundred to few thousand AU from the core [6, 14, 8]. This has made disk-jet connection
the hardest to understand for superluminal radio jets [18, 19].
In Figure 2, we plot the total 3 -- 150 keV X-ray flux during the preceding plateau vs the
integrated radio flux of the superluminal flare. We derive a correlation coefficient of 0.99 sug-
4
Figure 2: Correlation between the total 3-150 keV X-ray flux from spectral analysis of RXTE
PCA/HEXTE X-ray data during preflare plateau state and the integrated flux of the super-
luminal radio flare. The solid line is the linear fit to the data points.
gesting a strong connection between the total X-ray flux during the preceding plateau state
and the integrated radio flux of the superluminal flare. The total X-ray flux is dominated
by the Compton scattered emission or coronal emission (≥ 85%). Using the total Compton
scattered emission flux instead of the total X-ray flux improves this correlation with a correla-
tion coefficient of 0.996. In Figure 3, we plot the QPO frequency from our timing analysis of
X-ray data during the plateau state as a function of the decay time constant of the following
superluminal radio flare. This also shows a tight correlation, with correlation coefficient of
0.98. The QPOs are believed to be associated with the coronal flow. The remarkable feature
of our findings here is that the parameters calculated using completely independent spectral
and timing analysis bring out a clear connection between the accretion disk during the plateau
state and the following superluminal radio flare.
In the last column of Table 1, we give the calculated absorption column density NH . The
NH ranges from 10×1022 to 15×1022 cm−2 which are higher than the commonly used NH ∼
5 -- 6×1022 cm−2 for spectral analysis of GRS 1915+105 . The calculated NH shows a tight
correlation with the CompTT flux and not with the disk blackbody flux. This rules out the
possibility that the enhanced NH may be due to overestimation of the disk normalisation.
Yadav (2006) has discussed this in detail and compared these with results obtained from
Chandra and ASCA data [27, 28]. Lee et al. (2002) have analysed Chandra & RXTE X-
ray data during a radio plateau state and have suggested presence of disk wind. Kotani et
al. (2000) also came to a similar conclusion using ASCA data. The wind terminal velocity is
estimated to be of the order of 108 cm s−1 (β ≤ 0.01). The lower limit of bolometric luminosity
Lbol ∼ LX = 6.4×1038 ergs s−1 which is 0.35 of Eddington Luminosity LEdd, for a black hole
5
Figure 3: Correlation between the quasi-periodic oscillation (QPO) frequency from timing
analysis of RXTE/PCA X-ray data during preflare plateau state and the decay time constant
of the superluminal radio flare. The dotted line is the linear fit to the data.
of mass 14 M⊙ [27]. The corresponding lower limit of bolometric luminosity for the X-ray flux
listed in Table 1 falls in the range 0.35 -- 0.48 of LEdd. When GRS 1915+105 is accreting near
LEdd, the presence of a radiation-driven wind is always expected and the wind density should
be a strong function of the disk luminosity. Our derived values of NH show strong dependence
on observed total X-ray flux with correlation coefficient of 0.995 [30].
The calculated power density spectra shown in Figure 4 also independently lend support
to the presence of wind during the plateau state. Figure 4 shows PDS spectra observed on
2001 July 11 and 1998 April 28. The Compton scattered flux increases from 1.5×10−8 ergs
cm−2 s−1 on 2001 July 11 to 1.9×10−8 ergs cm−2 s−1 on 1998 April 28 while the observed
QPO frequency decreases from 2.4 Hz on 2001 July 11 to 1.4 Hz on 1998 April 28. It is clear
from Figure 4 that the power at frequency ν > 0.2 Hz is less in the PDS observed on 1998
April 28 than that observed on 2001 July 11. The fast variability is suppressed by photon
scattering in the enhanced wind on 1998 April 28, hence reducing the power in the PDS at
higher frequencies. Shaposhnikov & Titarchuk (2006) have discussed the decrease in the PDS
power observed in Cyg X-1 at higher frequency (ν > 0.1 Hz) as the wind increases. In Cyg X-
3, the PDS power in the low hard state drops to below 10−3 (rms/mean)2 Hz−1 at frequencies
ν > 0.1 Hz as a dense wind from the companion always envelops the compact object [30].
These results support the internal shock model for the origin of superluminal flares [31].
The internal shock should form in the previously generated slowly moving wind from the
accretion disk with β ≤ 0.01 as the fast moving discrete jet with β ∼ 1 [8] catches up and
interacts with it. Both the components, slow moving wind and fast moving jet, are related
to the accretion disk during plateau state and the strength & speed of these two components
6
Figure 4: Normalised power density spectra in 0.09 -- 9 Hz frequency range observed on 1998
April 28 and 2001 July 11. Strong QPOs are seen at 1.4 Hz and 2.4 Hz in PDS observed on
1998 April 28 and 2001 July 11 respectively. The total Compton scattered flux is also given.
will determine the power of the internal shock. The wind deposits a large amount of energy
as mwind approaches maccr prior to the switch-on of superluminal flare [20]. Thus, the internal
shock model can easily accommodate high jet power requirement ≥ 1038 ergs s−1 [14] and can
explain the shifting from thick to thin radio emission during superluminal flares [31]. Our
results in Figure 5 which shows the peak flux of superluminal radio flares as a function of NH
strongly support our description of superluminal radio flares. A fit to the data (dotted line)
suggests that for wind strength corresponding to NH ≤ 8.3±1.5×1022 cm−2, no superluminal
jet will be produced. The absence of superluminal jets during the class β in GRS 1915+105 is
attributed to the absence of wind [20]. The calculated NH during class β is below this critical
NH value [30]. This model of superluminal jets can provide simple explanations for (1) why
the direction of superluminal jets may differ from the direction of the compact jets, (2) why
the phase lag is complex during plateaux, and (3) why compact (discrete) jets are absent in
Cyg X-3 [20].
In Figure 6, we show a sketch of the heliosphere due to disk wind and mark the locations
of various types of radio emission seen in GRS 1915+105 . All the radio flares observed in
GRS 1915+105 can be broadly put into two groups on the basis of their flux, radio spectrum
and spatial distribution; (1) the superluminal flares (200 -- 1000 mJy) which have steep radio
spectra and are seen at large distances (≥ 240 AU), and (2) all other flares (5 -- 360 mJy)
which include the preplateau flares, radio oscillations & discrete flares and the steady radio
emission during the plateaux. All these flares have flat radio spectra and are observed close to
the compact object. All these radio flares are consistent with the ejection of an adiabatically
7
Figure 5: Correlation between the absorption column density, NH and the peak flux of the
superluminal radio flares. The peak flux of a superluminal flare is calculated using exponential
profile fitting if there is a gap in radio data. The dotted line is a linear fit to the data.
expanding self absorbing synchrotron cloud from the accretion disk and this cloud is supposed
to consist of coronal mass. During the radio plateau, it may be confined expansion due to the
presence of a dense wind which agrees with the source size [8]. The IR & radio oscillations
(5 -- 150) are periodic ejections of adiabatically expanding self absorbing synchrotron clouds
[13, 32]. The preplateau radio flares (50 -- 360 mJy) are discrete ejections which are closely
spaced in time and hence produce overlapped radio flares. The preplateau flares have been
modeled as adiabatically expanding self absorbing clouds ejected from the accretion disk (like
in the case of oscillations & discrete jets) which explain reasonably well all the available
observed data of time delay for radio emission at lower frequency [32]. All these types of radio
emission in GRS 1915+105 starts somewhere close to the corona.
As discussed above, the maccr, as inferred from the X-ray flux, is very high during the radio
plateaux and Lbol approaches LEdd. It is suggested that such hot accretion disk during the
radio plateaux always accompanies with a radiation-driven wind. The NH is tightly correlated
with the COMPTT flux. This is analogous to the solar wind originating from the dense solar
corona. Since the disk wind power & speed are higher than those of solar wind (the average
solar wind speed ∼ 450 km/s), the size of heliosphere due to the disk wind may be around
200-300 AU (the solar heliosphere size is of 100-150 AU and it is supposed to vary with the
solar activity). It is expected that the size of disk heliosphere will increase as the wind power
increases with disk luminosity as discussed earlier. As shown in Figure 6, superluminal radio
8
Figure 6: Sketch diagram (not to the scale) of the heliosphere created by disk wind similar
to the solar heliosphere. The locations are marked where different types of radio jets are
produced. Superluminal jets start at the boundary of the disk heliosphere while all other
types of the jets start somewhere close to the corona.
jets are suggested to appear at the boundary of disk heliosphere. It is expected that strong
superluminal jets should appear at larger distances than the distance where weak superluminal
jets appear. A superluminal jet with peak flux of ∼ 200 mJy was reported around 240 AU on
2001 July 16[23] while strong superluminal jets are suggested to appear around 500 AU (the
peaks of these superluminal flares are not observed but the data are consistent with the above
suggestion) [8]. This can be easily tested in future with VLBA and other high resolution radio
data if we catch the start of superluminal flares in GRS 1915+105 as well as in other LMXBs.
4 Conclusions
We have provided a tight correlation between accretion disk and superluminal jet parameters.
We find that maccr during the plateaux is very high and suggest that the accretion disk during
the radio plateaux is always associated with a radiation-driven wind. The internal shock forms
in the previously generated slowly moving wind (during plateau) with β ≤ 0.01 as the fast
moving discrete jet (usually at the end of plateau) with β ∼ 1 catches up and interacts with
9
it. The strength & speed of these two components determine the power of the internal shock
as well as of the superluminal jets. The peak flux of the superluminal jets is tightly correlated
with the wind power. We discuss the implications of this work and some of these can be
checked in future.
Acknowledgment: The author thanks the RXTE PCA/HEXTE and NSF-NRAO-NASA
Green Bank Interferometer teams for making their data publicly available. The Green Bank
Interferometer is a facility of the National Science Foundation operated by the NRAO in
support of NASA High Energy Astrophysics programs.
References
[1] Mirabel, I. F., & Rodriguez, L. F. 1994, Nature, 371, 46
[2] Morgan, E. H., Remillard, R. A. & Greiner, J. 1997, ApJ, 482, 1010
[3] Muno, M. P., Morgan, E. H. & Remillard, R. A. 1999, ApJ, 527, 321
[4] Yadav, J. S., Rao, A. R., Agrawal, P. C, et al. 1999, ApJ, 517, 935
[5] Belloni, T., Klein-Wolt, M., Mendez, M., et al. 2000, A&A, 355, 271
[6] Muno, M. P., Remillard, R. A., Morgan, E. H., et al. 2001, ApJ, 556, 515
[7] Fuchs, Y., Rodriguez, J., Mirabel, I. F. et al. 2003, A&A, 409, L35
[8] Dhawan, V., Mirabel, I. F. & Rodriguez, L. F. 2000, ApJ, 543, 373
[9] Ribo, M., Dhawan, V. & Mirabel, I. F. 2004, Proc. 7th European VLBI network sympo-
sium, p. 111
[10] Gallo, E., Fender, R. P., & Pooley, G. G. 2003, MNRAS, 344, 60
[11] Pooley, G. G. & Fender, R. P. 1997, MNRAS, 292, 925
[12] Eikenberry, S. S., Matthews, K., Morgan, E. H., et al. 1998, ApJ, 494, L61
[13] Mirabel, I. F., Dhawan, V., Chaty, S. et al. 1998, A&A, 330, L9
[14] Fender, R. P., Garrington, S. T., McKay, D. J. et al. 1999, MNRAS, 304, 865
[15] Hjellming, R. M., & Rupen, M. P. 1995, Nature, 375, 464
[16] Wu, K., Soria, R., Campbell-Wilson, D., et al. 2002, ApJ, 565, 1161
[17] Orosz, J. A., Kuulkers, E., van der Klis, M. 2001, ApJ, 555, 489
[18] Fender, R. P., Belloni, T. M. & Gallo, E. 2004, MNRAS, 355, 1105
10
[19] Fender, R. P. & Belloni, T. M. 2004, ARA&A, 42, 317
[20] Yadav, J. S., 2006, ApJ, 646, 385
[21] Klein-Wolt, M., Fender, R. P., Pooley, G. G., et al. 2002, MNRAS, 331, 745
[22] Vadawale, S. V., Rao, A. R., Naik, S., Yadav, J. S., et al. 2003, ApJ, 597, 1023
[23] Dhawan, V., Muno, M. P., Remillard, R., et al. 2003 preprint.
[24] Rau, A., & Greiner, J. 2003, A&A, 397, 711
[25] Rothstein, D. M., Eikenberry, S. S. & Matthews, K. 2005, ApJ, 626, 991
[26] Yadav, J. S. 2001, ApJ, 548, 876
[27] Lee, J. C., Reynolds, C. S., Remillard, R. , et al. 2002, ApJ, 567, 1102
[28] Kotani, T., Ebisawa, K., Dotani, T., et al. 2000, ApJ, 539, 413
[29] Shaposhnikov, N., & Titarchuk, L. 2006, ApJ, 643, 1098
[30] Yadav, J. S., 2006, "Relativistic superluminal radio jets in microquasars in our galaxy"
this conferece.
[31] Kaiser, C. R., Sunyaev, R. & Spruit, H. C. 2000, A&A, 356, 975
[32] Ishwara-Chandra, C. H., Yadav, J. S., & Pramesh Rao, A. 2002, A&A, 388, L33
11
|
0805.1452 | 1 | 0805 | 2008-05-10T06:40:49 | The Molecular Ridge Close to 30 Doradus in the Large Magellanic Cloud | [
"astro-ph"
] | With the ATNF Mopra telescope we are performing a survey in the 12CO(1-0) line to map the molecular gas in the Large Magellanic Cloud (LMC). For some regions we also obtained interferometric maps of the high density gas tracers HCO+ and HCN with the Australia Telescope Compact Array (ATCA). Here we discuss the properties of the elongated molecular complex that stretches about 2 kpc southward from 30 Doradus. Our data suggests that the complex, which we refer to as the ``molecular ridge,'' is not a coherent feature but consists of many smaller clumps that share the same formation history. Likely molecular cloud formation triggers are shocks and shearing forces that are present in the surrounding south-eastern HI overdensity region, a region influenced by strong ram pressure and tidal forces. The molecular ridge is at the western edge of the the overdensity region where a bifurcated velocity structure transitions into a single disk velocity component. We find that the 12CO(1-0) and HI emission peaks in the molecular ridge are typically near each other but never coincide. A likely explanation is the conversion of warmer, low-opacity HI to colder, high-opacity HI from which H2 subsequently forms. On smaller scales, we find that very dense molecular gas, as traced by interferometric HCO+ and HCN maps, is associated with star formation along shocked filaments and with rims of expanding shell-like structures, both created by feedback from massive stars. | astro-ph | astro-ph |
The Molecular Ridge Close to 30 Doradus in the Large Magellanic
Cloud
Jurgen OttA,B,L, Tony WongC, Jorge L. PinedaD, Annie HughesE,F, Erik
MullerF,M, Zhi-Yun LiG, Min WangH, Lister Staveley-SmithI, Yasuo FukuiJ,
Axel WeissK, Christian HenkelK, & Ulrich KleinD
A National Radio Astronomy Observatory, 520 Edgemont Road, Charlottesville, VA, 22903, USA
B California Institute of Technology, 1200 E. California Blvd., Caltech Astronomy 104 -- 25, Pasadena,
CA, 91125 -- 2400, USA
C Department of Astronomy, University of Illinois, 1002 W. Green St., Urbana, IL 61801, USA
D Argelander Institut fur Astronomie, Universitat Bonn, Auf dem Hugel 71, 53121 Bonn, Germany
E Centre for Supercomputing and Astrophysics, Swinburne University of Technology, Hawthorn VIC
3122, Australia
F CSIRO Australia Telescope National Facility, Cnr Vimiera & Pembroke Roads, Marsfield, NSW
2122, Australia
G Department of Astronomy, University of Virginia, PO Box 400325, Charlottesville, VA 22903-4325,
USA
H Purple Mountain Observatories, CAS, 2 West Beijing Road, Nanjing 210008, China
I School of Physics M013, University of Western Australia, Crawley WA 6009, Australia
J Department of Astrophysics, Nagoya University, Furocho, Chikusaku, Nagoya 464-8602, Japan
K Max-Planck-Institut fur Radioastronomie, Auf dem Hugel 69, 53121 Bonn, Germany
L Jurgen Ott is a Jansky Fellow of the National Radio Astronomy Observatory. E-mail: [email protected]
M Bolton Fellow
Abstract:
With the ATNF Mopra telescope we are performing a survey in the 12CO(1-0) line to map the molecular
gas in the Large Magellanic Cloud (LMC). For some regions we also obtained interferometric maps of
the high density gas tracers HCO+ and HCN with the Australia Telescope Compact Array (ATCA).
Here we discuss the properties of the elongated molecular complex that stretches about 2 kpc southward
from 30 Doradus. Our data suggests that the complex, which we refer to as the "molecular ridge," is not
a coherent feature but consists of many smaller clumps that share the same formation history. Likely
molecular cloud formation triggers are shocks and shearing forces that are present in the surrounding
south-eastern H i overdensity region, a region influenced by strong ram pressure and tidal forces. The
molecular ridge is at the western edge of the the overdensity region where a bifurcated velocity structure
transitions into a single disk velocity component. We find that the 12CO(1-0) and H i emission peaks
in the molecular ridge are typically near each other but never coincide. A likely explanation is the
conversion of warmer, low-opacity H i to colder, high-opacity H i from which H2 subsequently forms.
On smaller scales, we find that very dense molecular gas, as traced by interferometric HCO+ and HCN
maps, is associated with star formation along shocked filaments and with rims of expanding shell-like
structures, both created by feedback from massive stars.
Keywords: ISM: evolution -- ISM: molecules -- galaxies: ISM -- galaxies: individual(Large Magel-
lanic Cloud) -- radio lines: ISM -- (galaxies:) Magellanic Clouds
1
Introduction
The study of star formation (SF) in molecular clouds
is an area which has seen rapid progress in recent
years and the Magellanic Clouds are increasingly seen
as one of the most important laboratories for this re-
search. At a distance of about 50 -- 60 kpc and an in-
clination angle of ∼ 35◦ (van der Marel & Cioni 2001)
for the Large Magellanic Cloud (LMC), current and
future instruments can resolve individual star-forming
regions on sub -- parsec scales and, at the same time,
assess the global picture on galactic scales. The Mag-
ellanic Clouds also contain a large range of very differ-
ent environments that can be observed to test theories
of star formation. The 30 Doradus (30 Dor) region in
the LMC, for example, is the most vigorous site of
current SF in the Local Group. Consequently, the sur-
roundings of 30 Dor are dominated by a strong far -- UV
radiation field. Possible triggering of SF on expand-
ing shells can be studied, e.g., in the LMC 4 region
1
2
Publications of the Astronomical Society of Australia
toward the northeast of the LMC (e.g. Braun et al.
1997; Yamaguchi et al. 2001; Cohen et al. 2003).
In
addition, the LMC and the Small Magellanic Cloud
(SMC) are gravitationally interacting with each other
and with the Galaxy (cf.
the numerical models of,
e.g., Murai & Fujimoto 1980; Gardiner et al. 1994;
Staveley-Smith et al. 2003; Connors et al. 2006), and
the motion of the LMC through the Galactic halo
causes strong ram pressure effects (e.g. de Boer et al.
1998a). It should be noted, however, that tidal or ram
pressure models alone cannot explain the full morphol-
ogy and dynamics of the gas and stars in the Magel-
lanic System, and a synthesis of both is likely required
(Mastropietro et al. 2005). All of these external inter-
actions, as well as factors internal to the LMC (e.g.,
shocks, density waves, and bar-induced motions), in-
fluence the ability to form stars on local and global
scales and can potentially be studied in the Magellanic
Clouds in great detail.
Also, with metallicities of ∼ 30% and ∼ 10% so-
lar in the LMC and SMC, respectively, cooling and
chemistry of gas in the Magellanic Clouds are clearly
different than in the Galaxy and the conditions may re-
semble those at higher redshifts, at times when galax-
ies were generally less evolved than today. Another
important property of the Magellanic Clouds and in
particular the LMC is that the numbers of resolved gi-
ant molecular clouds with associated star formation is
large. For example, Fukui et al. (1999) counted 107
molecular complexes in the 2.6′ (=37 pc) resolution
NANTEN 12CO(1-0) survey of the LMC. It is there-
fore possible to derive statistically meaningful results
for a large sample of regions, all at the same distance
and therefore resolution.
The location of 30 Dor and the surrounding mas-
sive atomic and molecular gas reservoir dominating
the south-eastern part of the LMC had led to many
speculations on its structure and star forming history.
Fujimoto & Noguchi (1990), for example and, more re-
cently, Bekki & Chiba (2007) were able to model this
region based on dynamical and hydro -- dynamical sim-
ulations of the LMC -- SMC interaction, not including
the Galaxy. Gardiner et al. (1998), on the other hand
attempted to simulate the amount of star formation
based on the dynamics of the off-center bar in the
LMC. Their simulations, though not designed to pre-
dict the gas distribution, do show that asymmetries
can be induced by the bar. de Boer et al. (1998a)
pointed out that the ages of structures in the LMC
increase clockwise from the eastern 30 Dor region to-
ward the north. They argued that the LMC may
undergo SF triggered by a bow-shock at the eastern
edge, where the ram pressure due to the LMC's mo-
tion through the Galactic halo is expected to be great-
est. The clockwise rotation of the LMC then turns the
newly formed, but now aging stellar populations to-
ward the north. In other words, ram pressure creates
a 'hot spot' of SF (currently 30 Dor) that remains in
the eastern part of the LMC while the galaxy rotation
brings fresh gas toward and newly formed stars away
from it. If this interpretation is correct, one may infer
that conditions upstream from 30 Dor, i.e., from the
east toward the south, may be those encountered just
before stars eventually form. This would be a very for-
tunate situation as it is usually difficult to predict the
timescales and locations of future SF in galaxies.
It is common knowledge that stars form out of
molecular gas, and indeed the region south of 30 Dor
contains a remarkably straight, elongated CO struc-
ture which we refer to as the "molecular ridge." 12CO(1-
0) maps from the Columbia survey (Cohen et al. 1988),
the NANTEN telescope (Fukui et al. 1999), and the
SEST key project (Kutner et al. 1997; Johansson et al.
1998) show that, with a mass of 107 M⊙, about 1/3
of all of the molecular gas that is traced by CO is
found in this structure stretching ∼ 2 kpc south of
30 Dor. Johansson et al. (1998) provides an analysis of
the N 159 and 30 Dor regions and they find kinetic tem-
peratures in the 10-50 K range. Kutner et al. (1997)
were able to describe the southern part of the molecu-
lar ridge but they restricted their analysis to a descrip-
tion of the general morphology of the feature and a
comparison of line strength to those in Galactic clouds.
In this paper we present first results of a compre-
hensive, high angular resolution 12CO(1-0) survey of
the entire molecular ridge obtained with the ATNF
Mopra1 telescope. The observations are the pilot for
the MAGMA project ('The Magellanic Mopra Assess-
ment') which is currently underway and which will
provide high resolution 12CO(1-0) maps of molecular
clouds across the entire LMC and SMC, mapping all
regions where NANTEN has detected emission, but
with a beam size ∼ 16 times smaller in area. In terms
of sensitivity and resolution, MAGMA is comparable
to the SEST key program but MAGMA encompasses
all CO traced molecular clouds rather than a selected
ensemble. In Sect. 2 we describe the observations and
data reduction, which is followed by the presentation
of our results in Sect. 3. A discussion of the results is
provided in Sect. 4 and a summary in Sect. 5.
2 Observations and Data Re-
duction
During the 2005 southern winter season, we observed
the molecular ridge close to 30 Dor with the single dish
Mopra telescope in the 12CO(1-0) molecular transition
(rest frequency: 115.271 GHz). Mopra was used in
the on -- the -- fly mapping mode and we observed about
∼ 120 5′
× 5′ maps in two orthogonal scanning direc-
tions with 1.8 h integration time each. The total time
on source was therefore of order 250 h. The Mopra
AT correlator was set up to observe at a bandwidth
of 64 MHz split into 1024 channels. This corresponds
to a velocity coverage of ∼ 166 km s−1 and a resolu-
tion of ∼ 0.16 km s−1 and the spectral window was
centered on a velocity of 230 km s−1 LSR. The ob-
servation of each map was preceded by pointing cali-
bration on the bright SiO maser R Dor, with typical
corrections less than 5". T ∗
A system temperatures of
the observations were usually ∼ 600 K and the data
1The Mopra radio telescope is part of the Australia Tele-
scope which is funded by the Commonwealth of Australia
for operation as a National Facility managed by CSIRO.
www.publish.csiro.au/journals/pasa
3
were calibrated against a warm absorber that was in-
serted every ∼ 30 minutes. The final data set has a
size of ∼ 0.6◦
× 2.0◦ which, at an assumed distance
of the LMC of 50 kpc, corresponds to 0.5 kpc×1.8 kpc,
the long side along declination. A map of the survey
region is shown in Fig. 1. The natural spatial resolu-
tion of Mopra at ∼ 115 GHz is about 35". The data
was reduced with the ATNF package livedata to derive
the quotient against line -- free reference positions up
to ∼ 0.5◦ away from the observed fields. First -- order
baselines fitted to line -- free channels were subtracted
from the resulting spectra. All data were regridded
to produce a position-position-velocity 3-dimensional
data cube using gridzilla. The 'inner' Mopra error
beam is in the range of 40"-80" which is similar to
the size of the emission and for the analysis here we
applied the extended beam efficiency of ηxb = 0.55
(Ladd et al. 2005) to convert T ∗
A to extended beam
brightnesses Txb. For better signal -- to -- noise ratios, the
data were eventually smoothed to a resolution of 45",
which corresponds to a physical length of ∼ 11 pc. The
rms noise of the data is about 0.4 K in a 0.16 km s−1
velocity channel. A comparison with earlier SEST data
at a similar resolution (Johansson et al. 1998) toward
the molecular gas in N 159 shows that the Mopra and
SEST fluxes agree within 10%.
3 Results
Maps of the molecular ridge derived from the 12CO(1-
0) data cube are displayed in Fig. 2. The integrated in-
tensity map clearly shows that the molecular ridge is a
slim (width: ∼ 100 − 200 pc) but long (∼ 1.8 kpc) fea-
ture along the declination axis. The brightest clumps
are toward the N159 region (see Fig. 2) and about half
a degree south of it at the 'kink' of the ridge. Be-
tween N159 and 30 Dor relatively little molecular gas
is observed. Using the Strong et al. (1988) 'standard'
Galactic CO intensity -- to -- H2 column density (XCO)
conversion factor of 2.3 × 1020 cm−2 (K km s−1)−1 the
detection limit (3 consecutive channels with 3σ detec-
tion each) is at about 1.3 × 1020 cm−2 (correspond-
ing to a mass of ∼ 200 M⊙ within a beam) and for
the molecular ridge we derive a total molecular gas
mass of about 3.2 × 106 M⊙. But note that studies
of CO with the Columbia 1.2 m, the SEST and NAN-
TEN telescopes have shown that the LMC XCO factor
may be a few times larger than this Galactic value
(Cohen et al. 1988; Israel 1997; Fukui et al. 1999, but
see Pineda et al.
2008). For their NANTEN sur-
vey data, Fukui et al. (1999) use XCO of 9 × 1020 NH2
(K km s−1)−1 and they derive a value of 4 × 107 M⊙
for the entire molecular gas in the LMC. If there are
no variations of XCO within the LMC, the ratio of the
fluxes should equal the ratio of masses and if we com-
pare our values to those of NANTEN, we confirm that
the molecular ridge contains about ∼ 1/3 of the total
molecular mass of the LMC.
4 Discussion
The very thin and long shape of the molecular ridge
would argue for it to be a coherent structure. A com-
parison of the H i and CO velocities in various position --
velocity diagrams along the ridge (Fig. 3, ATCA2 H i
observations are taken from Kim et al. 1998) shows,
however, that the molecular clumps are not confined to
a specific velocity component of the atomic gas. In ad-
dition, the molecular clumps are somewhat separated
from each other, only connected by some faint common
molecular envelope. This is even more obvious when
the 3-dimensional position-position-velocity data cube
is rotated around the declination axis. In a more global
context the molecular gas in the ridge finds itself on the
western edge of the south-eastern high-column density
region in the LMC, at a position where the H i column
densities are still high. However, the H i emission is
bifurcated in two velocity components, called the disk
and 'L' component (Luks & Rohlfs 1992). The veloc-
ity separation of the two components is very large to-
wards the eastern edge of the LMC, but they start to
merge into a single velocity component at about the
position of the molecular ridge (see the second moment
image in Fig. 1). This suggests that the clumps in the
molecular ridge were created by the same large-scale
triggering event and thus have similar formation his-
tories, not only in environmental terms but also on
similar timescales.
4.1 Molecular Cloud Formation
In Fig. 4, CO contours are overlaid over an integrated
H i column density map. The CO peaks are usually
found very close to H i peaks but they virtually never
coincide. Such a morphology was already indicated
by the lower resolution NANTEN data (resolution:
∼ 2.6′) and with the high angular resolution of Mo-
pra, this is unambiguously confirmed. If, instead of the
integrated H i column density, the H i peak flux is com-
pared to the 12CO(1-0) intensity, the peaks of the two
gas tracers are getting slightly closer to each other but
they still never coincide. To quantify this effect, pixel-
to-pixel correlations between the CO luminosity and
the H i column density and H i peak brightness tem-
perature are shown in Fig. 5. The CO bright pixels are
at intermediate H i column densities and brightnesses
and not at the largest values.
In particular, the H i
peak brightness values where the brightest CO emis-
sion is observed hovers around 60 to 100 K. In the fol-
lowing, we will discuss different mechanisms that may
be responsible for the shift of the CO and H i emission
peaks:
1. Rapid conversion of warm H i into H2. At a
certain threshold, the atomic gas combines into molec-
ular hydrogen and more than just the gas above the
threshold is converted, creating slight depressions in
the map of atomic H i. If the H i gas remains optically
2The Australia Telescope Compact Array is part of the
Australia Telescope which is funded by the Commonwealth
of Australia for operation as a National Facility managed
by CSIRO.
4
Publications of the Astronomical Society of Australia
thin one can determine the entire amount and distri-
bution of hydrogen atoms from both, H i and H2 in
the gas. In this scenario one would expect a smooth
distribution of hydrogen, some in atomic, and some in
molecular form. As an example we consider N159. The
atomic gas in this region has typical column densities
of ∼ 7 × 1021 cm−2 whereas the H i at the positions
of the CO peaks have a somewhat lower column of
N(H i)∼ 5 × 1021 cm−2. The CO peaks (smoothed to
the resolution of the H i map, 60′′) have intensities of
∼ 35 K km s−1. Converted with a Galactic XCO fac-
tor, this becomes an H2 column density of N(H2)∼
8 × 1021 cm−2 which adds to the aforementioned H i
column density of ∼ 5×1021 cm−2. In total, the hydro-
gen column density at the positions of the molecular
clouds is determined to be N(Htotal)=N(H i)+2 N(H2)
≈ 21 × 1021 cm−2. This is a factor of ∼ 3 larger than
the H i column that surrounds the molecular clumps
but where no CO is detected. Therefore, the total (H i
+ 2 H2) hydrogen column density map exhibits a jump
rather than a smooth transition whenever molecular
gas is present. This sharp edge cannot be softened by
a variation of the XCO factor. E.g., Fukui et al. (1999)
and Israel (1997) find that the LMC has an XCO fac-
tor a few times larger than the Galactic value. Using
such a higher XCO would result in larger H2 columns
and therefore in an even steeper rise of total hydrogen
column densities. If the XCO factor is changed to pro-
duce smooth hydrogen maps, where the H2 traced by
CO fills in the 'holes' of H i, one requires an XCO of
about 3 times less than the Galactic value. No study
of the LMC or of any other low metallicity object has
shown such low XCO factors before.
This comparison, however, is only valid if all molec-
ular gas is traced by CO. If there are extended re-
gions with molecular gas that is not traced by CO,
e.g., at low densities when CO may be dissociated by
the surrounding radiation field but H2 is not, the sharp
contrast in the total hydrogen map may be smoothed
out. Such an H2 phase is hard to detect and one of
the best approaches is UV absorption spectroscopy.
Tumlinson et al. (2002) performed such a survey to-
ward many sightlines in the LMC with FUSE. They
find that the molecular hydrogen, with a fraction of
∼ 1% (cf. Galaxy: ∼ 10%), is not a huge contribu-
tor to the entire gaseous ISM (see also the ORFEUS
UV absorption observation presented in de Boer et al.
1998b). They were also observing sightlines close to
30 Doradus where the UV radiation is very strong. If
there are large amounts of H2 that are not traced by
CO due to the higher dissociation energy of H2 they
should be most obvious toward that region.
Indeed,
the total column of H2 they find close to 30 Dor is with
∼ 1020 cm−2 one of the largest in their survey. How-
ever, the H i column density towards the same sight-
line is ∼ 70 times larger. Similar fractions are found
all across the LMC and the column of H2 in absorp-
tion never exceeds that of H i. It is therefore unlikely
that the contrast between the H i and H2 column den-
sities as traced by CO could be smoothed out with
extended layers of molecular gas that are undetected
in CO observations. This is in contrast to dust extinc-
tion studies which suggest that there may be large, not
CO traced reservoirs of molecular gas in the LMC (e.g.
Imara & Blitz 2007).
To conclude, the shifts of the H i and the CO peaks
are probably not simply a result of warm hydrogen
atoms (WNM) combining into H2. However, gravita-
tion may accumulate a lot of material quickly in the
densest molecular cores. Typical shifts between CO
and H i peaks are of order 2′, which corresponds to
∼ 30 pc. Within a typical free fall time of a Myr,
self-gravitation would accelerate the gas to a veloc-
ity of about 30 km s−1, a velocity gradient that is not
ruled out by the position velocity diagrams displayed
in Fig. 3.
2. WNM to CNM to H2 conversion. As shown
in Fig. 5, the H i brightness temperature, at the loca-
tion where most and the brightest molecular material
resides, falls into the range of 60 to 100 K which cor-
responds to the kinetic temperature of the cold neutral
medium in the Galactic disk (CNM, e.g., Dickey & Lockman
1990). A conversion of WNM to CNM can be achieved
by a shock.
Indeed, the atomic gas to the east of
the molecular ridge is prone to ram pressure effects
as this side is the leading edge towards the Milky Way
halo. The region is also influenced by tidal effects (e.g.
Staveley-Smith et al. 2003; Mastropietro et al. 2005)
and it has even been speculated that this region is
responsible for the origin of the Magellanic Stream
(Nidever et al. 2007). The second moment map of the
H i in Fig. 1 is in fact not dominated by the intrinsic
line width of the neutral gas but by the separation of
disk and 'L' velocity components. At the eastern edge
the separation is very large and the two components
become closer in velocity space toward the western di-
rection. At the position of the molecular ridge the
disk and 'L' components start to merge. To date the
origin of the bifurcation is still under discussion (e.g.
Bekki & Chiba 2007; Nidever et al. 2007) but whether
it is ram pressure or tidal effects or both, the shear
and shocks from such an interaction introduces energy
into the ISM that may reduce the likeliness to form
molecular clouds. Indeed, some molecular complexes,
the 'arc' clouds, are found in the region (Mizuno et al.
2001) but the bulk of molecular gas, the molecular
ridge, is found where the disk and 'L' components start
to merge into a single velocity component. If this in-
dicates that they are also spatially merging, one may
speculate that only the combination of both compo-
nents can provide enough shielding for the gas to cool
down and to become molecular.
In other words, shocks and/or shearing forces prop-
agating from the eastern edge may convert some of the
WNM into CNM. When the disk and 'L' component
merge they may provide enough shielding to convert
part of the CNM into molecular gas and the remain-
ing, cold H i becomes optically thick and thus levels off
at a brightness temperature close to the kinetic tem-
perature of the CNM.
3. H i accretion. The above process does not have
to be static. As discussed earlier, we cannot rule out
gas moving at speeds that are required to match free-
fall timescales. H i accretion after the formation of
www.publish.csiro.au/journals/pasa
5
the first molecular cores may be responsible for the
pile -- up of neutral gas close to the molecular clouds.
In this scenario, once the atomic gas comes close to
the molecular clumps, the gas is shielded, cooled, and
converted into H2 (see also Fukui 2007). H i accretion
may therefore be described as a cooling flow. Cooling
flows would result in decreasing H i line widths toward
the molecular cores. We do not observe this on the
scales of the H i-CO peak offsets but the most signifi-
cant cooling may happen only very close to the molec-
ular cores, too close to be resolved with the ATCA H i
observations at 21 cm.
4. Mechanical feedback from massive stars. Newly
formed, massive stars deposit large amounts of energy
into the surrounding interstellar medium (ISM) in the
form of strong stellar winds and supernovae. Virtu-
ally all star-forming galaxies exhibit an ISM that is
dominated by a wealth of expanding shells of all sizes.
Kim et al. (2007) catalog the expanding shells in the
LMC. Along the molecular ridge, the number density
of the shells seems to be lower than in the rest of the
LMC. It should therefore not have a significant effect
on the distribution of the atomic and molecular gas.
For some regions like the N 159 region, however, some
shell -- like structure is observed in H i and dust maps
(cf. Fig. 6) and the CO emission toward this position
may coincide with the inner side of the ring, similar
to what is observed toward other expanding Galactic
shells or LMC 4 (Yamaguchi et al. 2001; Cappa et al.
2005).
The scenario that best fits the observations is the
second one, in which shocks transform WNM into op-
tically thick CNM followed by conversion into H2, with
or without H i accretion. The very sharp rise of the to-
tal hydrogen column density inferred at the position of
the molecular gas is likely an artifact of high H i opti-
cal depth. On the other hand, the confinement of CO
emission to clumpy structures suggests an important
role for self-gravity in shaping the molecular clouds.
4.2 Star Formation
Typical H i column density thresholds in massive and
dwarf galaxies, above which SF is observed hover around
a canonical value of ∼ 1 × 1021 cm−2 (e.g. Walter et al.
2007). Below this threshold SF appears to be sup-
pressed. For most of the LMC this rule holds well.
However, the south-eastern region, dominated by ram
pressure and tidal forces of the LMC/SMC/Milky Way
interaction, appears to be different. Virtually no SF,
as traced by Hα, is observed in this region at positions
where the H i column density is below 5 × 1021 cm−2.
This may be due to the tidal and ram pressure forces
in the two gas components of the south -- eastern LMC.
On smaller scales, SF occurs in the very densest
molecular clumps. We conducted a survey with the
ATCA of the high-density gas tracers HCO+ and HCN
towards the star-forming regions N159 (40 pointing
mosaic) in the molecular ridge and toward N113 (7
pointing mosaic) at the western tip of the LMC bar.
The observations have a resolution of ∼ 7′′ which, at
the distance of the LMC, corresponds to a physical
length of ∼ 1.7 pc. Both of the observed molecules are
only excited at volume densities exceeding ∼ 104 cm−3.
HCO+, however, is also a tracer for photo-dominated
regions and can also form in more diffuse material
where the C+ abundance is high. Integrated HCO+ in-
tensity maps overlaid on 8µm Spitzer maps are shown
in Fig. 6. In both regions, the dense molecular gas typ-
ically overlaps or is adjacent to bright infrared (IR)
emission peaks. These peaks are the locations where
individual stars and stellar clusters eventually form
within the larger molecular complexes traced by CO.
Comparing the HCN and HCO+ maps, we find that,
after beam deconvolution, the HCO+ emission is about
20-30% more extended than the HCN clumps, in agree-
ment with the results presented in Wong et al. (2006).
The 8µm IR morphology of N159 appears to be ring --
like and indeed it is a 1-2 Myr old wind -- blown bubble
created by a central, massive star (Jones et al. 2005).
Most of the dense molecular gas is situated at the
rim of this shell, with relatively little observed at pro-
jected sightlines towards the interior. N113, in con-
trast, is a long, bent string of several individual knots,
both visible in the IR and in the HCO+ and HCN
maps. The string of knots curves from the south-east
towards the north. This morphology appears remi-
niscent of a circular arc, and close to the center of
this hypothetical circle, towards the north -- east, one
indeed finds HD 269219, a ∼ 30 M⊙ supergiant B star
and HD 269217, an emission-line star. Thus, the dens-
est pockets of molecular gas and the associated star-
forming regions appear to be located on a shock front
created by the two energy -- injecting massive stars, not
unlike the situation of the shell in N159.
The two examples, and the above discussion on
molecular cloud formation argues that shocks are the
major mechanism to trigger both, the formation of
low-density molecular clouds and the formation of star --
forming, high-density molecular clumps within these
larger structures.
5 Summary
The Large Magellanic Cloud is a unique object to study
molecular cloud and star formation on galaxy-wide
scales with high spatial detail and in a large range of
very different environments. A very prominent molec-
ular feature stretches from 30 Dor ∼ 1.8 kpc south-
ward;
it contains about 1/3 of the entire molecular
content of the LMC, as traced by CO. Our Mopra data
of this molecular ridge reveals the following:
• The structure is most likely not a coherent, large
molecular complex but consists of smaller molec-
ular clouds that are observed to be embedded
in different velocity components of the neutral
atomic gas. However, the global picture of the
LMC shows that the molecular ridge is at the in-
nermost edge of the south -- eastern high-density
region of the LMC, a region where tidal and
ram pressure forces have a major influence on
the state of the gas. The ridge coincides with
H i at high column density and is just where
6
Publications of the Astronomical Society of Australia
the disk and 'L' H i velocity components start to
merge. Thus, the molecular clouds in the ridge
are likely to share the same formation histories
and timescales.
• The local H i and CO peaks are typically dis-
placed by ∼ 30 pc. They are usually near each
other but hardly ever coincident. A likely expla-
nation for this morphology is that shocks propa-
gate from the eastern edge of the LMC and turn
warm H i into partly optically thick, cold H i
and eventually into molecular hydrogen. This
may happen in a static sense or via H i accre-
tion. The hydrogen column densities within the
molecular clumps are much larger than in the
surrounding material, which indicates that the
clouds are self -- gravitating. Other mechanisms,
such as feedback from massive stars may also be
responsible for the shift of H i and CO peaks but
are likely only playing a role for a small number
of molecular complexes.
• Very dense pockets of molecular gas, as traced
by HCO+ and HCN, are found near and coin-
ciding with strong IR emission. This strengthens
the suggested correlation of the very dense gas
and actual star formation rates. For the N159
and N113 regions we find the high density gas
peaks to be mostly located on shock fronts and
on the rims of expanding shells, powered by cen-
tral, massive stars.
Over the coming years, the study of molecular gas
in the LMC and SMC enters a new era. Our team
will continue to map all the molecular gas in both sys-
tems with the Mopra (MAGMA survey) and ATCA
telescopes.
In addition, many new facilities are cur-
rently under construction or have just been completed
that can be used to observe the LMC, such as APEX,
ASTE, NANTEN2, ALMA, as well as satellites such
as Herschel and Spitzer. Studying molecular gas in
the LMC provides a unique link between Galactic and
extragalactic observations and will be germane to the
full understanding of star formation on all scales.
References
Bekki, K., & Chiba, M. 2007, PASA, 24, 21
Braun, J. M., Bomans, D. J., Will, J.-M., & de Boer,
K. S. 1997,A&A, 328, 167
Cappa, C., Niemela, V. S., Mart´ın, M. C., & McClure-
Griffiths, N. M. 2005, A&A, 436, 155
Cohen, R. S., Dame, T. M., Garay, G., Montani, J.,
de Boer, K. S., Braun, J. M., Vallenari, A., & Mebold,
U. 1998, A&A, 329, L49
Dickey, J. M., & Lockman, F. J. 1990, ARA&A, 28,
215
Fujimoto, M., & Noguchi, M. 1990, PASJ, 42, 505
Fukui, Y., et al. 1999, PASJ, 51, 745
Fukui, Y. 2007, IAU Symposium, 237, 31
Gardiner, L. T., Sawa, T., & Fujimoto, M. 1994, MN-
RAS, 266, 567
Gardiner, L. T., Turfus, C., & Putman, M. E. 1998,
ApJL, 507, L35
Imara, N., & Blitz, L. 2007, ApJ, 662, 969
Israel, F. P. 1997, A&A, 328, 471
Johansson, L. E. B., et al. 1998, A&A, 331, 857
Jones, T. J., Woodward, C. E., Boyer, M. L., Gehrz,
R. D., & Polomski, E. 2005, ApJ, 620, 731
Kim, S., Staveley-Smith, L., Dopita, M. A., Freeman,
K. C., Sault, R. J., Kesteven, M. J., & McConnell,
D. 1998, ApJ, 503, 674
Kim, S., et al. 2007, ApJS, 171, 419
Kutner, M. L., et al. 1997, A&AS, 122, 255
Ladd, N., Purcell, C., Wong, T., & Robertson, S. 2005,
PASA, 22, 62
Luks, T., & Rohlfs, K. 1992, A&A, 263, 41
Mastropietro, C., Moore, B., Mayer, L., Wadsley, J.,
& Stadel, J. 2005, MNRAS, 363, 509
Mizuno, N., et al. 2001, PASJ, 53, 971
Murai, T., & Fujimoto, M. 1980, PASJ, 32, 581
Nidever, D. L., Majewski, S. R., & Butler Burton,
W. 2007, ArXiv e-prints, 706, arXiv:0706.1578 (ApJ
submitted)
Pineda, J. L., Ott, J., Klein, U., Wong, T., Muller, E.,
& Hughes, A. 2008, ApJ, submitted
Staveley-Smith, L., Kim, S., Calabretta, M. R.,
Haynes, R. F., & Kesteven, M. J. 2003, MNRAS,
339, 87
Strong, A. W., et al. 1988, A&A, 207, 1
Tumlinson, J., et al. 2002, ApJ, 566, 857
Rubio, M., & Thaddeus, P. 1988, ApJL, 331, L95
van der Marel, R. P., & Cioni, M.-R. L. 2001, AJ, 122,
Cohen, M., Staveley-Smith, L., & Green, A. 2003, MN-
1807
RAS, 340, 275
Walter, F., et al. 2007, ApJ, 661, 102
Connors, T. W., Kawata, D., & Gibson, B. K. 2006,
Wong, T., Whiteoak, J. B., Ott, J., Chin, Y.-n., &
MNRAS, 371, 108
Cunningham, M. R. 2006, ApJ, 649, 224
de Boer, K. S., Richetr, P., Bomans, D. J., Heithausen,
A., & Koorneef, J. 1998, A&A, 338, L5
Yamaguchi, R., Mizuno, N., Onishi, T., Mizuno, A., &
Fukui, Y. 2001, ApJL, 553, L185
www.publish.csiro.au/journals/pasa
7
Figure 1: H i column density map of the LMC
(taken from Kim et al. 1998, beam size: 1′) with
the contours of our molecular ridge CO data over-
laid in black. The white polygon outlines the re-
gion observed. The yellow boxed insert is the same
region as the one marked on the column density
map but it displays the second moment of the
H i data (for better signal -- to noise, this map was
smoothed to three times the resolution of the inte-
grated H i map.). Note that toward the south-east
the second moment is dominated by the velocity
difference of the disk and 'L' components rather
than by the line widths of the neutral gas.
8
Publications of the Astronomical Society of Australia
Figure 2: Mopra 12CO(1-0) maps of the molecular
ridge close to 30 Dor. From left to right:
Inte-
grated intensity in K km s−1, intensity -- weighted
velocities (moment 1), dispersions (moment 2).
www.publish.csiro.au/journals/pasa
9
Figure 3: Position velocity cuts through the CO
and H i data cubes. The arrows in the left panel
indicate the position, direction, and length where
the cuts are taken. The corresponding position-
velocity plots are shown in the middle and to the
right. Arrows of the same color belong together;
they are also coded by letters 'A'-'D'. The color
scale are data from the H i cube and the blue con-
tours are the CO data. Note that the CO at the
30 Dor region exhibits H i in absorption (start of
cut 'C').
10
Publications of the Astronomical Society of Australia
Figure 4: CO contours of the region close to N159
overlaid as contours on an integrated H i column
density map. Note that the CO peaks are close to
but never coincident with nearby peaks of H i.
www.publish.csiro.au/journals/pasa
11
Figure 5: Pixel -- based scatter plots of the 12CO(1-
0) luminosity as a function of H i column density
(top) and H i peak brightness (bottom). The con-
tours are levels of constant point density in loga-
rithmic spacings.
12
Publications of the Astronomical Society of Australia
Figure 6: The star-forming regions N159 (left) and
N113 (right) in the LMC. The contours show the
integrated HCO+ emission as observed with the
ATCA on top of Spitzer 8µm maps. The green
lines outline the mapped regions.
|
0805.4289 | 1 | 0805 | 2008-05-28T09:03:31 | V405 Aurigae: A High Magnetic Field Intermediate Polar | [
"astro-ph"
] | Our simultaneous multicolor (UBVRI) circular polarimetry has revealed nearly sinusoidal variation over the WD spin cycle, and almost symmetric positive and negative polarization excursions. Maximum amplitudes are observed in the B and V bands (+-3 %). This is the first time that polarization peaking in the blue has been discovered in an IP, and suggests that V405 Aur is the highest magnetic field IP found so far. The polarized flux spectrum is similar to those found in polars with magnetic fields in the range B ~ 25-50 MG. Our low resolution circular spectropolarimetry has given evidence of transient features which can be fitted by cyclotron harmonics n = 6, 7, and 8, at a field of B = 31.5 +- 0.8 MG, consistent with the broad-band polarized flux spectrum. Timings of the circular polarization zero crossovers put strict upper limits on WD spin period changes and indicate that the WD in V405 Aur is currently accreting closely at the spin equilibrium rate, with very long synchronization timescales, T_s > 10^9 yr. For the observed spin to orbital period ratio, P_{spin}/P_{orb} = 0.0365, and P_{orb} ~ 4.15 hr, existing numerical accretion models predict spin equilibrium condition with B ~ 30 MG if the mass ratio of the binary components is q_1 ~ 0.4. The high magnetic field makes V405 Aur a likely candidate as a progenitor of a polar. | astro-ph | astro-ph |
DRAFT VERSION OCTOBER 25, 2018
Preprint typeset using LATEX style emulateapj v. 08/22/09
V405 AURIGAE: A HIGH MAGNETIC FIELD INTERMEDIATE POLAR
V. PIIROLA1, T. VORNANEN, & A. BERDYUGIN
Tuorla Observatory, University of Turku, FI-21500 Piikkiö, Finland
AND
G.V. COYNE, S.J.
Vatican Observatory, V-00120 Città del Vaticano
Draft version October 25, 2018
ABSTRACT
Our simultaneous multicolor (UBV RI) circular polarimetry has revealed nearly sinusoidal variation over the
WD spin cycle, and almost symmetric positive and negative polarization excursions. Maximum amplitudes
are observed in the B and V bands (±3%). This is the first time that polarization peaking in the blue has been
discovered in an IP, and suggests that V405 Aur is the highest magnetic field IP found so far. The polarized flux
spectrum is similar to those found in polars with magnetic fields in the range B ∼ 25-50 MG. Our low resolution
circular spectropolarimetry has given evidence of transient features which can be fitted by cyclotron harmonics
n= 6, 7, and 8, at a field of B = 31.5 ± 0.8 MG, consistent with the broad-band polarized flux spectrum. Timings
of the circular polarization zero crossovers put strict upper limits on WD spin period changes and indicate that
the WD in V405 Aur is currently accreting closely at the spin equilibrium rate, with very long synchronization
timescales, Ts > 109 yr. For the observed spin to orbital period ratio, Pspin/Porb=0.0365, and Porb ∼ 4.15 hr,
existing numerical accretion models predict spin equilibrium condition with B ∼ 30 MG if the mass ratio of
the binary components is q1 ∼ 0.4. The high magnetic field makes V405 Aur a likely candidate as a progenitor
of a polar.
Subject headings: stars: binaries (close) -- novae, cataclysmic variables -- stars: magnetic fields -- polariza-
tion -- accretion -- stars individual: V405 Aur
1. INTRODUCTION
V405 Aur (RX J0558.0+5353) was discovered in the
ROSAT all-sky survey (Haberl et al. 1994) and was identified
with a V ∼ 14.5 Cataclysmic Variable (CV), showing char-
acteristics of an Intermediate Polar (IP): the soft X-ray flux
was modulated at 272.7 s, which suggested that the system
consists of a spinning magnetic white dwarf (WD) accreting
matter from a cool low-mass companion. An orbital period of
4.15 h was deduced from optical spectroscopy. Further anal-
ysis of the ROSAT data at energies higher than 0.7 keV re-
vealed a period of 545 s, which is the true spin period of the
WD, and therefore the soft X-ray light curve has a double-
peaked structure (Allan et al. 1996; Evans & Hellier 2004).
Evidence of two-pole accretion has been found also from op-
tical spectroscopy (Still et al. 1998; Harlaftis & Horne 1999).
V405 Aur belongs to a small group of IPs which are unusu-
ally bright in soft X-rays, and have also been found to emit
polarized radiation in the optical and/or near IR, resembling
what we see in the strictly (or nearly) synchronous magnetic
CVs, polars. The polarized IPs known to date are BG CMi
(Penning et al. 1986; West et al. 1987), PQ Gem (Piirola et
al. 1993; Rosen et al. 1993; Potter et al. 1997), V2400 Oph
(Buckley et al. 1997), V405 Aur (Shakhovskoy & Kolesnikov
1997), V2306 Cyg (Uslenghi et al. 2001; Norton et al. 2002),
and RX J2133.7+5107 (Katajainen et al. 2007). Of these IPs,
PQ Gem was the first found to show spin-modulated circu-
lar and linear polarization variations, arising from cyclotron
emission from two accretion regions near the opposite mag-
netic poles of the spinning WD.
IPs are generally believed to have lower magnetic field (B <
Electronic address: [email protected]
1 Visiting Astronomer, Vatican Observatory, V-00120 Città del Vaticano
10 MG, see e.g. Warner 1995) than polars, where the deter-
mined fields range from 7 MG (V2301 Oph, Ferrario et al.
1995) to 230 MG (AR UMa, Schmidt et al. 1996). Typical
values for polars are a few tens of MG (see Cropper 1990 for
a review), and the magnetic field is sufficiently high to pre-
vent the formation of an accretion disk. The stream of matter
from the companion is coupled to the magnetic field lines and
channeled via accretion columns onto the WD surface. In IPs
accretion disk may exist, but is truncated at some inner ra-
dius where flow to the WD takes place via accretion curtains
(Rosen et al., 1988; Ferrario et al., 1993).
IPs predominantly have longer orbital periods (4-5 h) than
polars (typically < 2 h), though there is considerable over-
lap. Close binaries lose angular momentum via gravitational
radiation and magnetic wind braking, and therefore evolve
towards shorter orbital periods. For sufficiently high field
strengths and small binary separations the magnetic lock-
ing torque will start to spin-down the WD and synchronism
may be reached, the system thus becoming a polar. For IPs
magnetic field measurements are difficult, but from polarized
flux spectrum estimates have been made for PG Gem (8-18
MG, Piirola et al. 1993; 9-21 MG, Väth et al. 1996; Pot-
ter et al. 1997), V2400 Oph (9-20 MG, Väth 1997), and RX
J2133.7+5107 (& 20 MG, Katajainen et al. 2007). These val-
ues are well within the range of the magnetic fields seen in
polars and suggest that some of the polarized IPs may evolve
into polars. So far only one of the polarized IPs, PQ Gem, has
been shown to have the WD spinning down (Mason 1997).
In the present paper we report extensive multicolour
(UBV RI) polarimetric observations of V405 Aur, carried out
in order to put constraints on the magnetic field strength of the
WD and on the system geometry. These observations were
complemented with low-resolution circular spectropolarime-
2
Piirola et al.
try. We have also searched for possible spin period changes
by a long-term monitoring of phase shifts in the circular po-
larization curves.
2. OBSERVATIONS
Broad-band multicolor polarimetry was carried out at the
2.5 m Nordic Optical Telescope (NOT) at Roque de los
Muchachos Observatory on La Palma. The instrument (Tur-
Pol) provides strictly simultaneous measurements in five pass-
bands close to the UBV RI system, by using four dichroic fil-
ters to split the light into the different spectral regions (Piirola
1988). In the circular polarimetry mode the superachromatic
quarter-wave (λ/4) plate was rotated in 90◦ steps above the
plane-parallel calcite plate polarizing beam splitter. For each
orientation of the retarder, the two polarized beams were in-
tegrated with a 25 Hz chopping frequency for a 5 s total inte-
gration time each. With some dead time involved in the chop-
ping procedure, the resulting time resolution is about 12 s for
photometry. One circular polarization observation consists of
two integrations with the λ/4 plate in orthogonal orientations,
and the time resolution is correspondingly about 24 s. In the
simultaneous linear and circular polarization mode the λ/4
plate was rotated in 22.5◦ steps. One complete observation of
linear and circular polarization consists of eight integrations,
and the time resolution is correspondingly about 1.6 minutes.
A summary of the polarization observations is given in Table
1.
For WD spin period timing we made circular polariza-
tion observations of V405 Aur also at the 60 cm reflector of
Tuorla observatory, using a newly developed CCD polarime-
ter DIPOL (Piirola et al. 2005). Because of the smaller aper-
ture of the telescope, the observations were done in unfiltered
light. The polarimeter is equipped with a thinned back illumi-
nated CCD (Marconi) which has a high blue sensitivity. The
time resolution of these observations is about 1 min. Some
complementary CCD-photometric observations were carried
out also at the 60 cm KVA telescope on La Palma.
Low resolution circular spectropolarimetry was done with
the ALFOSC+FAPOL instrument at the NOT. With the grism
#11 and 1.8" slit the spectral resolution was about 40Å. An
achromatic λ/4 plate was rotated in 90◦ steps above the slit
by the polarimetry unit (FAPOL), controlled via the ALFOSC
user interface. A plane parallel calcite plate below the slit pro-
duced two perpendicularly polarized spectra on the CCD. The
two polarized beams passed through the collimator, grism,
and the camera onto the CCD. To improve time resolution, cir-
cular polarization spectra were deduced from individual 120 s
exposures. Null calibration was done with the help of unpolar-
ized standard stars. To check for the circular polarization sign
and scale, observations of the highly polarized star Grw+70
8247 were made.
3. RESULTS AND DISCUSSION
3.1. Multicolor photopolarimetry
Figure 1 illustrates our simultaneous UBV RI circular po-
larimetry. The data collected on two different nights show
nearly sinusoidal and remarkably symmetric modulation, with
extreme values near ±3% in the B and V , and about ±2% in
the U and R bands. In the I band the modulation is smaller,
about ±1%. This is the first time that polarization peaking in
the blue has been detected in an IP, and suggests that V405
Aur is the highest field IP discovered so far. Its magnetic field
is probably similar to what is typically found in polars (25-
50 MG). Nearly symmetric positive and negative excursions
FIG. 1. -- Simultaneous UBVRI circular polarimetry of V405 Aur on 1997
Nov 27 (left) and 2003 Sep 22 (right). Phases refer to the spin period of the
WD given in Eq. 2.
in the circular polarization curves indicate that two opposite
magnetic poles (positive and negative) are accreting at similar
rate in V405 Aur.
The left hand panels in Figure 1 show evidence of a small
dip in circular polarization near phase 0.25, where the positive
magnetic pole points closest to us. This is due to cyclotron
beaming effect. Cyclotron emission intensity is at maximum
in the direction perpendicular to the magnetic field, and de-
creases to zero when looking along magnetic field lines. Un-
polarized background radiation from thermal sources dilutes
the observed degree of circular polarization when the polar-
ized cyclotron flux decreases.
Our simultaneous UBV RI light curves in Figure 2 show
minima near the phases 0.25 and 0.75 where we are look-
ing closest along the field lines, and maxima at the phases
0.0 and 0.5 where our line sight is perpendicular to the field
lines, as seen from the zero-crossings of circular polariza-
tion. This photometric behaviour is in accordance with the
cyclotron beaming effect, i.e., maximum emission takes place
in directions perpendicular to the magnetic field.
Many, but not all, polars show a pulse of linear polariza-
tion when our line of sight is nearly perpendicular to the field
lines. We have found no evidence of any linear polarization
pulse in V405 Aur. Linear polarization appears to be con-
stant at the level of PB ∼ 0.36 ± 0.03%, θ ∼ 138◦± 4◦and is
probably of interstellar origin. Fitting the modified Serkowski
law for the wavelength dependence of interstellar polarization
(Whittet et al. 1992) to the average polarization in the UBV RI
bands gives maximum polarization Pmax = 0.44 ± 0.04%, at
λmax = 0.56 ± 0.11µm.
The zero-crossings of circular polarization give an accu-
V405 Aur: A High Magnetic Field IP
3
SUMMARY OF POLARIMETRIC OBSERVATIONS
TABLE 1
UT Date
HJD-24450000
Passbands
Telescope
Instrument Mode
N
1997 Nov 25
Nov 27
Nov 28
2001 Feb 04
Feb 05
2002 Jan 25
2003 Jan 10
Jan 13
Jan 16
Jan 30
Jan 31
Feb 07
Feb 19
Feb 20
Mar 23
Apr 27
Sep 20
Sep 21
Sep 22
Oct 18
2004 Feb 26
Feb 26
Feb 28
Feb 29
Mar 01
778.5566 - .6621
780.5380 - .6806
781.5224 - .6103
1945.4355 - .5224
1946.4423 - .5214
2300.3837 - .5019
2650.2704 - .3418
2653.3405 - .4482
2656.3313 - .4104
2670.3613 - .4923
2671.2289 - .3641
2678.2729 - .3721
2690.2659 - .3872
2691.2158 - .2945
2722.2757 - .3732
2757.3800 - .3950
2903.6210 - .6660
2904.6075 - .7285
2905.6025 - .7385
2931.5679 - .5967
3062.3398 - .3597
3062.3713 - .4192
3064.4495 - .4805
3065.4080 - .4176
3066.3603 - .3921
UBV RI
UBV RI
UBV RI
UBV RI
UBV RI
UBV RI
no filter
"
"
"
"
"
"
"
"
"
B
"
"
UBV RI
UBV RI
UBV RI
no filter
spectrop.
no filter
NOT
NOT
NOT
NOT
NOT
NOT
T-60
T-60
T-60
T-60
T-60
T-60
T-60
T-60
T-60
NOT
NOT
NOT
NOT
NOT
NOT
NOT
NOT
NOT
NOT
TurPol
TurPol
TurPol
TurPol
Turpol
Turpol
DIPOL
DIPOL
DIPOL
DIPOL
DIPOL
DIPOL
DIPOL
DIPOL
DIPOL
ALFOSC
Turpol
Turpol
Turpol
ALFOSC
ALFOSC
ALFOSC
ALFOSC
ALFOSC
ALFOSC
SLC
SLC
L
SLC
SLC
SLC
L
C
C
C
C
C
C
C
C
C
C
C
C
L
C
C
C
C
C
46
198
101
36
30
96
148
136
100
150
156
116
154
100
118
48
41
192
216
30
64
24
96
32
96
in light curves take place at the circular polarization zero-
crossings, and our ephemeris from circular polarization is ac-
curate enough to solve without ambiguity which of the two
maxima corresponds to the positive cross-over of circular po-
larization. Weighted second order polynomial fit to the tim-
ings obtained over about 12 year interval yields a greatly im-
proved ephemeris for the WD spin:
T◦ = HJD 2 449 681.46389(5)
+ 0.0063131474(4) × E + 4(4) × 10- 16 × E2
(1)
The numbers in parentheses give the error estimates in units
of the last digit of the determined coefficient. No statistically
significant second-order term is detected. A weighted linear
fit to the timings gives the following ephemeris for the positive
crossover of circular polarization:
T◦ = HJD 2 449 681.46387(2)
+ 0.0063131476(3) × E.
(2)
The residuals (O-C) of the observed times of positive
crossover from those predicted by the linear ephemeris are
shown in Figure 3. There is no clear trend suggesting any sig-
nificant higher-order terms. The rather tight limits for WD
spin period changes imply very long spin-down timescales
(Ts > 109 yr) for V405 Aur in its current evolutionary and
accretion state.
3.2. Cyclotron model fittings
Observed polarized fluxes are independent of any unpolar-
ized background sources (accretion stream, disk, WD photo-
sphere) and therefore very useful for comparing with exist-
ing cyclotron models. Broad-band polarized flux spectrum
may give rough estimates of the magnetic field in the cy-
clotron emission region, as cyclotron emission intensity and
polarization are characteristic to each harmonic n = ω/ωc of
the fundamental cyclotron frequency ωc (see e.g. Wickra-
masinghe & Meggitt 1985). The wavelength λc which cor-
responds to ωc is in turn related to the magnetic field strength
FIG. 2. -- Simultaneous UBVRI photometry of V405 Aur. Phases refer to
the spin period of the WD given in Eq. 2.
rate measure of the rotation of the WD. We have used these
timings to search for possible changes in the WD spin pe-
riod and evidence of spin-up or spin down. Table 2 lists
our observations of the time of the positive cross-over of
circular polarization. We have included in Table 2 also a
few photometric timings to extend the time span. Maxima
4
Piirola et al.
TABLE 2
SPIN CYCLE TIMINGS FOR V405 AUR
Cycle No. HJD-24400000
O-C(d)
σ(d)
Filter Note
1.
1397.
2997.
13127.
14252.
14443.
16292.
16628.
16758.
16917.
173780.
173780.
173780.
173780.
173780.
173780.
173780.
173780.
174093.
174093.
174093.
174093.
174093.
174093.
174093.
174093.
358613.
358613.
358613.
358613.
358613.
358613.
358613.
358772.
358772.
414836.
414836.
414836.
414836.
470258.
470745.
471218.
473441.
473578.
474694.
476594.
476744.
481664.
487224.
510545.
510545.
510545.
510545.
510545.
510545.
510545.
510545.
510703.
510703.
510703.
510703.
510703.
510703.
510703.
510703.
535530.
535530.
535864.
536016.
536016.
536167.
702502.
49681.47030
49690.28330
49700.38450
49764.33660
49771.43890
49772.64480
49784.31760
49786.43880
49787.25960
49788.26340
50778.56263
50778.56256
50778.56254
50778.56256
50778.56285
50778.56280
50778.56282
50778.56272
50780.53873
50780.53868
50780.53861
50780.53858
50780.53866
50780.53864
50780.53861
50780.53860
51945.44078
51945.44077
51945.44080
51945.44078
51945.44069
51945.44073
51945.44072
51946.44450
51946.44458
52300.38478
52300.38479
52300.38474
52300.38476
52650.27201
52653.34659
52656.33269
52670.36694
52671.23165
52678.27720
52690.27218
52691.21913
52722.27984
52757.38082
52904.60991
52904.60991
52904.60989
52904.60985
52904.60973
52904.60979
52904.60975
52904.60979
52905.60741
52905.60736
52905.60731
52905.60728
52905.60735
52905.60734
52905.60734
52905.60733
53062.34390
53062.34373
53064.45240
53065.41211
53065.41195
53066.36538
54116.46291
0.00012
-0.00003
0.00013
0.00004
0.00005
0.00014
-0.00007
-0.00009
0.00000
0.00001
-0.00004
-0.00011
-0.00013
-0.00011
0.00018
0.00013
0.00015
0.00005
0.00005
-0.00001
-0.00007
-0.00011
-0.00003
-0.00005
-0.00007
-0.00009
0.00009
0.00008
0.00011
0.00009
0.00000
0.00004
0.00003
0.00002
0.00010
-0.00002
-0.00001
-0.00006
-0.00003
-0.00006
0.00002
0.00000
0.00013
-0.00007
0.00001
0.00001
-0.00001
0.00001
-0.00011
0.00006
0.00006
0.00004
0.00000
-0.00012
-0.00006
-0.00010
-0.00006
0.00009
0.00004
-0.00001
-0.00004
0.00003
0.00002
0.00002
0.00001
0.00006
-0.00011
-0.00003
0.00008
-0.00008
0.00006
0.00018
0.00005
0.00005
0.00005
0.00005
0.00005
0.00005
0.00005
0.00005
0.00005
0.00005
0.00009
0.00004
0.00004
0.00007
0.00006
0.00005
0.00005
0.00008
0.00004
0.00003
0.00003
0.00004
0.00007
0.00005
0.00005
0.00006
0.00007
0.00007
0.00009
0.00002
0.00002
0.00003
0.00003
0.00010
0.00014
0.00003
0.00004
0.00002
0.00002
0.00005
0.00007
0.00004
0.00007
0.00003
0.00009
0.00004
0.00002
0.00004
0.00002
0.00003
0.00002
0.00003
0.00004
0.00003
0.00002
0.00003
0.00003
0.00002
0.00002
0.00002
0.00003
0.00002
0.00002
0.00002
0.00002
0.00005
0.00004
0.00003
0.00005
0.00005
0.00002
0.00008
no
no
no
no
no
no
no
no
no
no
U
B
V
R
U
B
V
R
U
B
V
R
U
B
V
R
B
V
R
U
B
V
R
B
V
U
B
V
R
no
no
no
no
no
no
no
no
no
no
U
B
V
U
B
V
R
I
U
B
V
U
B
V
R
I
no
no
no
no
no
no
no
1
1
1
1
1
1
1
1
1
1
2
2
2
2
3
3
3
3
2
2
2
2
3
3
3
3
2
2
2
3
3
3
3
2
2
3
3
3
3
4
4
4
4
4
4
4
4
4
5
2
2
2
3
3
3
3
3
2
2
2
3
3
3
3
3
5
6
5
5
6
5
7
NOTE. -- (1) Photometry from Allan et al. 1996, (2) UBV RI polarimetry at the
NOT, (3) UBV RI photometry at the NOT (4) CCD polarimetry at T-60, (5) CCD
polarimetry at the NOT, (6) CCD photometry at the NOT, (7) CCD photometry at
the KVA-60
V405 Aur: A High Magnetic Field IP
5
i, of the WD spin axis and the colatitude, β, of the cyclotron
emission region(s). This adds further uncertainty in the mod-
eling of the cyclotron flux spectrum (Figure 4), which is de-
pendent on the angle between our line of sight and the mag-
netic field. The symmetric shape of the circular polarization
curves over the WD spin period (Figure 1) requires that β is
large. This was noted already by Shakhovskoy & Kolesnikov
(1997), who suggested a nearly equatorial magnetic field in
V405 Aur.
The small dip at the circular polarization maximum (Fig-
ure 1, left panel) indicates that when the emission region is
pointing closest to us the viewing angle, α, reaches a value
where the beaming effect starts to reduce cyclotron flux, but
does not yet drop it substantially. Less clear evidence of a
beaming effect is seen on the right hand panel of Figure 1 sug-
gesting that probably the emission region had moved slightly
towards larger α values, and consequently the observed cy-
clotron emission was less influenced by the beaming effect on
that night.
With the above constraints from circular polarization
curves, we have tried to reproduce the observed polarization
and light curves of V405 Aur with a geometric model involv-
ing extended emission strips on the WD surface. We have
applied the viewing angle dependence of the cyclotron flux
polarization and intensity from Wickramasinghe & Meggitt
(1985), as done in the earlier work by Piirola et al. (1990,
1993, 1994), where the models are also explained in more
detail. The minimum total flux determined from each of the
UBV RI light curves was adopted as the wavelength depen-
dent unpolarized background flux in the model computations.
This is a good approximation as the relative amount of cy-
clotron emission in the total observed flux is small in V405
Aur. Centered, or slightly de-centered, dipole field was as-
sumed for computing the direction of field lines relative to the
normal to the WD surface for the emission arcs located off
from the magnetic poles.
The nearly symmetric rising and descending parts of the
circular polarization curves suggest that the tilt of magnetic
field lines with respect to the normal to the WD in the emis-
sion region is not large in the direction of longitude, i.e., the
cyclotron region is not located far from the magnetic pole in
that direction. Such accretion geometry may be favored in
spin equilibrium case. In contrast, for PQ Gem Potter et al.
(1997) have found the cyclotron arc located ahead of the mag-
netic pole in longitude, which is consistent with the spin-down
of the WD in this system. Estimates from X-ray spectral fits
(Evans & Hellier 2004) give rather small values of the black-
body emitting area (< 0.001 of the WD surface) associated
with the accretion region. Hence, we assume relatively nar-
row cyclotron emission arcs in our model.
The main features of the circular polarization and light
curves can be reproduced reasonably well (Figures 5 and 6),
but the cyclotron flux from the constant temperature models
drops too much when moving away from the wavelength of
the peak polarized flux. Hence, the amplitude of the circu-
lar polarization and intensity variations becomes too small for
the U and I bands (top and bottom panels, respectively). Also
the beaming effect, the dip of polarization at phases 0.25 and
0.75, becomes too pronounced at high harmonics. Producing
a broader cyclotron flux spectrum would require inhomoge-
neous emission regions with spread in the physical parame-
ters, the electron density, temperature, and the magnetic field.
Such modeling is, however, beyond the scope of the present
paper.
FIG. 3. -- Spin cycle timings of V405 Aur. Residuals from the weighted
linear fit (Eq. 2) are plotted vs. cycle number. Different symbols denote
photometry (◦) and polarimetry (+).
FIG. 4. -- Observed peak circularly polarized fluxes of V405 Aur, com-
pared with those deduced from existing cyclotron models. The solid line
corresponds to the extended emission region model (Wickramasinghe et al.
1991) with a field of 26 MG and average view angle ∼60◦. Dot-dashed and
dashed lines give fluxes from the 20 kEV constant temperature, Λ = 106,
model (Wickramasinghe & Meggitt 1985) for viewing angles α = 70◦ and
50◦.
by λc ∼ 108µm/B[MG]. The method has been applied to the
polarized IPs PQ Gem (Piirola et al. 1993; Väth et al. 1996;
Potter et al. 1997) and V2400 Oph (Väth 1997) and the sug-
gested field strengths are in the range 8-20 MG in both of
these objects.
Polarized flux spectrum is strongly dependent on the pa-
rameters of the adopted cyclotron model and the system ge-
ometry and this limits the usefulness of the method. How-
ever, it may be the only way of getting any magnetic field
estimates at all for IPs which do not show measurable cy-
clotron harmonic features or photospheric Zeeman features in
their spectra, because of the overlaying strong disk and stream
emission.
Figure 4 gives the peak circularly polarized fluxes in the
UBV RI bands, computed from the data displayed in Figures 1
and 2, compared with those deduced from existing cyclotron
models. With constant temperature Λ = 106 models (Wick-
ramasinghe & Meggitt 1985) the polarized flux distribution is
too narrow (dashed and dotted lines). Broader distribution and
better fit is obtained for extended ribbon-like accretion shocks
model of Wickramasinghe, Wu, & Ferrario (1991), where al-
lowance is made for field spread and for the change in shock
height as a function of specific accretion rate. The fit shown
in Figure 4 (solid line) corresponds to a field of B ∼ 26 MG.
Without measurable linear polarization from the cyclotron
source it is not possible to put strict constraints on the geo-
metric model, which is defined basically by the inclination,
6
Piirola et al.
FIG. 5. -- Examples of simulated circular polarization curves for i = 38◦
and two nearly equatorial cyclotron emission regions extended by ∆λ = 24◦
in longitude and confined between colatitudes β1 = 80-92◦and β2 = 76-88◦
for the positive and negative regions, respectively. Each of the emission re-
gions has been divided into four equidistant strips with 4◦ intervals in β for
the modeling shown in the left panel. The right hand panel displays curves
from the accretion model shown in Figs. 7 and 8, compared with the circular
polarization normal points of 2003 Sep 22 in the UBV RI bands (from top to
bottom).
The results from our computations shown in Figures 5 and 6
support the earlier suggestion by Shakhovskoy & Kolesnikov
(1997) that the magnetic dipole axis is nearly equatorial in
V405 Aur (Figures 7 and 8). With such a large value of the
colatitude, β ∼ 90◦ , best fits to our circular polarization and
light curves are obtained with the WD spin axis inclination in
the range i ∼ 30-50◦. To put better constraints on the system
geometry of V405 Aur, further efforts to establish possible
variability of the linear polarization (P, θ) would be required.
3.3. Comparison with X-ray data
Soft X-rays in polarized IPs are dominated by blackbody
emission from the heated WD surface near the accretion
shock. The double-peaked structure of the soft X-ray light
curve in V405 Aur indicates that both accretion regions con-
tribute. The equality of the two maxima requires that the angle
between the magnetic and spin axes is high (Evans & Hellier
2004), as also suggested earlier by the circular polarization
curve (Shakhovskoy & Kolesnikov 1997) and our modeling
in the present paper (Sect 3.2).
The details of the geometric model of Evans & Hellier
(2004) from X-ray data are not fully supported by the circular
FIG. 6. -- Simulated light curves for the same models as in Figure 5, com-
pared with the light curves from 2003 Sep 22 in the UBV RI bands (dots, from
top to bottom).
polarization curves. With their values of spin and magnetic
axis inclination (i=65◦ and β1=60◦) the upper magnetic pole
passes near the center of the visible disk of the WD when
pointing closest to us, and the corresponding viewing angles,
α ∼ 0◦ would give prominent cyclotron beaming effects: cir-
cular polarization would drop to zero in all wavebands. Our
circular polarization curves (Figure 1) show only minor de-
pressions at phases 0.25 and 0.75 in the U band and marginal
evidence in the B and V bands, suggesting that the viewing
angle, α > 40◦ throughout the spin cycle. The symmetric po-
larization and light curves require the two accretion regions
be nearly at the same latitude. For nearly centered dipole this
means nearly equatorial magnetic field. More exotic possi-
ble configurations might be a grossly de-centered (in latitude)
dipole, or quadrupole field with one positive and one negative
pole visible (for parts of the spin cycle) and nearly at the same
latitude.
Our improved ephemeris (Eq. 2) allows us to phase some
of the published X-ray curves with our (UBV RI) circular po-
larization and light curves. In hard X-rays V405 Aur shows
a single peak, and the epoch of the maximum given by de
Martino et al. (2004), HJDmax = 2 450 364.159695, occurs at
the phase 0.74 of our ephemeris. This is very near the phase
(0.75) where the negative magnetic pole is pointing closest
to us, and means that this pole is seen brighter in hard X-
rays. Evans & Hellier (2004) define phase zero of their XMM-
V405 Aur: A High Magnetic Field IP
7
FIG. 7. -- Illustration of the model used to compute the circular polarization and light curves shown in Figs. 5 and 6. The spinning WD is seen at inclination
i=38◦ and the magnetic dipole angle is β=82◦. Positive magnetic pole points closest to us at the phase Φ ∼0.25 and negative pole at Φ ∼0.75. Brighter light
curve maximum takes place at Φ ∼ 0.0 and another maximum at Φ ∼ 0.5. Cyclotron emission arcs and the dipole axis are drawn with thick lines. The emission
arcs are off from the magnetic poles by ∆β1=-3◦ and ∆β2=+4◦ for the positive and negative pole, respectively. The centers of the 24◦ long emission arcs are
ahead of the poles in longitude by ∆λ1=5◦ and ∆λ2=1◦ . The accretion model geometry is shown in Fig. 8.
FIG. 8. -- Side-on schematic view of the accretion model, shown for i=38◦ and magnetic dipole angle β=82◦. The dipole is decentered by 0.1RW D towards the
upper rotational pole. Arrows give viewing directions at phases 0.25 and 0.75. For a larger inner radius of the disc the accreting field lines hit the WD surface
closer to the magnetic poles.
8
Piirola et al.
Newton data as the peak of the larger of the soft X-ray max-
ima, which corresponds to HJD 2 452 187.55904. With our
ephemeris this takes place at the phase 0.44, which is near
the second maximum in the optical light curves (phase 0.5),
where we look nearly perpendicular to the magnetic field.
This would be in accordance with the standard accretion cur-
tain model (see e.g. Hellier et al. 1991), where the opacity of
the column causes X-rays to emerge preferentially perpendic-
ular to the curtain, and argues against the other plausible ex-
planation that foreshortening of the accretion polecaps when
viewed face on and near the limb of the WD would be the
dominating effect on the soft X-ray flux observed in V405
Aur. However, neither de Martino et al. (2004) nor Evans
& Hellier (2004) found the absorption dip expected from the
accretion-curtain dip model, and the soft X-ray behaviour of
V405 Aur vs. the magnetic phase requires further attention.
3.4. Spectropolarimetry
If cyclotron harmonics are detected in the spectrum, they
provide an accurate method for determining the magnetic field
strength in the emission region. However, thermal broadening
effects and magnetic field spread in an inhomogeneous region
may smear out the harmonics and make them undetectable.
Also the overlying emission from other than the cyclotron
source need to be carefully modeled and removed (see e.g.
Cropper et al. 1988). Because of these complications, cy-
clotron harmonic humps have not been seen in many of the
known polars, and even if detected are visible only at some
phase angle intervals during the WD spin cycle.
In V405 Aur the bright thermal emission from the disk and
the accretion stream dilute the cyclotron flux at least by an
order of magnitude when compared with strongly polarized
polars which have Pc ∼ 20-40%. This makes the detection
of cyclotron humps in the intensity spectrum of V405 Aur
unlikely. Therefore, we have carried out circular spectropo-
larimetry with the aim to search for cyclotron harmonics in the
polarized flux spectra, where the effects from thermal emis-
sion sources are largely eliminated.
Most of the polarized spectra we obtained show no clear cy-
clotron harmonic pattern. However, at some phase angles of
the WD spin period we see transient structures which suggest
a possible cyclotron origin. A single 120 s exposure of the
spectrum of the negative emission region in Figure 9 shows
three peaks (downwards), the inverse wavelengths of which
can be fitted with cyclotron harmonics n = 6, 7, and 8, at
magnetic field of B = 31.5±0.8 MG. If confirmed, this would
make V405 Aur the first IP with a direct measurement of mag-
netic field, and the value found is typical of that of polars.
3.5. Is V405 Aur a progenitor of a polar?
The broad-band (UBV RI) circular polarization wavelength
dependence (Fig. 1) is similar to what is seen in moderately
high-field polars (B ∼ 25-50MG), such as VV Pup (Piirola
et al. 1990; Schwope & Beuermann 1997) and V834 Cen
(Piirola 1995; Ferrario et al. 1992). Our magnetic field es-
timates for V405 Aur from the polarized flux spectrum (B ∼
25 MG, Sect. 3.2) and spectropolarimetry (B ∼30 MG, Sect.
3.4) are in accordance with the values determined for those
polars from cyclotron humps or Zeeman spectroscopy. This
makes V405 Aur a likely candidate as a progenitor of a polar.
Using a numerical model of magnetic accretion, Norton et
al.
(2004) have predicted that IPs with magnetic moments
µ1 ≥ 5 × 1033 G cm3 and Porb > 3 hr will evolve into polars.
FIG. 9. -- Circular spectropolarimetry of V405 Aur on 2004 Feb 26. The
negative circular polarization spectrum (120 s exposure, top panel) shows
features which can be fitted on the 1/λ scale with cyclotron harmonics n = 6,
7, and 8 from a field of B = 31.5±0.8 MG. The degree of circular polarization
has been corrected for the diluting unpolarized flux from the emission lines
and smoothed with a ∆λ ∼ 70 Å wide window filtering. The bottom panel
gives a raw intensity spectrum (ADU).
For a solar-mass WD a surface field of B = 30 MG corre-
sponds to µ1 ∼ 5 × 1033 G cm3, which gives further support
that V405 Aur may evolve into a polar.
The ratio of the spin to the orbital period Pspin/Porb = 0.0365
indicates that V405 Aur is still far from synchronism. Our
spin period analysis (Fig. 3 and Eqs. 1-2) gives Pspin which is
constant within tight limits, and the corresponding spin-down
timescales are very long (Ts > 109 yr). Therefore, V405 Aur
is currently accreting closely at the spin equilibrium rate. For
small Pspin/Porb values this condition occurs in disk-like ac-
cretion (Norton et al. 2004, 2008). The angular momentum
gain from accretion onto the WD balances the magnetic brak-
ing torque, and equilibrium is reached.
In Figure 10 we show the location of V405 Aur in the Pspin
vs. magnetic moment, µ1, diagram from Norton et al. (2008)
for Porb = 4 hr, and three different mass ratios, q=0.2, 0.5, and
0.9. The inferred magnetic field strength depends strongly on
the value of q. For the value of Pspin/Porb = 0.0365 these nu-
merical model computations predict a field B = 20 MG for q =
0.5 and the spin equilibrium condition, defined in these graphs
by the boundary line between the D (disk) and P (propeller),
or the S (stream) and P (propeller) type accretion. Lower q
implies larger B (60 MG for q = 0.2). B=30 MG would be ob-
tained for q ∼ 0.4 from these models, assuming a solar mass
WD in the conversion of µ1 to the surface field.
Because of the angular momentum loss via gravitational
radiation and magnetic wind braking, V405 Aur will evolve
towards shorter orbital period and smaller binary separation,
a. It is expected that when small enough a is reached, the
magnetic braking torque will overcome the spin-up accretion
torque and the WD will start to spin-down towards synchro-
nization, the system thereby becoming a polar.
4. CONCLUSIONS
Our simultaneous multicolor (UBV RI) circular polarimetry
of the intermediate polar V405 Aur has revealed polarization
which peaks in the B and V passbands, similar to relatively
high field (B = 25 - 50 MG) polars. This is the first time that
a polarized flux spectrum reaching maximum at such short
wavelengths has been found in an IP, and suggests that V405
Aur is the highest field IP found so far. We have also detected
transient features in circularly polarized spectra which can be
V405 Aur: A High Magnetic Field IP
9
FIG. 10. -- The location of V405 Aur in the Pspin/Porb vs. magnetic moment µ1 diagram from Norton et al. (2008) for Porb = 4 hr and mass ratios q = 0.2, 0.5,
and 0.9. Spin equilibrium condition predicted by the numerical model is defined in these graphs by the boundary line between the D (disk) and P (propeller), or
the S (stream) and P (propeller) type accretion. The respective surface magnetic field values (MG) correspond to a solar-mass WD.
fitted on the wavenumber scale with cyclotron harmonics n=6,
7, and 8, at a magnetic field of B = 31.5 ± 0.8 MG. Such a
field is consistent with the broad-band polarized flux spectrum
observed, and this makes V405 Aur a likely candidate of a
progenitor of a polar.
Nearly sinusoidal and symmetric positive and negative cir-
cular polarization excursions in all of the UBV RI passbands
indicate that the two opposite magnetic poles are accreting
at very similar rate, and are located far from the rotational
poles, probably near the equator. For such a magnetic geom-
etry, our numerical model simulations suggest WD spin axis
inclination, i = 30-50◦. The broad-band circular polarization
spectrum is flatter than predicted by constant temperature cy-
clotron models, and requires an inhomogeneous emission re-
gion with a spread of the physical parameters (temperature,
electron density, magnetic field) and possibly a complicated
geometry.
Period analysis from timings of the zero-crossover of cir-
cular polarization puts strict constraints on the WD spin pe-
riod changes. The respective spin-down timescales are very
long, Ts > 109 yr, which means that the WD in V405Aur is
currently accreting very closely at the spin equilibrium rate.
With the spin to orbital period ratio, Pspin/Porb = 0.0365, such
a condition is achieved in disk-type accretion where the angu-
lar momentum gain from matter transferred onto the WD via
the disk balances the magnetic braking torque. Comparison
with numerical accretion model computations (Norton et al.
2008) shows that for V405 Aur equilibrium condition would
be obtained with mass ratio q1 ∼ 0.4, if the magnetic field is
B ∼ 30 MG.
The Nordic Optical Telescope is operated on the island
of La Palma jointly by Denmark, Finland, Iceland, Nor-
way, and Sweden, in the Spanish Observatorio del Roque
de los Muchachos (ORM) of the Instituto de Astrophysica
de Canarias. The KVA-60 telescope is operated by Tuorla
Observatory of the University of Turku, at ORM under the
agreement between the University of Turku, Finland, and the
Royal Academy of Sciences, Sweden (Kungliga Vetenskap-
sakademien).
REFERENCES
Allan, A., Horne, K., Hellier, C., et al. 1996, MNRAS, 279, 1345
Buckley, D.A., Haberl, F., Motch, C., et al. 1997, MNRAS, 287, 117
Cropper, M. 1990, Space Science Reviews, 54, 195
Cropper, M., Mason, K.O., Allington-Smith, J.R., et al. 1988, MNRAS, 236,
29P
de Martino, D., Matt, G., Belloni, T., et al. 2004, A&A, 415, 1009
Evans, P.A., & Hellier, C. 2004, MNRAS, 353, 447
Ferrario, L, Wickramasinghe, D.T., Bailey, J., et al. 1992, MNRAS, 256, 252
Ferrario, L., Wickramasinghe, D.T., Bailey, J., & Buckley, D. 1995,
MNRAS, 273, 17
Ferrario, L., Wickramasinghe, D.T., & King, A.R. 1993, MNRAS260, 149
Haberl, F., Thorstensen, J.R., Motch, C., et al. 1994, A&A, 291, 171
Harlaftis, E.T., & Horne, K. 1999, MNRAS, 305, 437
Hellier, C., Cropper, M., & Mason, K. 1991, MNRAS, 248, 233
Katajainen, S., Butters, O.W., Norton, A.J., et al. 2007, A&A, 475, 1011
Mason, K.O. 1997, MNRAS, 285, 493
Norton, A.J., Quaintrell, H., Katajainen, S., et al. 2002, A&A, 384, 195
Norton, A.J., Wynn, G.A., & Somerscales, R.V. 2004, ApJ, 614, 349
Norton, A.J., Butters, O.W., Parker, T.L., & Wynn, G.A. 2008, ApJ, 672, 524
Penning, W.R., Schmidt, G.D., & Liebert, J. 1986, ApJ, 301, 881
Piirola, V. 1988, in 'Polarized Radiation of Circumstellar Origin', eds. G.V.
Coyne et al. (Tucson: Univ. of Arizona Press), 735
Piirola, V. 1995, in ASP Conf. Ser. 85: Magnetic Cataclysmic Variables, ed.
D.A.H. Buckley & B. Warner, 31
Piirola, V., Berdyugin, A., Mikkola, V., & Coyne, G.V. 2005, ApJ, 632, 576
Piirola, V., Coyne, G.V., & Reiz, A. 1990, A&A, 235, 245
Piirola, V., Coyne, G.V., Takalo, L., et al. 1994, A&A, 283, 163
Piirola, V., Hakala, P., & Coyne, G.V. 1993, ApJ, 410, L107
Potter, S.B., Cropper, M., Mason, K.O., et al. 1997, MNRAS, 285, 82
Rosen, S.R., Mason, K.O., & Cordova, F.A., 1988, MNRAS, 231, 549
Rosen, S.R., Mittaz, J.P.D., & Hakala, P. 1993, MNRAS, 264, 171
Schwope, A.D., & Beuermann, K. 1997, Astron. Nachr., 318, 111
Schmidt, G.D., Szkody, P., Smith, P.S., et al. 1996, ApJ, 473, 483
Shakhovskoy, N.M., & Kolesnikov, S.V. 1997, IAU Circ. 6760, 2
Still, M., Duck, S.R., & Marsh, T.R. 1998, MNRAS, 299, 759
Uslenghi, M., Tommasi, L., Treves, A., Piirola, V., & Reig, P. 2001, A&A,
Väth, H. 1997, A&A, 317, 476
Väth, H., Chanmugan, G., & Frank, J. 1996, ApJ, 457, 407
Warner, B. 1995, Cataclysmic Variable Stars (Cambridge University Press),
372, L1
572
West, S.C., Berriman, G., & Schmidt, G.D. 1987, ApJ, 322, L35
Whittet, D.C.B., Martin, B.G., Hough, J.H., et al. 1992, ApJ, 386, 562
Wickramasinghe, D.T. & Meggitt, S.M.A. 1985, MNRAS, 214, 605
Wickramasinghe, D.T., Wu, K., & Ferrario, L. 1991, MNRAS, 249, 460
|
astro-ph/0403619 | 1 | 0403 | 2004-03-26T17:45:52 | Gravitational microlensing and dark matter problem in our Galaxy: 10 years later | [
"astro-ph"
] | Foundations of standard theory of microlensing are described, namely we consider microlensing stars in Galactic bulge, the Magellanic Clouds or other nearby galaxies. We suppose that gravitational microlenses lie between an Earth observer and these stars. Criteria of an identification of microlensing events are discussed. We also consider such microlensing events which do not satisfy these criteria (non-symmetrical light curves, chromatic effects, polarization effects). We describe results of MACHO collaboration observations towards the Large Magellanic Cloud (LMC) and the Galactic bulge. Results of EROS observations towards the LMC and OGLE observations towards the Galactic bulge are also presented. Future microlensing searches are discussed. | astro-ph | astro-ph |
GRAVITATIONAL MICROLENSING AND DARK MATTER IN
OUR GALAXY: 10 YEARS LATER
Institute of Theoretical and Experimental Physics, 117259, B. Cheremushkinskaya,
A.F. ZAKHAROVa
25, Moscow, Russia
Astro Space Centre of Lebedev Physics Institute, Moscow, Russia
Abstract. Foundations of standard theory of microlensing are described, namely
we consider microlensing stars in Galactic bulge, the Magellanic Clouds or other
nearby galaxies. We suppose that gravitational microlenses lie between an Earth
observer and these stars. Criteria of an identification of microlensing events are
discussed. We also consider such microlensing events which do not satisfy these
criteria (non-symmetrical light curves, chromatic effects, polarization effects). We
describe results of MACHO collaboration observations towards the Large Magel-
lanic Cloud (LMC) and the Galactic bulge. Results of EROS observations towards
the LMC and OGLE observations towards the Galactic bulge are also presented.
Future microlensing searches are discussed.
A standard microlens model is based on a simple approximation of a point
mass for a gravitational microlens. Gravitational lensing (gravitational fo-
cusing) results from the effect of light bending by a gravitating body (the
phenomenon was discussed by I. Newton, but in the framework of Newto-
nian gravity a formal derivation of the light bending angle was published by
J. Soldner [1]).
In the framework of general relativity (GR) using a weak gravitational field
approximation the correct bending angle is described by the following expres-
sion derived by Einstein in 1915 just after his formulation of GR
δϕ = −
4GM∗
c2p
.
(1)
The derivation of the famous Einstein's formulae for the bending angle of light
rays in gravitational field of a point mass M∗ is practically in all monographs
and textbooks on general relativity and gravity theory (see, for example books
[2, 3]).
The law was firstly confirmed by Sir A. Eddington for observations of light
ray bend by the Solar gravitational field near its surface. The angle is equal to
1.75′′, therefore Einstein prediction was confirmed by observations very soon
after its discovery.
The gravitational lens effect is a formation of several images instead of one
(see details in [4,5]). We have two images for a point lens model (Schwarzschild
lens model). The total square of the two images is larger than a source square.
The ratio of these two squares is called gravitational lens amplification A. That
is a reason to call gravitational lensing as gravitational focusing. The angular
distance between two images is about angular size of so-called Einstein's cone.
ae-mail: [email protected]
The angular size of Einstein's cone is proportional to the lens mass divided by
the distance between a lens and an observer. Therefore, if we consider a gravi-
tational lens with typical galactic mass and a typical galactic distance between
a gravitational lens and an observer then the angular distance between images
will be about few angular seconds; if we suppose that a gravitational lens has
a solar mass and a distance between the lens and an observer is about sev-
eral kiloparsecs then an angular distance between images will be about angular
millisecond.
If a separation angle is ∼ 1′′, then one may observe two images in optical band
although this problem is a complex one, but one cannot observe directly two
images by Earth's observer in the optical band if a separation angle is ∼ 0.001′′.
Therefore, the microlensing effect is observed on changing of a luminosity of a
source S.b
If the source S lies on the boundary of the Einstein cone, then we have A =
1.34. Note, that the total time of crossing the Einstein cone is T0. Sometimes
the microlensing time is defined as a half of T0 we suppose that Dd < Dds (here
we assume that Dds is the distance from the source S to the lens D; Dd is the
distance from the lens D to the observer O; Ds is the distance from the source
S to the observer O)
T0 = 3.5 months ·s M
M⊙
Dd
10 kpc
·
300 km/s
v
,
where v is the perpendicular component of a velocity of a dark body. If we
suppose that the perpendicular component of a velocity of a dark body is equal
to ∼ 300 km/s (that is a typical stellar velocity in Galaxy), then a typical time
of crossing Einstein cone is about 3.5 months. Thus, a luminosity of a source
S is changed with the time.
We will give numerical estimations for parameters of the microlensing effect.
If the distance between a dark body and the Sun is equal to ∼ 10 kpc, then
the angular size of Einstein cone of the dark body with a solar mass is equal to
∼ 0.001′′ or the linear size of Einstein cone is equal to about 10 astronomical
units.
It is clear that since typical distances between two images are about
Einstein diameters therefore is very difficult to resolve the images by ground
based telescopes at least in an optical band. It was a reason that both Einstein
and Chwolson thought if gravitational lenses and sources are stars then sepa-
ration angle is very small to be detectable. However, recently, a direct method
to measure Einstein angle φE was proposed to resolve double images generated
by microlensing with an optical interferometer (say VLTI) [6](see also [7] for a
bSince the angle is very small Einstein and Chwolson thought that gravitational lens effect
could not be detectable if sources and lenses are stars. Now there are chances to measure
such angles in IR band therefore there is a giant development of observational facilities.
discussion). Moreover, it is plan to launch astrometrical space probe, Ameri-
can SIMc and European GAIAd, these instruments will have precisions about
10 micro arc seconds and could determine Einstein radii for any microlensing
events.
Astrometric microlensing or motions of visible images due to influence of
a gravitational field of microlenses was analyzed in number of papers [8 -- 19],
although light bending in gravitational field was discussed by I. Newton (actu-
ally that is the same effect but authors presented detailed analysis and pushed
new ideas to use the phenomenon to detect even invisible astronomical objects
by shifts of images for background sources). An optical depth of microlensing
for distant quasars was discussed for different locations of microlenses (see, for
example, [20] and references therein.
For observations of extragalactic gravitational lens a typical time for changes
of light curve is very long (∼ 105 years) for its direct observations. Therefore,
extragalactic gravitational lenses are discovered and observed by resolving dif-
ferent optical components (images) since typical angular distances between
images are about some angular seconds because of a great mass of a gravita-
tional lens. If a gravitational lens is a galaxy cluster then the angular distances
between images may be about several minutes. For an identification of gravita-
tional lenses, observers compare typical features and spectra of different images.
It is clear that one cannot to resolve different components during microlensing
but it is possible to get and analyze a light curve in different spectral bands.
One of the basic criterion for microlensing event identification is the symme-
try of a light curve. If we consider a spherically symmetric gravitational field
of a lens, a point source and a short duration of microlensing event then the
statement about the symmetry of a light curve will be a strong mathematical
conclusion, but if we consider a more complicated distribution of a gravitational
field lens or an extensive light source then some deviations of symmetric light
curves may be observed and (or) the microlensing effect may be chromatic [4,5].
More than 70 years ago it was found that densities of visible matter is about
10% of total density in galactic halos (the invisible is called as dark matter
(DM) [21, 22])e Thus baryonic density is a small fraction of total density of the
Universe. Probably galactic halos is "natural" places to store not only baryonic
DM, but non-baryonic DM also. If DM forms objects with masses in the range
[10−5, 10]M⊙ microlensing could help to detect such objects. Thus, before
intensive microlensing searches it was a dream that microlensing investigations
could help us to solve DM problem for Galactic halo at least.
For the first time a possibility to discover microlensing using observations of
chttp://sim.jpl.nasa.gov/whatis/
dhttp://astro.estec.esa.nl/GAIA
eNow it is known that the matter density (in critical density units) is Ωm = 0.3 (including
baryonic matter Ωb ≈ 0.05 − 0.04, but luminous matter Ωlum ≈ 0.001), Λ-term density
ΩΛ = 0.7.
star light curves was discussed in the paper by Byalko in 1969 [23]. Systematic
searches of dark matter using typical variations of light curves of individual
stars from millions observable stars started after Paczynski's discussion of the
halo dark matter discovery using monitoring stars from Large Magellanic Cloud
(LMC) [24]. We remark that in the beginning of the nineties new computer
and technical possibilities providing the storage and processing of huge volume
of observational data were appeared and it promoted at the rapid realization
of Paczynski's proposal. Griest suggested to call the microlenses as Machos
(Massive Astrophysical Compact Halo Objects) [25]. Besides, MACHO is the
name of the project of observations of the US-English-Australian collaboration
which observed the LMC and Galactic bulge using 1.3 m telescope of Mount
Stromlo observatory in Australia.f
The first papers about the microlensing discovery were published by the
MACHO collaboration [26] and the French collaboration EROS (Exp´erience
de Recherche d'Objets Sombres) [27].g
First papers about the microlensing discovery toward Galactic bulge were
published by US-Polish collaboration (Optical Gravitational Lens Experiment),
which used 1.3 m telescope at Las Campanas Observatory. Since June 2001,
after second major hardware upgrade OGLE entered into its third phase, OGLE
III as a result the collaboration observes more than 200 millions stars observed
regularly once every 1 -- 3 nights. Last two years OGLE III detected more than
four hundreds microlensing event candidates each year [29].h
MOA (Microlensing Observations in Astrophysics) is collaboration involving
astronomers from Japan and New Zealand [30, 31].i
To investigate Macho distribution in another direction one could use searches
toward M31 (Andromeda) Galaxy lying at 725 kpc (it is the closest galaxy
for an observer in the Northern hemisphere).
In nineties two collaborations
AGAPE (Andromeda Gravitational Amplification Pixel Experiment, Pic du
Midi, France)j and VATT started to monitor pixels instead of individual stars
[28, 33]. These teams reported about discoveries of several microlensing event
candidates.
The event corresponding to microlensing may be characterized by the fol-
lowing main features, which allow to distinguish the microlensing event and a
stellar variability [4, 34, 35].
• Since the microlensing events have a very small probability, the events
f MACHO stopped since end 1999.
gEROS experiment stopped in 2002 [28].
hhttp://www.astrouw.edu.pl/ ogle/ogle3/ews/ews/html
ihttp://www/roe.ac.uk/%7Eiab/alert/alert/alert/html
jNew collaboration, POINT-AGAPE started in 1999 and uses INT (2.5 v Isaac Newton
Telescope) [54].
should never repeat for the same star. The stellar variability is connected
usually with periodic (or quasi-periodic) events of the fixed star.
• In the framework of a simple model of microlensing when a point source is
considered, the microlensing effect must be achromatic (deviations from
achromaticity for non-point source were considered, for example in the
paper by Bogdanov & Cherepashchuk [36]), but the proper change of
luminosity star is connected usually with the temperature changes and
thus the light curve depends on a colour.
• The light curves of microlensing events are symmetric, but the light curves
of variable stars are usually asymmetric (often they demonstrate the rapid
growth before the peak and the slow decrease after the peak of a lumi-
nosity).
• Observations of microlensing events are interpreted quite well by the sim-
ple theoretical model, but some microlensing events are interpreted by
more complicated model in which one can take into account that a source
(or a microlens) is a binary system, a source has non-vanishing size, the
parallax effect may take place.
The typical features of the light curve of the first microlensing event observed
by the MACHO collaboration in the LMC are shown in Fig. 1, where the light
curves are shown for two spectral bands (a more recent MACHO fit to the
observed amplification of this event gives Amax = 7.2). The light curve (in two
bands) is fitted by the simple model well enough, but the ratio of luminosities
for the bands is shown in the lower panel of figure (the ratio shape is adjusted
with the event achromaticity). However, one can note that near the maximal
observable luminosity the theoretical curve fits the data of observations not
very well.
Now one can carry out accurate testing the achromaticity and moreover
the stability of the source spectrum during a microlensing event with the Early
Warning systems implemented both by the MACHO and OGLE collaborations.
This allows one to study the source properties using large telescopes and to
organize intense follow-up studies of light curves using telescope network around
the globe.
In addition to the typical properties of individual microlensing events, Roulet
and Mollerach note that the population of observed events should have the
following statistical properties [4, 34].
• Unlike a star variability microlensing events should happen with the same
probability for any kind of star therefore the distribution of microlensing
Figure 1: The first microlensing event which was detected by the MACHO collaboration
during microlensing searches towards LMC [26].
events should correspond to the distribution of observed stars in the color-
magnitude diagrams.k
• The distribution of the maximal amplification factor Amax should cor-
respond to a uniform distribution of the variable umin = 1/b (b is the
dimensionless impact parameter).
• The distributions of the amplification Amax and the microlensing event
time T should be uncorrelated.
Since for the microlens searches one can monitor several million stars for
several years, the ongoing searches have focused on two targets: a) stars in the
Large and Small Magellanic Clouds (LMC and SMC) which are the nearest
galaxies having lines of sight which go out of the Galactic plane and well across
kHowever, Roulet and Mollerach noted that for observations in the bulge since observed
stars have non-negligible spread along the line of sight, the optical depth is significantly
larger for the star lying behind the bulge, thus the lensing probabilities should increase for
the fainter stars [34].
the halo; b) stars in the Galactic bulge which allow to test the distribution of
lenses near to the Galactic plane.l
Let us cite well established results of microlensing searches and discuss the
questions for which we have now different answers which do not contradict to
the observational data. Now it is generally recognized that the microlensing
searches towards the Galactic bulge or nearby galaxies are very important for
solutions of a lot of problems in astronomy and cosmology. As Paczynski noted,
the most important is the consensus that the microlensing phenomenon has
been discovered [37]. Now it is impossible to tell which part of the microlensing
event candidates is actually connected with the effect since probably there are
some variable stars among the event candidates, it could be stellar variability
of an unknown kind.m
1. Observed light curves are achromatic and their shapes are interpreted
by simple theoretical expressions very well, however, there is not com-
plete consent about "very well interpretation" since even for the event
candidate MACHO # 1 the authors of the discovery proposed two fits.
Dominik and Hirshfeld suggested that the event could be fitted perfectly
in the framework of the binary lens model [38, 39], but Gurevich et al.
assumed that the microlensing event candidate could be caused by a non-
compact microlens [40].n
2. As expected, binary lenses have been detected and the observed rate of
the events correspond to expected value.
3. As expected, the parallax effect has been detected.
4. Since the observed optical depth is essentially greater than the estimated
value, the independent confirmation of the Galactic bar existence was
done.
5. Using photometric observations of the caustic-crossing binary lens mi-
crolensing event EROS BLG-2000-5, PLANET collaboration reported
about the first microlens mass determination, namely the masses of these
components are 0.35 M⊙ and 0.262 M⊙ and the lens lies within 2.6 kpc
of the Sun [47].
6. Bennett et al. discovered gravitational microlensing events due to stel-
lar mass black holes [48]. The lenses for events MACHO-96-BLG-5 and
MACHO-96-BLG-6 are the most massive, with mass estimates M/M⊙ =
6+10
−3 and M/M⊙ = 6+7
−3, respectively.
lIn this paper we do not discuss microlensing for distant quasars.
mThe microlensing event candidates proposed early by the EROS collaboration ( #1 and
#2) and by the MACHO collaboration (#2 and #3) are considered now as the evidence of
a stellar variability [37].
nMicrolensing by non-compact objects considered also in papers [41 -- 46].
Now the following results are generally accepted:
1. The optical depth towards the Galactic bulge is equal to ∼ 3 × 10−6, so
it is larger than the estimated value [49].
2. Analysis of 5.7 years of photometry on 11.9 million stars in LMC by MA-
CHO collaboration reveals 13 -- 17 microlensing events [50] (recent results
of the MACHO collaboration on could find in [51]). The optical depth
towards the LMC is equal to τ (2 < t < 400 days) = 1.2+0.4
× 10−7, so,
−0.3
it is smaller than the estimated value. The maximum likelihood analysis
gives a MACHO halo fraction f=0.2. Alcock et al. (2000b) gives also
estimates of the following probabilities P (0.08 < f < 0.5) = 0.95 and
P (f = 1) < 0.05. The most likely MACHO mass M ∈ [0.15, 0.9]M⊙,
depending on the halo model and total mass in MACHOs out 50 kpc is
found to be 9+4
× 1010M⊙ EROS collaboration gives a consistent con-
clusion, namely, this group estimates the following probability P (M ∈
[10−7, 1]M⊙ & f > 0.4) < 0.05 [52, 53]. However, these conclusions are
based on assumptions about mass and spacial distributions of microlenses
but generally speaking these distributions are still unknown.
−3
However there are different suggestions (which are not contradicted to the
observational data) about the following issues [37]:
What is the location of objects which dominate microlensing observed towards
the Galactic bulge?
Where are the most microlenses for searches towards LMC? The microlenses
may be in the Galactic disk, Galactic halo, the LMC halo or in the LMC itself.
Are the microlenses stellar mass objects or are they substellar brown dwarfs?
What fraction of microlensing events is caused by binary lenses?
What fraction of microlensing events is connected with binary sources?
Paczynski suggested that we shall have definite answers for some presented
issues after some years and since the optical depth towards the Galactic bulge
is essentially greater than the optical depth towards the LMC, we shall have
more information about the lens distribution towards the Galactic bulge, how-
ever, probably, some problems in theoretical interpretation will appear after
detections of new microlensing event candidates [37].
The main result of the microlensing searches is that the effect predicted the-
oretically has been confirmed. This is one of the most important astronomical
discoveries.
When new observational data would be collected and the processing methods
would be perfected, probably some microlensing event candidates lost their sta-
tus, but perhaps new microlensing event candidates would be extracted among
analyzed observational data. So, the general conclusion may be done. The
very important astronomical phenomenon was discovered, but some quanti-
tative parameters of microlensing will be specified in future. However, the
problem about 80% of DM in the halo of our Galaxy is still open (10 years
ago people believe that microlensing could give an answer for this problem).
Thus, describing the present status Kerins wrote adequately that now we have
"Machos and clouds of uncertainty" [54].
I thank prof. A.I. Studenikin for his kind invitation to present this contri-
bution at the XI Lomonosov Conference on Elementary Particle Physics.
References
[1] J.G. Soldner, Berliner Astron. Jahrbuch, 1804, 161 (1804).
[2] L.D. Landau, E. M. Lifshitz, 1975, The Classical Theory of Fields, Perg-
amon Press, Oxford (1975).
[3] C. Moller, The Theory of Relativity Oxford, Clarendon Press, 1972.
[4] A.F. Zakharov, Gravitatsionnie linzi i microlinzi (Gravitational Lenses
and Microlenses) Janus-K, Moscow (1997).
[5] A. F. Zakharov, M.V. Sazhin, Phys. Usp., 41, 945 (1998).
[6] F. Delplancke, K. Gorski, A. Richichi, Astron. and Astrophys., 375, 701
(2001).
[7] B. Paczynski, Preprint astro-ph/0306564 (2003).
[8] E. Hog, I.D.Novikov, A.G.Polnarev, Astron. and Astrophys., 294, 287
(1995).
[9] M.A. Walker, Astrophys. J., 453, 37 (1995).
[10] M. Miyamoto, Y.Yoshii, Astron. J., 110, 1427 (1995).
[11] M.V. Sazhin, Pis'ma v Astron. Zhurn., 22, 647 (1996).
[12] M.V. Sazhin et al., Montly Notices Roy. Astron. Soc., 300, 287 (1998).
[13] A.F. Boden, M. Shao, D. Van Buren, Astrophys. J., 502, 538 (1998).
[14] B. Paczynski, Astrophys. J., 494, L23 (1998).
[15] M. Honma, Publ. Astron. Soc. Japan, 53, 223 (2001).
[16] M. Honma, T. Kurayama, Astrophys. J., 568, 717 (2002).
[17] R. Takahashi, Astrophys. J., 595, 418 (2003).
[18] G.F. Lewis, R.A. Ibata, Astrophys. J., 501, 478 (1998).
[19] M. Treyer, J. Wambsganss, Preprint astro-ph/0311519.
[20] A.F. Zakharov, L. C. Popovi´c, P. Jovanovi´c, Astron. & Astrophys. (ac-
cepted); astro-ph/0403254 (2004).
[21] J. Oort, Bull. Astron. Instit. Heth., 6, 249 (1932).
[22] F. Zwicky, Helvetica Physica Acta, 11, 110 (1933).
[23] A.V. Byalko, Astron. Zhurn., 46, 998 (1969).
[24] B. Paczynski, Astrophys. J., 304, 1 (1986).
[25] K. Griest, Astrophys. J., 366, 412 (1991).
[26] C. Alcock et al., Nat., 365, 621 (1993).
[27] E. Aubourg et al., Nat., 365, 623 (1993).
[28] M. Moniez, Cosmological Physics with Gravitational Lensing, eds.
J. Tran Thanh Van, Y. Mellier & M. Moniez, Proc. of the XXXVth
Rencontres de Moriond, EDP Sciences, 3, 2001.
[29] A. Udalski, Acta Astron., 51, 175 (2002).
[30] I.A. Bond et al., MNRAS, 327, 868 (2001).
[31] J. Skuljan, Publ. Astron. Obs. Belgrade, 75, 37 (2003)
[32] E. Kerins et al., Preprint astro-ph/0002256 (2000).
[33] Y. Le Du, Cosmological Physics with Gravitational Lensing, eds. J. Tran
Thanh Van, Y. Mellier & M. Moniez, Proc. of the XXXVth Rencontres
de Moriond, EDP Sciences, 65, 2001.
[34] E. Roulet, S. Mollerach, Phys. Rep. 279, 2 (1997).
[35] A.F. Zakharov, Publ. Astron. Obs. Belgrade 75, 27 (2003); astro-
ph/0212009.
[36] M.B. Bogdanov, A.M. Cherepashchuk, Pis'ma v Astron. Zhurn., 21, 570
(1995).
[37] B. Paczynski, Ann. Rev. Astron & Astrophys., 34, 419 (1996).
[38] M. Dominik, A.C. Hirshfeld, Astron. and Astrophys., 289, L31 (1994).
[39] M. Dominik, A.C. Hirshfeld, Preprint DO-TH 95/19, Dortmund (1995).
[40] A.V. Gurevich, K.P. Zybin, V.A. Sirota, Phys. Lett. A 214, 232 (1996).
[41] A.F. Zakharov, Phys. Lett. A, 250, 67 (1998).
[42] A.F. Zakharov, Astron. Rep., 43, 325 (1999).
[43] A.F. Zakharov, Dark Matter in Astro- and Particle Physics, ed.
H.V. Klapdor-Kleingrothaus, Proc. of the Intern. Conf. DARK-2000,
Springer, 364 (2001).
[44] A.F. Zakharov, Cosmological Physics with Gravitational Lensing, eds.
J. Tran Thanh Van, Y. Mellier & M. Moniez, Proc. of the XXXVth
Rencontres de Moriond, EDP Sciences, 57 (2001).
[45] A.F. Zakharov, M. V. Sazhin, JETP Letters, 63, 937 (1996).
[46] A.F. Zakharov, M. V. Sazhin, JETP 83, 1057 (1996).
[47] J.H. An et al., 2002, Astroph. J., 572, 521 (2002)
[48] D.C. Bennett et al., Astroph. J., 579, 639 (2002).
[49] C. Alcock et al., Astroph. J., 541, 734 (2000).
[50] C. Alcock et al., Astroph. J., 542, 281 (2000)
[51] P. Popowski al., Preprint astro-ph/0304464 (2003).
[52] T. Lasserre et al., 2000, Astron & Astroph. , 355, L39 (2000).
[53] T. Lasserre, 2001, Dark Matter in Astro- and Particle Physics with Grav-
itational Lensing, ed. H.V. Klapdor-Kleingrothaus, Proc. of the Intern.
Conf. DARK-2000, Springer, 342, 2001.
[54] E. Kerins, Cosmological Physics with Gravitational Lensing, eds. J. Tran
Thanh Van, Y. Mellier & M. Moniez, Proc. of the XXXVth Rencontres
de Moriond, EDP Sciences, 43, 2001.
|
astro-ph/9907139 | 1 | 9907 | 1999-07-12T11:00:09 | Proposed identification of Hubble Deep Field submillimeter source HDF 850.1 | [
"astro-ph"
] | The IRAM interferometer has been used to detect the submm source HDF 850.1 found by Hughes et al. (1998) in the Hubble Deep Field. The flux density measured at 1.3mm is 2.2 mJy, in agreement with the flux density measured at the JCMT. The flux densities and upper limits measured at 3.4, 2.8, 1.3, 0.85, and 0.45 mm show that the emission is from dust. We suggest that the 1.3mm dust source is associated with the optical arc-like feature, 3-593.0, that has a photometric redshift of about 1.7. If HDF 850.1 is at this redshift and unlensed, its spectral energy distribution, combined with that of 3-593.0, matches closely that of the ultraluminous galaxy VII Zw 31. Another possibility is that the dust source may be gravitationally lensed by the elliptical galaxy 3-586.0 at a redshift of 1. | astro-ph | astro-ph | A&A manuscript no.
(will be inserted by hand later)
Your thesaurus codes are:
11 (11.09.1 HDF 850.1, 11.09.4, 11.11.1, 11.19.6)
ASTRONOMY
AND
ASTROPHYSICS
13.1.2018
9
9
9
1
l
u
J
2
1
1
v
9
3
1
7
0
9
9
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
Proposed identification of Hubble Deep Field submillimeter
source HDF 850.1
D. Downes1, R. Neri1, A. Greve1, S. Guilloteau1, F. Casoli2, D. Hughes3,4, D. Lutz5, K.M. Menten6,
D.J. Wilner7, P. Andreani8, F. Bertoldi6, C.L. Carilli9, J. Dunlop3, R. Genzel5, F. Gueth6, R.J. Ivison10,
R.G. Mann11, Y. Mellier2,12, S. Oliver11, J. Peacock3, D. Rigopoulou5, M. Rowan-Robinson11,
P. Schilke6, S. Serjeant11, L.J. Tacconi5, and M. Wright13
1 Institut de Radio Astronomie Millim´etrique, Domaine Universitaire, F-38406 St. Martin d'H`eres, France
2 DEMIRM, Observatoire de Paris, 61 av. de l'Observatoire, F-75014 Paris, France, and UMR 8540 du CNRS
3 Institute for Astronomy, University of Edinburgh, Royal Observatory, Blackford Hill, Edinburgh, EH9 3HJ, UK
4 Instituto Nacional de Astrofisica, Optica y Electronica (INAOE), Apartado Postal 51 y 216, 72000 Puebla, Pue., Mexico
5 Max-Planck-Institut fur extraterrestrische Physik, D-85748 Garching-bei-Munchen, Germany
6 Max-Planck-Institut fur Radioastronomie, Auf dem Hugel 69, D-53121 Bonn, Germany
7 Center for Astrophysics, 60 Garden St., Cambridge, MA 02138, USA
8 Dipartimento di Astronomia, Universit`a di Padova, vicolo dell'Osservatorio 5, I-35122 Padova, Italy
9 National Radio Astronomy Observatory, P.O. Box O, Socorro, N.M., 87801, USA
10 Dept. of Physics and Astronomy, University College London, Gower Street, London, WC1E 6BT, UK
11 Astrophysics Group, Imperial College London, Blackett Laboratory, Prince Consort Road, London SW7 2BZ, UK
12 Institut d'Astrophysique, 98bis, Bd Arago, 75014, Paris, France
13 Radio Astronomy Laboratory, University of California, Berkeley, CA94720, USA
received date; accepted date
Abstract. The IRAM Interferometer has been used
to detect the submillimeter source HDF 850.1 found
by Hughes et al. (1998) in the Hubble Deep Field.
The flux density measured at 1.3 mm (236 GHz) is
2.2±0.3(1σ) mJy, in agreement with the flux density mea-
sured at the JCMT. The flux densities and upper limits
measured at 3.4, 2.8, 1.3, 0.85, and 0.45 mm show that the
emission comes from dust. We suggest that the 1.3 mm
dust source is associated with the optical arc-like feature,
3-593.0, that has a photometric redshift z ≈ 1.7. If the
HDF 850.1 is at this redshift and unlensed, its spectral en-
ergy distribution, combined with that of 3-593.0, matches
closely that of the ultraluminous galaxy VII Zw 31. An-
other possibility is that the dust source may be gravita-
tionally lensed by the elliptical galaxy 3-586.0 at z ≈ 1.
The position of the dust source agrees within the errors
with that of the tentative VLA radio source 3651+1226.
Key words: galaxies: structure -- galaxies: individual
(HDF-850.1) -- cosmology -- galaxies: ISM
Send offprint requests to: D. Downes
1. Introduction
Hughes et al. (1998) observed the Hubble Deep Field with
the SCUBA multi-beam bolometer array at the James
Clerk Maxwell Telescope (JCMT) at 850 and 450 µm.
Apart from HDF 850.4 (and possibly HDF 850.3) the
SCUBA survey did not detect the mid-IR sources found
in the Hubble Deep Field with the ISO satellite (Rowan-
Robinson et al. 1997; Mann et al. 1997; Aussel et al. 1999).
In general these ISOCAM detections are associated with
z < 1 galaxies, while the submillimeter survey appears
to have revealed a higher-redshift population of dust en-
shrouded starburst galaxies. Further SCUBA surveys of
other fields (e.g., Barger et al. 1999a,b; Eales et al. 1999;
Lilly et al. 1999) indicate however that many of the sub-
mm sources have redshifts smaller than 1, and that much
of the star-forming activity in galaxies has occurred rela-
tively recently, at z ∼ 2.
Richards (1999) argued against high redshifts (z > 2)
for the HDF sub-mm sources, from plausible associations
of the SCUBA sources with nearby radio sources detected
in his VLA survey at 20 cm (1.4 GHz) and the identifica-
tion of some of these non-thermal radio continuum sources
with optical galaxies in the Hubble Deep Field. From the
radio luminosity function of starburst galaxies, Richards
argued that if the submm dust sources had cm-radio detec-
2
D. Downes et al.: Proposed identification of Hubble Deep Field submillimeter source HDF 850.1
Table 1. Source Positions
Source
R.A.
(J2000)
Dec.
(J2000)
∆θ from
HDF850.1
Proposed identification:
HDF-850.1
3-593.0
3-593.2
3651+1226
12h36m51.98s ± 0.04s
12h36m51.86s ± 0.06s
12h36m51.94s ± 0.06s
12h36m51.96s ± 0.14s
62◦12′25.7′′ ± 0.3′′
62◦12′25.6′′ ± 0.4′′
62◦12′25.0′′ ± 0.4′′
62◦12′26.1′′ ± 1.0′′
Other nearby optical sources within 3′′:
3-586.0
3-577.0
3-613.0
12h36m52.10s ± 0.06s
12h36m52.26s ± 0.06s
12h36m52.37s ± 0.06s
62◦12′26.3′′ ± 0.4′′
62◦12′27.2′′ ± 0.4′′
62◦12′25.1′′ ± 0.4′′
Other nearby sources within 6′′:
3-550.2
3-633.1
3-659.1
3651+1221
PM3 29
tentative?
12h36m51.34s ± 0.06s
12h36m51.60s ± 0.06s
12h36m51.72s ± 0.06s
12h36m51.65s ± 0.02s
12h36m51.9s ± 0.4s
12h36m51.58s ± 0.06s
62◦12′26.9′′ ± 0.4′′
62◦12′22.5′′ ± 0.4′′
62◦12′20.2′′ ± 0.4′′
62◦12′21.4′′ ± 0.2′′
62◦12′21′′ ± 3′′
62◦12′20.5′′ ± 0.4′′
--
0.8′′
0.8′′
0.4′′
1.0′′
2.5′′
2.8′′
4.6′′
4.2′′
5.8′′
4.9′′
4.7′′
5.9′′
Adopted positions for phase calibrators:
1044+719
1125+596
1300+580
10h48m27.620s ± 0.002s
11h28m13.342s ± 0.002s
13h02m52.465s ± 0.002s
71◦43′35.93′′ ± 0.01′′
59◦25′14.78′′ ± 0.01′′
57◦48′37.62′′ ± 0.01′′
14.0◦
8.8◦
5.5◦
Redshift
Remark
Refs.
z
--
1.73 to 1.76
1.82
--
1.3 mm dust source
"arc 1"
"arc 2"
VLA (3 cm) ?
1.0 to 1.2
2.88, 2.89, 3.36 ?
1.64
elliptical
"counterimage"
blue irregular
1.72
1.72
0.299
--
--
--
Flux:
--
--
--
blue irregular
irregular ?
Sb spiral (Cowie 1999)
VLA (20 cm, 3 cm)
ISOCAM (15 µm)
3σ contour at 1.3 mm
1.3 mm
0.39 Jy
0.12 Jy
0.13 Jy
3.4 mm
0.68 Jy
0.20 Jy
0.20 Jy
1
2, 6, 7
2, 6
3
2, 6, 7
6, 7, 8
2, 6, 7
2, 6, 7
2, 6, 7
2, 6
3
4
1
5
5
5
Position refs.: (1) This paper; (2) Williams et al. (1996); (3) Richards et al. (1998); (4) Aussel et al. (1999);
(5) Patnaik et al. (1992). Redshift refs.: (6) Rowan-Robinson (1999); (7) Fern´andez-Soto et al. (1999); (8) Zepf et al. (1997)
tions, then they must be either z < 2 star-forming galaxies
or, if at higher redshift, heated by active galactic nuclei
(AGN). To clarify these issues, it is important to obtain
good positions for the dust sources by millimeter interfer-
ometry to examine possible coincidences with the Hubble
Deep Field optical galaxies and VLA radio sources.
The strongest object in the SCUBA survey is the
source HDF 850.1, which has a flux density of 7.0±0.4 mJy
at 850 µm. Follow-up photometry of HDF 850.1 by Hughes
et al. (1998) yielded a flux density of 2.1 ± 0.5 mJy at
a wavelength of 1.35 mm, confirming that the continuum
is optically thin emission by dust. At longer millime-
ter wavelengths, the Hubble Deep Field is known to be
empty of sources. With the BIMA interferometer, Wilner
& Wright (1997) found no sources at 2.8 mm, to a 5σ-
limit of 3.5 mJy. Because the 1.3 mm flux density of HDF
850.1 is close to the minimum detectable level in a 10-hr
integration at this wavelength, this strongest source in the
SCUBA survey may be the only source in the Hubble Deep
Field that can be detected at present in reasonable ob-
serving times with millimeter interferometers. To measure
more accurately the position of this source, we observed
it at 1.3 mm and 3 mm with the IRAM Interferometer on
Plateau de Bure, France.
2. Observations
The observations were made at 1.3 mm and 3 mm simulta-
neously, in the interferometer's compact configuration D
on November 17 and 22, 1998, and in the more extended
configuration C2 on December 15, 16, and 17, 1998. The
total integration time was about 40 hours, in excellent
weather, with precipitable water vapor content ≈ 1.5 mm,
and r.m.s. phase errors at 1.3 mm of ≤ 30◦. The five 15 m
dishes give 10 interferometer baselines from 24 m to 80 m
in configuration D, and 24 to 183 m in configuration C2.
At 1.3 mm, the observing frequency was 236.3 GHz, and
data were taken in upper and lower sidebands separated by
3 GHz. At 3 mm, data were taken in lower sideband only,
at 2.8 mm (105.7 GHz) on November 17, and at 3.4 mm
(88.7 GHz) on all the other dates. The SIS receivers had
equivalent system temperatures outside the atmosphere
of 150 K SSB at 3 mm, and 250 to 400 K SSB at 1.3 mm.
The spectral correlators covered 954 km s−1 at 3.4 mm and
670 km s−1 at 1.3 mm, with resolutions of 8 and 4 km s−1,
respectively. Flux densities were calibrated with 3C273
and MWC349, for which we adopted the following val-
ues: for 3C273, 11 Jy at 1.3 mm and 18 Jy at 3.4 mm; for
MWC349, 1.7 Jy at 1.3 mm and 0.9 Jy at 3.4 mm. The sys-
tematic uncertainties in the flux scales are estimated to be
±5% at 3.4 mm and ±10% at 1.3 mm.
D. Downes et al.: Proposed identification of Hubble Deep Field submillimeter source HDF 850.1
3
Fig. 1. IRAM interferometer map at 1.3 mm (236.3 GHz) of
the continuum emission in a 24′′ field centered on the SCUBA
position of the submm source HDF850.1 (cross; the arms of the
cross correspond to the 3σ positional uncertainty of Hughes et
al. 1998). The synthesized beam is 2.1′′ × 1.7′′ (box at lower
left). The map has not been corrected for the response of the
primary beam of the 15 m dishes at 1.3 mm (20′′ FWHM;
dashed circle). The first positive and negative (dashed) con-
tours are at ±0.5 mJy (2 σ), and the contours thereafter are in
steps of 0.25 mJy (1 σ).
Fig. 2. Interferometer map at 3.4 mm (88.7 GHz) of the con-
tinuum emission in a 64′′ field centered on the original SCUBA
position (cross) of the submm source HDF850.1. No source is
detected, to a limit of 0.36 mJy (3σ). The synthesized beam is
4.7′′ ×4.2′′ (lower left). The map has not been corrected for the
response of the primary beam at 3.4 mm (53′′ FWHM; outer
dashed circle). The inner dashed circle shows the primary beam
at 1.3 mm (see Fig. 1). The first positive and negative (dashed)
contours are at ±0.24 mJy (2 σ), and the contours thereafter
are in steps of 0.12 mJy (1 σ).
Phases were
radio sources
calibrated with the
1044+719, 1125+596, and 1300+580,
for which we
adopted the positions listed in Table 1. All three phase
calibrators were observed every 20 min, at 1.3 mm and
3 mm simultaneously, and the data were phase calibrated
with a weighting by the square of their signal-to-noise ra-
tios and by the inverse square of the system tempera-
ture. To correct the amplitudes and the phases at both
3 mm and 1.3 mm for atmospheric seeing, we used the
1.3 mm continuum total power measurements to compute
the changes in electrical path length due to short-term
fluctuations in the atmospheric water vapour content. For
the 1.3 mm data, we then adopted the phase vs. time cal-
ibration curve at 3 mm (scaled to 1.3 mm) and on top of
this, solved for a second-order phase calibration curve to
fit the phase residuals at 1.3 mm.
We used both uniform and natural weighting of the u, v
plane data to make maps. There is not much difference in
the final result, and the maps shown here (Figs. 1 & 2)
are the ones made with natural weighting. The 1.3 mm
map in Fig. 1 includes all the data, in both sidebands
and in the compact and extended configurations, and has
been deconvolved with the CLEAN algorithm by Clark
(1980), with 100 iterations.
3. Results
3.1. Detection at 1.3 mm
A continuum source was well detected at 1.3 mm
(236.3 GHz),
in the D configuration on November 17
and November 22 in the double sideband data, and in
the upper and lower sidebands separately. All of these
data subsets agree in the source position and flux den-
sity, within the errors. On the map that results from
adding all the data together (Fig. 1), the flux density
is 2.2 ± 0.3 mJy. This result agrees with the single-dish
value of 2.1 ± 0.5 mJy measured by Hughes et al. (1998)
at 1.35 mm.
The 1.3 mm source is at a position 2.4′′ west and 0.7′′
south of the SCUBA position of the submm source HDF
850.1 found by Hughes et al. (1998). The SCUBA detec-
tion was made with the 15 m JCMT single dish, with a
14.7′′ beam at 850 µm. Hughes et al. quoted r.m.s. posi-
tional uncertainties of 0.7′′ in each coordinate. The inter-
ferometer position differs from the SCUBA position at the
3.6 σ level, indicating that there may have been an unsus-
pected systematic error in the SCUBA positions, which in
turn may affect some of the other submm/optical associ-
ations suggested by Hughes et al. (1998).
4
D. Downes et al.: Proposed identification of Hubble Deep Field submillimeter source HDF 850.1
fluxes of HDF 850.1 would be only 0.13 mJy at 2.8 mm and
0.07 mJy at 3.4 mm.
3.4. No spectral lines detected
Although our main goal was the detection of dust emission
at 1.3 mm, we chose the frequency in the upper sideband,
237.824 GHz, to coincide with the CO(9 -- 8) line at z =
3.36, the tentative redshift assigned by Zepf, Moustakas,
& Davis (1997) to the optical source 3-577.0 located 1.4′′
east of the z ≈ 1 elliptical galaxy 3-586.0. No line was de-
tected in the upper sideband, to a limit of 6 mJy beam−1
in channels 32 km s−1 wide in the velocity range −300 to
+367 km s−1 relative to the frequency of the CO(9 -- 8) line
at this redshift. We also detected no line in the lower side-
band, 3 GHz lower in frequency.
At 3 mm, we initially chose 105.7 GHz to search for the
CO(4 -- 3) line at z = 3.36. At this frequency, no line was
seen, to a limit of 2 mJy beam−1 in channels 30 km s−1
wide, in the velocity range ±400 km s−1 around the fre-
quency of the CO(4 -- 3) line at this redshift. Later on,
we tuned the 3 mm receivers to 88.738 GHz to search for
the CO(1 -- 0) line at z = 0.299, the redshift of the bright
disk galaxy, 3-659.1, located 10′′ to the southwest of HDF
850.1. In velocity channels 34 km s−1 wide, no line was seen
to a limit of 2 mJy beam−1, anywhere in our 53′′ primary
beam, in the velocity range −450 to +460 km s−1 cen-
tered on 88.738 GHz. For a galaxy with lines ∼ 200 km s−1
wide, this corresponds to a limit of 0.4 Jy km s−1 for
the CO(1 -- 0) line flux. For comparison, the Milky Way
has a CO(1 -- 0) line luminosity of 4 × 108 K km s−1 pc2
inside the solar circle (Rivolo & Solomon 1988), so at
z ≈ 0.3, the Milky Way would have a CO(1 -- 0) line flux of
0.05 Jy km s−1.
4. Astrometry
We estimate the r.m.s. positional uncertainty to be 0.3′′ for
the centroid of the 1.3 mm dust source. This limit is set by
the baseline accuracy of the interferometer, the availabil-
ity of phase calibrators in this part of the sky, the signal to
noise ratio, and the seeing. Thus far, the best astrometry
done with the interferometer was for the source W3(OH),
with phases calibrated on the quasar 0224+671, located
5◦ away. In those projects (Wink et al. 1994; Wyrowski et
al. 1997; 1999), the global statistical positional repeatabil-
ity was 0.05′′ r.m.s., including the systematic errors in the
baseline geometry. To this must be added the 0.01′′ uncer-
tainty in the Jodrell Bank -- VLA position of the calibrator
source. For the current Hubble Deep Field project, our
strongest calibrator is 14◦ away (Table 1), so the r.m.s.
error in position would be three times worse, namely 0.2′′,
also with Jodrell Bank-VLA calibrator uncertainties of
0.01′′ (Patnaik et al. 1992).
To reduce errors, we chose the three phase calibrators
at earlier and later right ascensions, and lower and higher
Fig. 3. Visibility plot for the continuum source detected at
1.3 mm. The plot shows the real part of the visibility ampli-
tude vs. baseline, for u, v-plane data averaged in circular bins
of 20 m spacing, with 1σ error bars. The data do not have
a high enough signal-to-noise ratio to yield an angular size.
The "best-fit" (solid curve) is for a 2.2 mJy gaussian source of
FWHP diameter 0.7′′. The dashed curves show the visibility
functions at the 2 σ uncertainty of the fit, for a 1.7′′ source
(lower dashed curve) and a point source (upper dashed curve).
3.2. Size of the 1.3 mm source
The source is too weak to measure an angular size. In the
maps from the longer-baseline C2 configuration alone, the
flux at the source position is 1.3 ± 0.5 mJy, so the source
may be partially resolved. Our fits to the data in the u, v-
plane yield a "best fit" FWHP diameter of 0.7′′ ± 0.5′′. At
the 2 σ level, the visibility data (Fig. 3) agree with sizes
ranging from 1.7′′ to a point source. What size should
we expect? Since we observe mm/submm dust radiation,
the source cannot be very hot. For example, a circular
source at z ≈ 1.7 (see below) 1 with a dust temperature
of 35 K, which has a dust opacity of unity at an emitted
wavelength of 100µm, would yield a flux of 2.2 mJy at an
observed wavelength of 1.3 mm if its diameter were 0.3′′.
A less opaque source would be even larger.
3.3. No detection of 3 mm continuum
No source was detected at either 2.8 mm (105.7 GHz) or
at 3.4 mm (88.7 GHz), to 3σ limits of 0.5 mJy and 0.4 mJy
respectively (Fig. 2). This is consistent with the 1.3 mm
continuum being optically thin emission by dust. For a
typical dust continuum spectral index of 3.5, the expected
1 We use H0 = 50 km s−1 Mpc−1 and q0 = 0.5 in this paper.
D. Downes et al.: Proposed identification of Hubble Deep Field submillimeter source HDF 850.1
5
Fig. 4. IRAM interferometer map at 1.3 mm (236.3 GHz) of the dust continuum emission superimposed on a greyscale version
of the BVI image from the Hubble Deep Field. The map is the same as in Fig. 1, but here we only plot the 1.3 mm contours at
0.75 mJy (3 σ), 1.25 mJy (5 σ), and 1.75 mJy (7 σ). The cross marked SCUBA indicates the position and 3σ uncertainty of HDF
850.1 as given by Hughes et al. (1998). Small crosses with black dots indicate the positions and 1σ uncertainties of the VLA
sources (Richards 1999) and the large cross marked ISO indicates the position and 1σ uncertainty of the ISOCAM 15µm source
(Aussel et al. 1999). Optical objects are identified with their photometric redshifts, from the references in Table 1, except for
the optical source 3-577.0, which has a tentative spectroscopic redshift of 3.36 (Zepf et al. 1997).
6
D. Downes et al.: Proposed identification of Hubble Deep Field submillimeter source HDF 850.1
declinations than the Hubble Deep Field (Table 1). This
should formally reduce the noise in the baseline errors and
√2
in our weighted average of calibrators by a factor ∼
to √3. To check internal consistency of the reduction soft-
ware, we made maps of the calibrator sources, and on the
maps, the position errors were < 0.1′′. The uncertainties in
their measured cm-radio positions from the Jodrell-VLA
survey are ten times lower (Table 1). Two of the calibra-
tors are weak, ∼ 0.1 Jy at 1.3 mm, but their signal-to-noise
ratio is quite satisfactory at 3 mm. Because we subtract
the scaled 3 mm phase fluctuations from those at 1.3 mm,
we can improve the fit of the phase solution at 1.3 mm, but
there remains a residual instrumental phase error, which
brings the astrometric part of the position error to 0.2′′
r.m.s.
In addition to the astrometric error, there is a statis-
tical error related to the signal-to-noise ratio. For a point
source, this position error due to noise is
(1)
∆θ ≈ (B/2)/(S/N )
where B is the synthesized beam, and S/N is the signal-
to-noise ratio. For our map in configurations D + C2, the
beam B is 2.1′′, and the signal-to-noise ratio is S/N = 7,
so the position error due to noise is ∆θ = 0.2′′ r.m.s. The
convolution of the astrometric error (0.2′′), and the noise
error (0.2′′) then gives a root sum square error of 0.3′′.
5. Source Identification
The identification of the 1.3 mm dust source with an opti-
cal galaxy or a cm-radio source in the Hubble Deep Field
depends strongly on the accuracy of the measured posi-
tions. The radio-optical registration of the Hubble Deep
Field is described by Williams et al. (1996). Their esti-
mated accuracy of this absolute registration is 0.4′′. The
estimated accuracy of the 1.3 mm IRAM position is 0.3′′.
Radio: The new 1.3 mm position clearly rules out the
identification with the optical source 3-577.0 proposed by
Hughes et al. (1998) from the SCUBA data. It also rules
out the identification, suggested by Richards (1999), of the
1.3 mm dust source with the VLA source, 3651+1221, lo-
cated 4.9′′±0.4′′ to the southwest (see further discussion of
this source in the Appendix). The 1.3 mm dust source co-
incides with the tentative VLA source 3651+1226, which
is a 4.5σ detection at 3 cm (Richards et al. 1998).
Optical: On the Hubble Deep Field image (Fig. 4),
the closest object to the 1.3 mm dust source is the arc-like
feature comprising the optical sources 3-593.0 and 3-593.2.
From the optical coordinates published by Williams et al.
(1996), this arc-like feature is 0.8′′ ± 0.7′′ southwest of
the 1.3 mm dust source, where the 0.7′′ error is the sum
of the optical and 1.3 mm position errors (Table 1). We
must take the sum of the errors rather than the root sum
square because the radio-optical registration uncertainty
is a systematic error rather than a random error.
Fig. 5. Allowed values of the dust temperature and gas
mass, as functions of redshift, for an unlensed source, that
are consistent with the observed 1.3 mm flux of 2 mJy and the
450µm/1300µm flux ratio. "Low" redshift (z ≈ 1) solutions re-
quire low temperature (20 K) and high gas mass (6×1010 M⊙).
"High" redshift (z ≈ 3) solutions can have higher dust temper-
atures (60 K) and lower gas mass (8 × 109 M⊙).
The Elliptical: The next nearest optical object is the
source 3-586.0, located 1.0′′ ± 0.7′′ east of the 1.3 mm dust
source. The optical source 3-586.0 is an elliptical galaxy
at a photometric redshift of 1.0 ≤ z ≤ 1.2 (Hogg et al.
1996; Lanzetta et al. 1996 (their source number 3-306);
Mobasher et al. 1996; Gwyn & Hartwick 1996; Sawicki et
al. 1997; Wang et al. 1998; Fern´andez-Soto et al. 1999 --
their source 303). The optical photometry of this elliptical
galaxy is tabulated by Fasano et al. (1998), and its spectral
energy distribution (SED) is given by Cowie (1999). The
spectrum rises from the optical regime to a maximum near
2 µm, where it flattens out at a level of ∼ 10 µJy, consistent
with the non-detection at a level of 23 µJy at 15 µm (Oliver
et al, in prep., reported by Hughes et al. 1998; Aussel et
al. 1999).
From the position alone, the dust source could be the
arc-like feature or the elliptical. We argue, however, that
the 1.3 mm dust source is probably associated with the
arc-like feature. Although we cannot associate the dust
source with this optical object from the astrometry, we
now present several physical arguments that support this
association and disfavor the elliptical.
5.1. Dust flux ratios, dust temperature, and gas mass
The source HDF 850.1 is not detected at 450 µm to a
limit of 21 mJy (Hughes et al. 1998), nor is it in the IRAS
Faint Source Catalog, nor in the ISO images of the Hubble
Deep Field at 6.7 µm and 15 µm (Aussel et al. 1999). We
thus have a limit on the 450 µm to 1300 µm flux ratio,
where the emission is optically thin (the spectral index is
∼ 3). This ratio is S(450µm)/S(1300µm) ≤ 10. Let the
corresponding observed frequencies (667 and 236 GHz) be
D. Downes et al.: Proposed identification of Hubble Deep Field submillimeter source HDF 850.1
7
ν2 and ν1, respectively. Because the shape of the Planck
function is preserved with redshift, the ratio is the same
at the emitted frequencies, namely
S(ν2)
S(ν1)
= (cid:18) ν2
ν1(cid:19)3 (ehν1(1+z)/kTd − 1)
(ehν2(1+z)/kTd − 1)
(1 − e−τ (ν2(1+z)))
(1 − e−τ (ν1(1+z)))
(2)
where Td is the dust temperature and τ (ν) is the dust op-
tical depth. Experience in millimeter detection of dust at
high redshifts during the past decade shows that sources
are only detected when they contain large amounts of dust
(∼ 108 M⊙) -- as in the IR ultraluminous galaxies which
have dust opacities equal to unity near a wavelength of
100 µm (e.g., Downes, Solomon, & Radford 1993). At this
wavelength (frequency 3 THz), one may approximate the
dust opacity as τ (ν) = (ν/3)n, where ν is the emitted fre-
quency in THz, and n is the dust emissivity index, where
n ≈ 1.5 to 2. Regardless of whether the dust is optically
thick or thin, however, at "low" redshifts (z < 1), the ra-
tio in eq.(2) can be as low as the observed value of ≤ 10
only with dust temperatures Td ≤ 20 K. (Fig. 5).
From the mere detection of dust at 1.3 mm with a flux
density of 2.2 mJy, we can make a useful deduction about
the gas mass from the formula (cf. Downes et al. 1992)
Way were moved to z = 1, at an emission distance (angu-
lar size distance) of 1.75 Gpc, then the 1.3 mm observed
flux would be only 36 µJy, that is, 60 times lower than the
dust flux we detect.
At higher redshifts, the observed limit on the 450 µm
to 1300 µm flux ratio allows the dust temperature to be
higher. For example, at z = 3, the dust temperature can
be 60 K, comparable with the blackbody dust temperatures
deduced for IR ultraluminous galaxies from their IRAS
fluxes (e.g. Downes & Solomon 1998). The corresponding
molecular gas masses deduced from the observed 1.3 mm
flux also become plausible -- a few ×109 M⊙, even in the
absence of gravitational magnification. Because the line
of sight to the source passes so close to the elliptical --
a good gravitational lens -- the observed mm and submm
fluxes may be gravitationally magnified, and the real gas
mass even lower.
In summary, our argument, based on the mm and
submm fluxes alone, that the dust source cannot be a
"low" redshift (z ≤ 1) object, is as follows: a) the op-
tical depth is not high; b) the observed flux ratio may be
obtained at low redshift only if the dust temperature is
sufficiently low (< 20 K), and c) the observed flux plus a
low temperature requires an excessive gas mass.
M
[M⊙]
=
1.6 106
(1 + z)
Sνobs
[Jy] (cid:20) DL
[Mpc](cid:21)2(cid:20) Td
[K](cid:21)−1(cid:20) νem
[THz](cid:21)−(2+n)
(3)
5.2. Other arguments against identifying HDF 850.1 with
the elliptical galaxy 3-586.0
where M is the gas mass, S is the flux density, DL is
the luminosity distance, Td is the dust temperature, νem
is the emitted frequency, and n is the dust emissivity in-
dex, which we adopt to be n = 1.5. The main assump-
tion is that the emission comes from optically thin dust.
The equation is for a dust mass absorption coefficient of
0.11 m2 kg−1 of dust at an emitted wavelength of 1.3 mm,
as in dense molecular clouds (e.g., Krugel & Siebenmor-
gen 1994, their Fig. 12), and a gas-to-dust mass ratio of
100. The uncertainties are about a factor of three in each
direction -- higher and lower (for an analysis, see Hughes,
Dunlop, & Rawlings 1997). Higher gas-to-dust ratios mean
even larger gas masses.
To have a flux density of 2 mJy at 1.3 mm with the
low dust temperature derived above, a "low"-redshift
(z ≈ 1) source would need a very high molecular gas
mass, approaching 1011 M⊙. Lower dust temperatures or
less opaque dust or higher gas-to-dust ratios would all im-
ply even higher gas masses. Such an enormous quantity of
dust and molecular gas would have dramatic consequences
for star formation, and such a large, gas-rich galaxy at low
redshift would have been easily seen on the HDF images
or in the follow-up optical spectroscopy of these sources
(e.g. Zepf et al. 1997). For comparison, a large galaxy like
NGC 891, at a distance of 10 Mpc, has a total 1.3 mm dust
continuum flux of 730 mJy, corresponding to a molecular
gas mass of 1.5 × 109 M⊙ (Gu´elin et al. 1993), about the
same as that of the Milky Way. If NGC 891 or the Milky
A number of other facts argue against the optical source
3-586.0, an elliptical galaxy at z ≈ 1, being the optical
object associated with the 1.3 mm dust source.
1) The gas and dust masses of ellipticals are generally
too low for 1.3 mm dust detections (Mgas < 1010 M⊙).
Even in nearby ellipticals, there are very few detections
(see, e.g., the review by Knapp 1999). The three galaxies
with dust detections > 3σ reported by Wiklind & Henkel
(1995) have gas masses ranging from 1×105 to 7×108 M⊙.
The much larger dust and gas mass predicted by the phys-
ical arguments in the previous section would be totally
inconsistent with the gas content of an elliptical galaxy
like 3-586.0. Such galaxies typically have a total dynami-
cal mass (mostly stars) of 3 × 1010 M⊙ within a radius of
1 kpc, and, with a de Vaucouleurs profile, enclose a total
mass of 2× 1011 M⊙ at a radius of 7 kpc ( 0.85′′ at z ≈ 1).
This argument from the dust and gas mass needed to yield
2 mJy at 1.3 mm is equivalent to the empirical luminosity
argument by Hughes et al. (1998) -- that the mm/sub-
mm luminosity of the source HDF 850.1 is two orders of
magnitude too high to be emission by dust in the z ≈ 1
elliptical galaxy 3-586.0. Therefore the dust source cannot
be the z = 1 elliptical galaxy.
2) Dust has not yet been detected at 1.3 mm in any
"normal " galaxy at z ≈ 1. The reasons are firstly, that
the redshift is not high enough to shift warm dust emission
into the millimeter band, and secondly, that the emission
distance (angular size distance) is close to its maximum.
8
D. Downes et al.: Proposed identification of Hubble Deep Field submillimeter source HDF 850.1
A striking exception is the extremely red, z = 1.44 galaxy
HR 10 (or ERO J164502+4626.4), for which Cimatti et
al. (1998) and Dey et al. (1999) find mm/submm dust
fluxes corresponding to an astonishing far-IR luminosity
of ∼ 1013 L⊙, a molecular gas mass of ∼ 1011 M⊙, and
an apparent star formation rate of 1000 to 2000 M⊙ yr−1.
It will be important to investigate whether this object is
gravitationally lensed or not.
3) We argued above that if the object is at z ≈ 1, the
dust must be cold. So far, however, the mm detections of
dust in high-z objects are mainly of "warm" dust (T >
30 K), probably circumnuclear regions heated by an AGN
and/or a nuclear starburst. Therefore it is most unlikely
that we are detecting the elliptical at z ∼ 1.
6. Discussion
6.1. The dust source HDF850.1 as an unlensed source
Hughes et al. (1998) gave a very plausible interpretation
of HDF850.1 as a high-redshift starburst galaxy, similar
to the ultraluminous infrared galaxies at z < 0.3. As dis-
cussed above, from the interferometer position measure-
ment and the physical arguments about the gas mass, it
is likely that the dust source HDF850.1 is associated with
the optical arc-like feature 3-593.0/593.2. The photomet-
ric redshift estimates of the arc are z = 1.73 (Rowan-
Robinson 1999) and z = 1.76 (Fern´andez-Soto et al. 1999).
In this part of the sky there seems to be a cluster or a
group of galaxies at photometric redshifts z ≈ 1.7 (the
objects 3-593.0, 3-633.1, 3-550.2, and 3-613.0). It is con-
ceivable that the dust source is at this redshift as well, and
that it is part of the same object as the optical arc-like fea-
ture. Such a redshift for HDF850.1 would be well within
the redshift distribution of the identified sources detected
so far with the SCUBA array in other fields (e.g., Barger
et al. 1999a,b; Eales et al. 1999; Lilly et al. 1999).
No optical lines were detected in the arc by Zepf et
al. (1997). Why not? A simple explanation is that the ob-
ject is at 1.2 < z < 2.4, consistent with the photometric
redshifts, so that no strong lines are shifted into the op-
tical. As shown by the data of Zepf et al. (1997), the arc
3-593.0 is not likely to be at z ≥ 3, because it is too blue
-- it shows up nicely in the HST F450W band, and it has
significant flux in the F300W band. It is not a "U-band
dropout", that is, the hydrogen along the line of sight
does not greatly decrease the flux in the F300W band,
as it would if the object were at z ≥ 3, again consistent
with the photometric redshift of z ∼ 1.7. The arc is much
bluer than, for example, the object 3-577.0, which has a
photometric redshift of 2.88 (see, e.g, the colors tabulated
in Zepf et al. 1997).
In fact, the data on HDF 850.1, if combined with
the optical fluxes of the optical arc-like feature 3-593.0
and a redshift of 1.7, agree rather well with the spectral
energy distribution of the ultraluminous infrared galaxy
Fig. 6. Spectral energy distribution for the dust source HDF
850.1 and the optical source 3-593.0 if they are at z ≈ 1.7,
compared with the spectral energy distribution of the ultralu-
minous IR galaxy VII Zw 31 at z = 0.054. Data for HDF 850.1
are from this paper and Hughes et al. (1998). For 3-593.0, fluxes
are from Fern´andez-Soto et al. (1999). For the VLA source
3651+1226, cm-radio fluxes are from Richards et al. (1998) and
Richards (1999). For VII Zw 31, IR and optical data are from
Trentham et al. (1999); mm limits are from Downes & Solomon
(1998), and the 20 cm flux is from Condon et al. (1996). The
solid curve shows a spectrum of the far-IR emitting dust com-
ponent, from the model described in the text. The dashed line
is for an assumed synchrotron spectral index α = −0.75 for
VII Zw 31.
VII Zwicky 31. This object, which was interpreted by
Djorgovski et al. (1990) as a merger-induced starburst
galaxy, was recently selected for study with the HST by
Trentham, Kormendy, & Sanders (1999) because its "cool"
spectral energy distribution indicates it is an ultralumi-
nous galaxy that derives its IR luminosity from an em-
bedded starburst only, with no "warm" AGN component.
Figure 6 shows this comparison with VII Zw 31, if HDF
850.1 is associated with the optical object 3-593.0, that
has a photometric redshift of z ≈ 1.7. For VII Zw 31, we
took the IRAS and HST data from Trentham et al. (1999),
re-scaled for H0 = 50 km s−1 Mpc−1. The luminosities of
the far-IR peak of HDF 850.1 and the optical emission of
3-593.0, calculated for z = 1.7, correspond quite well to
the observed luminosity of VII Zw 31. The solid curve in
Fig. 6 shows the predicted spectrum of the cool dust com-
ponent from the radiative-transfer model for VII Zw 31
by Downes & Solomon (1998). In that model, the dust
temperature varies with galactocentric radius. The denser
dust that is opaque at 100 µm has a temperature of 50 K.
The millimeter dust emission is transparent over the full
extent of the source, for which the global spatial average
of the dust temperature is ∼ 35 K.
There is also a good agreement in the size of 3-593.0
and the dust region in VII Zw 31. On the Hubble Deep
Field image, the bright part of 3-593.0 has a radius of
D. Downes et al.: Proposed identification of Hubble Deep Field submillimeter source HDF 850.1
9
0.22′′, or ∼ 1.9 kpc at z ≈ 1.7. The total emission along
the arclike feature, including the faint optical object 3-
593.2, has an radius of 0.67′′, or 5.6 kpc. In VII Zw 31, the
relevant dimensions are nearly the same. The half-power
radius of the rotating molecular gas disk in VII Zw 31 is
1.7 kpc, with an outer radius of 5 kpc (Downes & Solomon
1998), and the ring of star-forming knots in VII Zw 31 has
a radius of 3.4 kpc (Mazzarella et al. 1999; image in Tren-
tham et al. 1999). This good match to the star-forming
region of VII Zw 31 in both size and spectral energy dis-
tribution, if HDF 850.1 is at z ≈ 1.7 and associated with
3-593.0, suggests a plausible interpretation of HDF 850.1
as an ultraluminous IR galaxy with the same intrinsic
8 µm to 1000 µm luminosity as VII Zw 31, namely, LIR
= 2 × 1012 L⊙.
Although the best-fit photometric redshift for the opti-
cal arc-like feature 3-593.0 is with an Scd galaxy template
at z = 1.73 to 1.76 (Fern´andez-Soto et al. 1999; Rowan-
Robinson 1999), the redshift likelihood function does have
a much smaller alias as a starburst galaxy at z = 2.6
(Fern´andez-Soto, private communication) to 3.0 (Rowan-
Robinson 1999). The fit is not as good as for z = 1.73, in
particular for the J-magnitude, but it is not implausible.
This is in fact the case discussed by Hughes et al. (1998;
see their Fig. 5). The spectral energy distribution would
still match that of VII Zw 31, as in Fig. 6, scaled up by
a factor of 3.6 (to correct for the greater luminosity dis-
tance), except in the optical range, where the arc source
3-593.0 would then be much more blue than VII Zw 31.
A better fit for both HDF 850.1 and 3-593.0 would be
obtained with the other two sources measured by Tren-
tham et al. (1999), namely, IRAS 12112+0305 and IRAS
F22491−1808. Because of the greater luminosity distance
at z = 3, the 8 µm to 1000 µm luminosity would have to
be LIR = 8.2 × 1012 L⊙, which would place HDF 850.1
among the most powerful of the ultraluminous galaxies,
like Mrk 231 (see, e.g., the sample of Sanders et al. 1988).
In summary, the arguments in favor of the interpreta-
tion as an unlensed source at z ∼ 1.7 are the fairly good
match in both size and spectral energy distribution with
VII Zw 31, the better fit of the optical template with a
photometric redshift of 1.7 for 3-593.0, the absence of any
strong lines in the optical spectrum, which is better un-
derstood if z ∼ 1.7, the apparent cluster of objects in the
range z = 1.6 to 1.8 in this region, the blue color of the
arc, and the fact that the total luminosity of 2 × 1012 L⊙
is not exceptional in the class of ultraluminous galaxies,
consistent with the absence of any indication of an em-
bedded optical quasar or Seyfert 1 nucleus (see, e.g., Kim
& Sanders 1998; Solomon et al. 1997).
The arguments in favor of an interpretation as an un-
lensed source at z ∼ 3 have been given by Hughes et al.
(1998) and Carilli & Yun (1999). To these arguments we
may add the poorer-fitting alias for a photometric red-
shift of 2.6 or 3.0. The consequence is that the IR lumi-
nosity would place the source among the most powerful of
the ultraluminous galaxies, those containing a quasar or a
Seyfert 1 nucleus. These objects are fairly rare in the exist-
ing samples of ultraluminous galaxies (see the luminosity
histogram of Kim & Sanders 1998; their Fig. 3), and usu-
ally have a small, bright, symmetric nucleus resembling a
reddened QSO (Surace et al. 1998). There is no evidence
for such a bright QSO-like nucleus in the optical image
of 3-593.0, which instead looks like a star-forming region
extended over several kpc.
6.2. The dust source HDF850.1 as a lensed source
Another possibility is that HDF 850.1 is a lensed source.
The relative abundance of lensed sources in the submm
is expected to be large, because the submm flux density-
redshift relations of distant, dusty, star-forming galaxies
are flat, leading to a larger fraction of brighter images
in the submm relative to the optical (Blain 1997). The
fraction of gravitationally lensed sources at the 10 mJy
level at 850 µm is > 0.01 (Blain 1998, his Fig. 4; Blain et
al. 1999), orders of magnitude greater than for the corre-
sponding optical sources. There is therefore a good chance
that we are detecting a lensed source, especially if the line
of sight is close to an elliptical galaxy.
Hogg et al. (1996) originally proposed that the arc-like
feature 3-593.0 was a source gravitationally lensed by the
elliptical galaxy at z ≈ 1 because of the 1.8′′ separation
of the arc from the elliptical 3-586, the mass of the el-
liptical as deduced from its luminosity, the appearance of
the arc, concave toward the elliptical, and the fact that
other gravitational lenses tend to be ellipticals. It is possi-
ble that the dust source is part of the same object as the
optical arc-like feature. and lies near a high-magnification
point of the lens.
The non-detection of optical lines in the arc is con-
sistent with 1.2 < z < 2.4 (S. Zepf, private communi-
cation and Zepf et al. 1997). The low end of this range
is close to the redshift of the elliptical, giving an unfa-
vorable geometry for lensing. The high end of the range,
however, yields nearly the same geometry as for IRAS
F10214+4724, where the elliptical is at z = 0.9 and the
source is at z = 2.3. If the optical arc-like feature is at
its photometric redshift (z ∼ 1.7), then the arc source
and the elliptical are in a favorable configuration for lens-
ing. The approximate radius of the circle of images may
be roughly estimated from the standard formula for the
Einstein radius for an isothermal, spherical lens
θE = (cid:18)
σv
186 km s−1(cid:19)2 DA(LS)
DA(S)
(4)
(e.g., Peacock 1999, his eq. 4.14), where σv is the line of
sight velocity dispersion of the lens in km s−1, and DA(LS)
and DA(S) are the angular size distances from the lens to
the source and from the observer to the source respec-
tively.
10
D. Downes et al.: Proposed identification of Hubble Deep Field submillimeter source HDF 850.1
Table 2. Comparison of the HDF 850.1/3-586.0/3-593.0 system with other lensed sources
-- -- -- -- -- Elliptical galaxy -- -- -- -- --
Lens
redshift
zlens
m814
r < 0.2′′
(mag)
m814 mH+K
total
total
(mag)
(mag)
σv
(km s−1)
Source
HDF 850.1 system:
Source
redshift
-- -- -- -- -- Lensed feature -- -- -- -- --
opt./UV
magnifi-
cation
crit.curve m814
total
(mag)
zs
radii
(arcsec)
HDF 3-586.0/3-593.0
1.0
--
23.3
19.5
< 340
∼ 1.7?
1.61
26.1
--
Other lenses:
IRAS F10214+4724
HST 14176+5226
Cloverleaf
0.9
0.81
∼ 1 ?
22.8
--
> 24
20.3
19.7
18.5
--
270
260
> 22.5 ∼ 20.5
< 350
2.286
3.4
2.558
1.2×0.8
2 × 1.5
0.7×0.5
20.44
25.8
17.5
∼ 100
--
18 to 30
HDF 3-586.0/3-593.0 data from Cowie (1999); lens model from Hogg et al. (1996).
IRAS F10214+4724 data and lens model from Eisenhardt et al. (1996); IR data from Evans et al. (1999).
HST 14176+5226 data from Ratnatunga et al. (1995); Crampton et al. (1996); lens model from Hjorth & Kneib (1999).
Cloverleaf data and lens model from Kneib et al. (1998a; 1998b), Turnshek et al. (1997), and Chae & Turnshek (1999).
In the lower-mass,
isothermal, elliptical-potential
model of Hogg et al. (1996), where only 3-593.0 and
3.577.0 were a candidate lensed pair, the optical arc was
close to the critical radius of 1.6′′, from which they de-
rived a velocity dispersion of ≥ 340 km s−1 if the elliptical
galaxy is at z = 1. If the dust source is at z ≈ 1.7, and
is a lensed image 0.8′′ from the elliptical, then the very
rough estimate from eq.(4) also yields a velocity disper-
sion of ∼ 330 km s−1. Hogg et al. (1996) tried to explain
not only the apparent arc 3-593.0, concave toward the el-
liptical, but also the object 3-577.0 as its counterimage,
elongated toward the elliptical. The observations of Zepf
et al. (1997) made this candidate counterimage unlikely,
however, because of the tentative detection of a Lyα line
at z = 3.36 in 3-577.0, and the different B450 − V606 colors
in the arc and the "counterimage" (the arc is bluer than
its candidate "counterimage"; for an alternative interpre-
tation, see Dickinson 1998).
How could the HDF 850.1 dust source and the optical
arc-like feature be at different positions? There are sev-
eral ways this could happen. In the source that is lensed,
the dust emission may not be in the same place as the
UV/optical emission. The dust source and the optical arc
could be different parts of the same galaxy, or a merger of
two galaxies, with their light paths deflected to different
spots by the elliptical. They would intersect the ellipti-
cal's cusp and fold caustics differently, and have different
magnifications, and different image shapes on the sky. If
the source is at z ≈ 1.7 and lensed, its luminosity would
still be that of a starburst galaxy. These objects are often
highly distorted, with double nuclei, or multiple knots of
emission -- good candidates to yield non-coincident opti-
cal and mm/submm images if gravitationally lensed.
Non-coincident images in different wavebands can be
expected quite naturally, as shown by the comparison with
other lensed sources in the next section.
6.3. Comparison with other lensed systems
More than half of the high-redshift objects in which dust
and/or CO has been detected at 1.3 mm are known to be
gravitationally lensed, with magnification factors ≈ 10 for
the mm dust emission in some of the sources (e.g. IRAS
F10214+4724, the Cloverleaf quasar, APM 08279+5255,
MG 0414+0534, and very likely BR 1202−07). Other
submm-selected sources are also lensed, with lower am-
plification factors (e.g., SMM J02399-0136 and SMM
J14011+0252: Ivison et al. 1998; Frayer et al. 1998; 1999).
Table 2 compares the lower-mass model for the ellipti-
cal 3-586.0 from Hogg et al. (1996) with similar models for
IRAS F10214+4724, HST 14176+5226, and the Cloverleaf
quasar, in which the lensing elliptical galaxies are at com-
parable redshifts. An important point in these examples
is that a gravitational lens can image different parts of a
source to different spots on the sky in different wavebands.
In IRAS F10214+4724, the combination of an ellip-
tical at z = 0.9 and a source at z = 2.3 gives strong
magnification with a 1.18′′ asymptotic critical radius for
the UV light redshifted to the optical. The optical arc in
10214+4724 contains two peaks. In the model of Eisen-
hardt et al. (1996), the brighter east peak is interpreted
as the blending of two images merging on the critical
curve, while the west peak is interpreted as the third im-
age. A possible analogy in the HDF 850.1 system might
be the brighter optical source 3-593.0 as the two-image
blend, and the weaker peak 3-593.2 as the third image.
The 10214+4724 counterimage is 100 times fainter than
the main arc, and has not been detected in the cm-radio
or mm-dust emission. The models by Eisenhardt et al.
(1996) for 10214+4724 predict a 0.4′′ radial displacement
in the K-band arc, due to the larger source region at longer
wavelengths. They note that an 0.5′′ infrared source would
be imaged into an elliptical ring with a position angle per-
pendicular to that of the optical arc, with the center of the
D. Downes et al.: Proposed identification of Hubble Deep Field submillimeter source HDF 850.1
11
ring offset from the elliptical. In fact, the HST NICMOS
observations of 10214+4724 by Evans et al. (1999) show
that the peak position of the arc changes even between
wavelengths of 1.1 and 2.1 µm.
In HST 14176+5226, the lens redshift, total magni-
tude of the elliptical galaxy in the HST F814W band, and
calculated line-of-sight velocity dispersion of the elliptical
are very similar to those of the elliptical that lenses IRAS
F10214+4724. In the detailed model by Hjorth & Kneib
(1999) for the lensing of HST 14176+5226, the magnified
images in the Einstein cross configuration actually lie out-
side the critical line, which is an elongated ellipse.
In the Cloverleaf quasar, the lensing galaxy has re-
cently been detected in NICMOS images (Kneib et al.
1998b), and may belong to a galaxy cluster or group at
z ≈ 0.9. The lensing effect may be due to a combination
of the elliptical and the galaxy group. The different source
size in different spectral regions leads to quite different im-
ages. In the Cloverleaf, the four optical images are circular
spots, while the four millimeter sources are arcs (Kneib et
al. 1998a).
6.4. Possible identification of the dust source with the ten-
tative VLA source 3651+1226 and implications from
the radio-FIR relation
From the positions in Table 1 and Fig. 4, the 1.3 mm
dust source appears to coincide with the weak VLA radio
source 3651+1226, listed in the Supplementary Catalogue
of Richards et al. (1998; their Table 5). At a wavelength
of 3.5 cm, this 4.5σ source has a (revised) sky flux den-
sity of 7.5±2.2 µJy (E. Richards, private communication).
From the map by Richards (1999) one may set a 3σ up-
per limit of 23 µJy at 20 cm. The dust emission extrap-
olated from the 2 mJy measured at 1.3 mm, would yield
only 0.02 µJy at 3.5 cm, so the cm-radio source must be
mainly synchrotron emission, not dust. Using this VLA
source in relation with the cm-radio-to-far-IR correlation
in low-redshift star-forming galaxies, Carilli & Yun (1999)
concluded that HDF 850.1 could be at a redshift z ≥ 3,
because the 350/1.4 GHz index α350
1.4 is > 1.
With regard to this submm-to-radio ratio, it is instruc-
tive to compare the observed flux density of HDF 850.1
with that expected from the ultraluminous IR galaxy
Arp 220, if the latter were at high redshift. The result is
shown in Table 3, which indicates that Arp 220 could not
be detected much beyond z ≈ 1.7 at 20 cm, even with deep
integrations with the VLA, a fact stressed by Richards
(1999). An unlensed Arp 220 would not be detectable with
the IRAM interferometer at 1.3 mm to the detection lim-
its reported in this paper, in the range 1.7 < z < 4. It
would also be 3.5 to 6 times weaker at 850 µm than the
flux observed by Hughes et al. (1998) for HDF 850.1. Nor
would it have been detected by the ISOCAM instrument
at 15 µm, in the same range of redshift.
Table 3. Expected unlensed flux densities of Arp 220 at high-z
Arp 220 expected flux at wavelength:
VLA
20 cm
(µJy)
3 × 105
37
14
8
Redshift
z
0.01818
1.7
3.0
4.0
HDF 850.1
IRAM SCUBA ISOCAM
1.3 mm 850 µm
(mJy)
(mJy)
15 µm
(µJy)
8 × 105
180
0.3
0.5
0.8
800
1.2
2.0
2.1
18
1.7
0.8
< 23
2.2
7.0
observed
Values are scaled to Arp 220's observed spectrum:
radio: Sopp & Alexander (1991),
millimeter: Downes & Solomon (1998),
far-IR: Rigopoulou et al. (1996); Fischer et al. (1998);
Klaas et al. (1997), near-IR: Soifer et al. (1999).
< 23
As noted by Carilli & Yun (1999), however, the
submm-to-radio index has a large scatter, so it should be
regarded more as a redshift indicator rather than a red-
shift estimator in the sense of the optical photometric red-
shifts. In particular, for the high-z dust sources detected
to date, there is not only a large scatter in the 350/1.4 GHz
index, but also possible contamination of the radio flux by
emission from an AGN, and modification of the ratio due
to differential magnification effects of gravitational lens-
ing. For example, IRAS F10214+4724, at z = 2.3, has an
index α350
1.4 = 0.7, and APM 08279+5255, at higher red-
shift (z = 3.9) has nearly the same index, α350
1.4 = 0.8. Both
of these indices would be compatible with z < 2, given the
scatter in the index data.
7. Conclusions
1) The IRAM interferometer has detected the source HDF
850.1, the strongest source in the submm survey of the
Hubble Deep Field by Hughes et al. (1998). We measure
a flux of 2.2± 0.3 mJy, in agreement with the flux density
measured at 1.35 mm at the James Clerk Maxwell Tele-
scope. The 1.3 mm flux, together with our non-detections
at 2.8 and 3.4 mm, and the SCUBA flux at 850 µm, are all
consistent with the interpretation by Hughes et al. (1998)
that the radiation is optically thin emission by dust.
2) The improved position from the interferometer
shows the dust source cannot be the optical source 3-577.0,
as suggested by Hughes et al. (1998). The measured posi-
tion also rules out the z = 0.299 galaxy 3-659.1 and the
low surface brightness galaxy 3-633.1 at z = 1.72 sug-
gested by Richards (1999) as an indentification for HDF
850.1. The dust source may coincide with the tentative
VLA source 3651+1226.
3) The dust source may be associated with the optical
arc-like feature 3-593.0/3-593.2, located 1.8′′ to the south-
west of the z ≈ 1 elliptical galaxy 3-586.0. The arguments
in favor of this interpretation are the 0.8′′ ± 0.7′′ sepa-
12
D. Downes et al.: Proposed identification of Hubble Deep Field submillimeter source HDF 850.1
ration of the dust source and 3-593 on the sky and the
temperature-mass argument developed in the text, which
disfavors an association of the dust source with the ellip-
tical. If HDF 850.1 is at z ≈ 1.7, the photometric redshift
of the optical source, then it could be an unlensed galaxy
with a spectral energy distribution and linear dimensions
that are similar to those of the z = 0.054 ultraluminous
starburst galaxy VII Zw 31. Hence, if at z ≈ 1.7 and un-
lensed, HDF 850.1 must be an ultraluminous IR galaxy,
with an IR luminosity of LIR = 2 × 1012 L⊙.
4) Because of the close proximity on the sky of the el-
liptical 3-586.0, the dust source HDF 850.1 may be gravi-
tationally lensed. The arguments in favor of this interpre-
tation are the arclike morphology of the optical source that
suggests a gravitational mirage; the similarity of the prop-
erties of the elliptical galaxy to the ones responsible for
the lensing in IRAS F10214+4724, HST 14176+5226, and
the Cloverleaf; and the fact that many of the other high-
redshift objects detected to date at mm/sub-mm wave-
lengths in dust and CO are magnified by gravitational
lensing. The arguments against this interpretation are that
the arc morphology is not conclusive, and that the candi-
date counterimage, 3-577.0, has a different B − V color,
and apparently a different redshift.
The discussion in this paper illustrates the absolute
necessity of sensitive millimeter/submillimeter interferom-
eters for identifying sources of dust emission with cosmo-
logically distant optical galaxies. The next step in this di-
rection is the construction of a mm/submm interferometer
with more than an order of magnitude greater sensitivity
than presently available.
Acknowledgements. We thank the operators on Plateau de
Bure for their help in observing, M. Bremer (IRAM) for
some of the image processing, and H. Aussel (CEA-Saclay),
M. Dickinson (STSci), A. Fern´andez-Soto (Univ. New South
Wales), J.-P. Kneib (Obs. Midi-Pyr´en´ees), E. Richards (Univ.
Virginia), P.A. Shaver (ESO), and S. Zepf (Yale Univ.) for very
useful comments. We also thank the referee, A. Omont, for his
helpful suggestions on improving the paper.
Appendix: The cm-radio source 3651+1221
Richards et al. (1998) tentatively associated the VLA
source 3651+1221 with the z = 0.299 galaxy 3-659.1
(cf. Fig. 4). This galaxy may also be the 15 µm source
HDF PM3 29 detected with ISOCAM (Aussel et al. 1999).
Richards (1999) showed however that the radio emission
is confined to the faint emission north of the spiral, about
1′′ south of the optical object 3-633.1, which has a pho-
tometric redshift of 1.72 (Fern´andez-Soto et al. 1999).
Since then, this optical object has been detected in J and
H bands, having colors consistent with a high-z galaxy
(Dickinson 1998). Richards (1999) noted that if the ra-
dio source is at z ∼ 1.7, the implied radio luminosity is
substantially higher than that of the most extreme local
starbursts like Arp 220, which suggests the radio source
may be an AGN.
There is a 3-sigma contour near the spiral galaxy on
the IRAM 1.3 mm map (Fig. 4), which we list in Ta-
ble 1 as a questionable source. The 1.3 mm position is
1′′ south of the VLA source. The apparent 1.3 mm flux
density is 0.75 mJy, but this would have to be corrected
for the primary beam attenuation to ∼ 1.2 mJy if the
source were real. In this case, the emission could not be
the extrapolation of the synchrotron spectrum from cen-
timeter wavelengths, because the flux density at 3.4 cm is
only 16 µJy (Richards et al. 1998). Could it be dust emis-
sion from the spiral? The problem with dust emission at
low redshifts (z ∼ 0.3) is that most of the energy is at
shorter wavelengths, not redshifted into the mm /submm
bands. The dust in the central parts of the Milky Way,
or in NGC 891, corresponding to a molecular gas mass of
1.5 × 109 M⊙, would yield a 1.3 mm flux of only 68 µJy
at z = 0.299, or ∼ 20 times lower than the apparent flux
on the 1.3 mm map. The apparent 3σ signal would thus
have to come from an extremely gas-rich galaxy, which is
not in any way indicated by the HST optical image of this
relatively nearby object. Since there is also a 3σ negative
peak at a similar radius from the center of the primary
beam (Fig. 1), the apparent 3σ positive signal may not
be real.
References
Aussel, H., Cesarsky, C.J., Elbaz, D., & Starck, J.L. 1999,
A&A, 342, 313
Barger, A.J., Cowie, L.L., & Sanders, D.B. 1999a, ApJ, 518,
(June 10) in press
Barger, A.J., Cowie, L.L., Smail, I., Ivison, R.J., Blain, A.W.,
& Kneib, J.-P. 1999b, AJ, (June 1999) in press (astro-
ph/9903142)
Blain, A.W. 1997, MNRAS, 290, 553
Blain, A.W. 1998, MNRAS, 295, 92
Blain, A.W., Kneib, J.-P., Ivison, R.J., & Smail, I. 1999, ApJ,
512, L87
Carilli, C.L., & Yun, M.S. 1999, ApJ, 513, L13
Chae, K.-H., & Turnshek, D.A. 1999, ApJ, 514, 587
Cimatti, A., Andreani, P., Rottgering, H., & Tilanus, R. 1998,
Nature, 392, 895
Clark, B.G. 1980, A&A, 89, 377
Condon, J.J., Helou, G., Sanders, D.B., & Soifer, B.T. 1996,
ApJS, 103, 81
Cowie,
L.
1999,
http://www.ifa.hawaii.edu/∼cowie/tts
/hdf42.html
Crampton, D., Le F`evre, O., Hammer, F., & Lilly, S.J. 1996,
A&A, 307, L53
Dey, A., Graham, J.R., Ivison, R.J., Smail, I., Wright, G.S., &
Liu, M.C. 1999, ApJ, in press (astro-ph/9902044)
Dickinson, M. 1998, in The Hubble Deep Field, ed. M. Livio,
S.M. Fall, & P. Madau, Cambridge: Cambridge Univ. Press,
219
Djorgovski, S., de Carvalho, R.R., & Thompson, D.J. 1990,
AJ, 99, 1414
Downes, D., Radford, S.J.E., Greve, A., Thum, C., Solomon,
P.M., & Wink, J.E. 1992, ApJ, 398, L25
D. Downes et al.: Proposed identification of Hubble Deep Field submillimeter source HDF 850.1
13
Downes, D., Solomon, P.M., & Radford, S.J.E. 1993, ApJ, 414,
Richards, E.A. 1999, ApJ, 513, L9
L13
Downes, D., & Solomon, P.M. 1998, ApJ, 507, 615
Eales, S., Lilly, S., Gear, W., Dunne, L., Bond, J.R., Hammer,
(http://www.cv.nrao.edu/∼jkempner/vla-hdf)
Richards, E.A., Kellermann, K.I., Fomalont, E.B., Windhorst,
R.A., & Partridge, R.B. 1998, AJ, 116, 1039
F., Le F`evre, O., & Crampton, D. 1999, ApJ, 515, 518
Rigopoulou, D., Lawrence, A., & Rowan-Robinson, M. 1996,
Eisenhardt, P.R., Armus, L., Hogg, D.W., Soifer, B.T., Neuge-
MNRAS, 278, 1049
bauer, G., & Werner, M.W. 1996, ApJ, 461, 72
Evans, A.S., Scoville, N.Z., Dinshaw, N., Armus, L., Soifer,
B.T., Neugebauer, G., & Rieke, M. 1999, ApJ, in press
(astro-ph/9812196)
Fasano, G., Christiani, S., Arnouts, S., & Filippi, M. 1998, AJ,
115, 1400
Fern´andez-Soto, A., Lanzetta, K.M., & Yahil, A. 1999, ApJ,
513, 34 (http://bat.phys.unsw.edu.au/∼fsoto/hdfcat.html)
Fischer, J., Satyapal, S., Luhman, M.L., Melnick, G., Cox, P.,
Cernicharo, J., Stacey, G.J., Smith, H.A., Lord, S.D., &
Greenhouse, M.A. 1998, in ISO to the Peaks, ed., M. Kessler
& M. Perry, Noordwijk: ESTEC, in press
Frayer, D.T., Ivison, R.J., Scoville, N.Z., Yun, M., Evans, A.S.,
Rivolo, A.R., & Solomon, P.M. 1988, in Molecular Clouds in
the Milky Way and External Galaxies, ed. R.L. Dickman,
R.L. Snell, & J.S. Young, Heidelberg: Springer, 42
Rowan-Robinson, M. 1999, in preparation
Rowan-Robinson, M., et al. 1997, MNRAS, 289, 490
Sanders, D.B., Soifer, B.T., Elias, J.H., Madore, B.F.,
Matthews, K., Neugebauer, G., & Scoville, N.Z. 1988, ApJ,
325, 74
Sawicki, M.J., Lin, H., & Yee, H.K.C. 1997, AJ, 113, 1;
(http://www.astro.utoronto.ca/∼sawicki/)
Soifer, B.T., Neugebauer, G., Matthews, K., Becklin, E.E.,
Ressler, M., Werner, M.W., Weinberger, A.J., & Egami,
E. 1999, ApJ, 513, 207
Smail, I., Blain, A.W., & Kneib, J.-P. 1998, ApJ, 506, L7
Solomon, P.M., Downes, D., Radford, S.J.E., & Barrett, J.W.
Frayer, D.T., Ivison, R.J., Scoville, N.Z., Evans, A.S., Yun,
M.S., Smail, I., Barger, A.J., Blain, A.W., & Kneib, J.-P.
1999, ApJ, 514, L13
1997, ApJ, 478, 144
Sopp, H.M., & Alexander, P. 1991, MNRAS, 251, 112
Surace, J.A., Sanders, D.B., Vacca, W.D., Veilleux, S., & Maz-
Gu´elin, M., Zylka, R., Mezger, P.G., Haslam, C.G.T., Kreysa,
zarella, J.M. 1998, ApJ, 492, 116
E., Lemke, R., & Sievers, A.W. 1993, A&A 279, L37
Trentham, N., Kormendy, J., & Sanders, D.B. 1999, AJ, in
Gwyn, S.D.J., & Hartwick, F.D.A. 1996, ApJ, 468, L77
press (astro-ph/9901382)
(http:/uvastro.phys.uvic.ca/grads/gwyn)
Turnshek, D.A., Lupie, O.L., Rao, S.M., Espey, B.R., & Sirola,
Hjorth, J., & Kneib, J.-P. 1999, ApJ, in press
Hogg, D.W., Blandford, R., Kundi´c, T., Fassnacht, C.D., &
C.J. 1997, ApJ, 485, 100
Wang, Y., Bahcall, N., & Turner, E.L. 1998, AJ, 116, 2081;
Malhotra, S. 1996, ApJ, 467, L73
ftp.astro.princeton.edu, cd elt/:HDF
Hughes, D.H., Dunlop, J.S., & Rawlings, S. 1997, MNRAS,
289, 766
Hughes, D., et al. 1998, Nature, 394, 241
Ivison, R.J., Smail, I., Le Borgne, J.-F., Blain, A.W., Kneib,
J.-P., & B´ezecourt, J., Kerr, T.H., & Davies, J.K. 1998,
MNRAS, 298, 583
Kim, D.-C., & Sanders, D.B. 1998, ApJS, 119, 41
Klaas, U., Haas, M., Heinrichsen, I., & Schulz, B. 1997, A&A,
325, L21
Knapp, G.R. 1999, in Star Formation in Early-Type Galaxies,
ed. P. Carral & J. Cepa, San Francisco: Astron. Soc. Pacific,
in press
Kneib, J.-P., Alloin, D., Mellier, Y., Guilloteau, S., Barvainis,
R., & Antonucci, R. 1998a, A&A, 329, 827
Kneib, J.-P., Alloin, D., & Pell´o, R. 1998b, A&A, 339, L65
Krugel, E., & Siebenmorgen, R. 1994, A&A, 288, 929
Lanzetta, K.M., Yahil, A., & Fern´andez-Soto, A. 1996, Nature,
381, 759
Lilly, S.J., Eales, S.A., Gear, W.K.P., Hammer, F., Le F`evre,
O., Crampton, D., Bond, J.R., & Dunne, L. 1999, ApJ, in
press (astro-ph/9901047)
Mann, R.G., et al. 1997, MNRAS, 289, 482
Mazzarella, J.M., et al. 1999, in preparation
Mobasher, B., Rowan-Robinson, M., Georgakakis, A., & Eaton,
N. 1996, MNRAS, 282, L7
Patnaik, A.R., Browne, I.W.A., Wilkinson, P.N., & Wrobel,
J.M. 1992, MNRAS, 254, 655
Peacock, J.A. 1999, Cosmological Physics, Cambridge: Cam-
bridge Univ. Press, 106
Ratnatunga, K.U., Ostrander, E.J., Griffiths, R.E., & Im, M.
1995, ApJ, 453, L5
Wiklind, T., & Henkel, C. 1995, A&A, 297, L71
Williams, R.E., et al. 1996, AJ, 112, 1335
Wilner, D.J., & Wright, M.C.H. 1997, ApJ, 488, L67
Wink, J.E., Duvert, G., Guilloteau, S., Gusten, R., Walmsley,
C.M., & Wilson, T.L. 1994, A&A, 281, 505
Wyrowski, F., Hofner, P., Schilke, P., Walmsley, C.M., Wilner,
D.J., & Wink, J.E. 1997, A&A, 320, L17
Wyrowski, F., Schilke, P., Walmsley, C.M., & Menten, K.M.
1999, ApJ, 514, L43
Zepf, S.E., Moustakas, L.A., & Davis, M. 1997, ApJ, 474, L1
This article was processed by the author using Springer-Verlag
LaTEX A&A style file L-AA version 3.
|
astro-ph/9711209 | 1 | 9711 | 1997-11-18T20:01:43 | Fluctuations in the Extragalactic Background Light: Analysis of the Hubble Deep Field | [
"astro-ph"
] | (Abridged) Statistical analysis of the unresolved light in the Hubble Deep Field (HDF) strongly constrains possible sources of the optical Extragalactic Background Light (EBL). To test for the statistical signature of previously undetected sources, we estimate the auto, cross, and color correlations of the ``sky'' in the HDF that remains after masking objects brighter than I=30 mag. This measurement yields the most stringent limits to date on small-scale structure in the night sky; analysis of shallower imaging would be dominated by galaxies now detected by the HDF. Unless there is a truly uniform optical background, the mean EBL is likely to be within a small fraction of the surface brightness from detected galaxies. No currently plausible sources of additional EBL satisfy the constraints that they (1) would not have already been detected, (2) contribute EBL comparable to that from detected galaxies, and (3) do not produce EBL fluctuations in excess of the upper limits set by correlations in the HDF. | astro-ph | astro-ph |
Fluctuations in the Extragalactic Background Light:
Analysis of the Hubble Deep Field1
Submitted to ApJ
Michael S. Vogeley2
Princeton University Observatory
Princeton, NJ 08544-1001
[email protected]
ABSTRACT
Statistical analysis of the unresolved light in the Hubble Deep Field (HDF)
strongly constrains possible sources of the optical Extragalactic Background
Light (EBL). This constraint is crucial for determining the spectrum of the
EBL because reported upper limits on the optical EBL are several times larger
than the surface brightness from detected galaxies, suggesting the possibility of
additional galaxy populations. To test for the statistical signature of previously
undetected sources, we estimate the auto, cross, and color correlations of
the "sky" in the HDF that remains after masking objects brighter than
I814 = 30mag. Auto and cross correlations of surface brightness in the V606 and
I814 bandpasses are well-fitted by ω(θ) ∼ 10−6(θ/1′′)−0.6 up to 10′′. Probable
contributions of several instrumental systematics ensure that these correlations
are firm upper limits on the true EBL fluctuations. This measurement yields the
most stringent limits to date on small-scale structure in the night sky; analysis
of shallower imaging would be dominated by galaxies now detected by the HDF.
Unless there is a truly uniform optical background, the mean EBL is likely to
be within a small fraction of the surface brightness from detected galaxies. No
currently plausible sources of additional EBL satisfy the constraints that they
(1) would not have already been detected, (2) contribute EBL comparable to
that from detected galaxies, and (3) do not produce EBL fluctuations in excess
of the upper limits set by correlations in the HDF. These constraints admit only
1 Based on observations with the NASA/ESA Hubble Space Telescope, obtained at the Space Telescope
Science Institute, which is operated by AURA, under NASA contract NAS 5-26555
2 Hubble Fellow
-- 2 --
a confusion-limited population of extremely low surface brightness objects that
is disjoint from the parameter space of all detected galaxies. Extrapolation of
detected galaxy counts to zero flux would add only a few percent to the EBL.
Diffuse intergalactic light clustered similarly to faint galaxies could explain some
of the observed correlations but would contribute at most a few ×10% to the
mean EBL.
Subject headings:
evolution -- methods: statistical
cosmology: observations -- galaxies: clustering -- galaxies:
1.
INTRODUCTION
The Extragalactic Background Light (EBL) is the surface brightness of the night sky
from all extragalactic sources, integrated out to the first epoch of star formation. The
ultimate goal of studying the EBL is to complete the resolution of Olbers's paradox by
measuring the spectrum of the EBL, identifying the sources that contribute to it, and
accounting for their evolution. Constraints on the optical EBL are particularly important
because these bandpasses are sensitive to rest-frame UV flux from galaxies at z > 1, when
we now believe the bulk of star formation in the universe occurred (Madau et al. 1996;
Connolly et al. 1997). Thus, measurement of the spectrum of the optical EBL is a critical
test for models of the star formation history of the universe (see, e.g., Whitrow & Yallop
1965; Partridge & Peebles 1967a,b; Fall, Charlot, & Pei 1996; Vaisanen 1996; Madau et al.
1997). Models for the sources of the EBL must also correctly predict statistical fluctuations
in the EBL, which reflect the surface brightness profiles and clustering of these sources.
[For an extensive review on the optical EBL that predates the HDF, see Tyson 1995.]
The average surface brightness of the sky from detected galaxies sets a firm lower limit
on the mean EBL, µEBL ≥ µgalaxies (here and throughout this paper, µ is an average surface
brightness). Tremendous progress has been made in detecting faint galaxies, through deep
surveys such as the Hubble Deep Field (Williams et al. 1996) and the near-IR survey
by Djorgovski et al. (1995), which reach as faint as V = 29.5mag and K = 24mag,
respectively. However, it is unclear what fraction of the mean optical EBL the detected
galaxies comprise, because the uncertainties in the mean level of the optical EBL are larger
than the surface brightness contributed by these galaxies. In addition, the systematic bias
of object detection against relatively diffuse, low surface brightness objects leaves open the
possibility that we may have missed a substantial fraction of the optical luminosity density
in the universe (Disney 1976; McGaugh et al. 1995; Dalcanton 1997).
-- 3 --
Figure 1 shows various observational constraints on the mean EBL. The inset box plots
only the optical limits, using linear axes. Solid symbols indicate lower limits on the EBL
in several bandpasses from the average sky brightness contributed by detected galaxies,
µgalaxies. In optical bandpasses, these lower limits include galaxy counts from the HDF and
brighter counts from previous surveys (compilation by Pozzetti et al. 1997). UV galaxy
counts are from Milliard et al. (1992). Open circles with 1σ uncertainties show a recent
detection of the mean optical EBL (Bernstein 1997; Bernstein, Freedman, & Madore 1998).
Arrows are upper limits from various experiments (Paresce 1990; Toller 1983; Dube, Wickes,
& Wilkinson 1977, 1979; Hauser 1996), where the upper bound in each case is set primarily
by uncertainties in foreground subtraction. Other bounds on the optical EBL (not plotted
here) include experiments by Roach & Smith (1968), Lillie (1968), Spinrad & Stone (1978),
Boughn & Kuhn (1986), and Mattila (1990). Precise measurement of the mean level of the
EBL is extremely difficult because several foreground sources, particularly Zodiacal light,
are at least an order of magnitude brighter than the mean optical EBL.
At optical wavelengths, uncertainties in the mean EBL leave room for considerable
surface brightness above the lower limits set by galaxy detections. The 2σ uncertainties
of the optical EBL detection by Bernstein et al. span the range from µEBL ∼ 0 to
µEBL ∼ 5µgalaxies. The gap between previous upper limits on the EBL and the galaxy
surface brightness is of order 4 − 10. These uncertainties prompt us to examine the
plausibility of an additional contribution to the EBL that is comparable to the surface
brightness from detected galaxies with magnitude V < 29.5mag. Could this flux arise from
fainter galaxies? The surface brightness per magnitude depends on the number counts as
dIν/dm = (dN/dm)(Igal(m)) ∝ 10(α−0.4)m. If we extrapolate the galaxy number counts
to infinitely faint limits, assuming the logarithmic count slope α = 0.2 at the HDF limit,
then all galaxies with V > 29.5mag would contribute only an additional 1.6% to the
EBL. If, strangely, the number counts turn up to α = 0.3 beyond V = 29.5mag, then
this contribution rises to 3.3%. Thus, it seems unlikely that galaxies fainter than the
HDF limit contribute significant flux. However, this does not rule out objects that evade
direct detection, e.g., galaxies with total magnitude V < 29.5 and very low central surface
brightness, or a diffuse intergalactic component.
If a substantial fraction of the EBL resides in sources that now lie below the detection
limits of deep optical imaging, this population might be inferred from its statistical signature
in the object-masked "sky." The internal profiles of sources and clustering among them will
cause correlations in the sky brightness. Differences between the spectra of these sources
and the average sky brightness will yield corresponding fluctuations in color (for HST
observations, the "average sky" is primarily Zodiacal light, for ground-based observations
there is also a strong atmospheric component).
-- 4 --
To constrain possible sources of additional EBL, we compute the angular autocorrelation
function of the unresolved light in the Hubble Deep Field. In other words, we constrain
additional sources of EBL by studying the correlation properties of the flux that cannot
be assigned with large statistical significance to individual objects. Fluctuation analysis of
the sky brightness is complementary to direct measurement of the mean EBL; although
the uncertainties in direct measurement would permit a large EBL component in addition
to detected galaxies, the small amplitude of the sky fluctuations rules out many plausible
candidates for such a component. For example, we can statistically test the proposition that
previously undetected low surface brightness galaxies comprise most of the EBL (Vaisanen
1996). This approach is insensitive to a truly uniform optical background, but the most
plausible source of the EBL is emission from stars, which are likely to be clustered. At large
redshift, we observe these in the rest-frame UV, where most of the stellar flux is from O
and B stars. Regions of active star formation are extremely clumpy, so we expect the flux
to be strongly correlated.
The approach of studying the correlation properties of the unresolved optical
background was pioneered by Shectman (1974), who used Schmidt plates to detect optical
EBL fluctuations on several arcminute scales. Shectman detected an EBL power spectrum
that was consistent with clustering of galaxies fainter than R = 18, as predicted by Gunn
(1965) and Shectman (1973). Tyson (1988) found evidence for residual surface brightness
fluctuations in deep B band CCD images after cleaning detected B < 27mag galaxies and
smoothing on a scale of 6′′. Cole, Treyer, & Silk (1992) examined how fluctuation analysis
of deep optical images might constrain faint galaxy populations. Martin & Bowyer (1989)
detected arcminute-scale fluctuations with a rocket-borne UV detector, which indicated the
presence of a UV background from galaxies. Boughn, Saulson, & Uson (1986) measured
the beam-to-beam variance of the sky flux at 2.2µm to set upper limits on galaxy counts
in the infrared, assuming a model for the galaxy clustering. Jenkins & Reid (1991) also
measured sky fluctuations at 2.2µm to test galaxy evolution models. Recently, Kashlinsky
et al. (1997) applied a fluctuation analysis to the DIRBE sky maps to set constraints on the
infrared background. Fluctuations at zero lag have long been used to constrain the number
count distribution of radio sources (Scheuer 1957; Hewish 1961; Condon 1974; see, e.g.,
Fomalont et al. 1988 for a recent result). Hasinger et al. (1993) computed limits on the
X-ray flux distribution in deep ROSAT images and several analyses have been performed to
measure the correlation function of the X-ray background, as observed by HEAO 1 (Persic
et al. 1989) and Einstein (Barcons & Fabian 1989).
In this paper we use the Hubble Deep Field to probe the correlation structure of the
night sky after masking all detected objects and thereby establish constraints on possible
sources of the optical EBL. Section 2 presents estimates of the auto and cross correlation
-- 5 --
of the unresolved sky in the Hubble Deep Field. Section 3 discusses possible contributions
to these measurement from foregrounds and instrumental systematics. In Section 4 we
use these results to set stringent limits on the types of additional galaxy populations that
might lie hidden below current detection limits and examine if some of the measured
correlation signal may be caused by weakly clustered intergalactic light. Section 5 reviews
our conclusions.
We quote all HDF magnitudes in the AB system (Oke & Gunn 1983), defined as
AB = −2.5 log fν − 48.60, where fν is in units of erg/s/cm2/Hz. Distances are in h−1Mpc,
where the Hubble constant is 100hkm s−1 Mpc−1.
2. EBL FLUCTUATIONS IN THE HUBBLE DEEP FIELD
2.1. Treatment of the HDF Data
We analyze the drizzled version 2 images of the HDF observed through the Wide
Field cameras (WF) in the F450W, F606W, and F814W filters (hereafter we refer to these
bandpasses as B450, V606, and I814, respectively). Except for corrections to the flat field
calibration (see below), we use the data as reduced and processed by the HDF team (Williams
et al. 1996). We analyze only the central one-quarter of the area (roughly 41′′ × 41′′) of
each drizzled image, to minimize the possible impact of residual large-scale gradients, and
structure in the dark counts due to fluorescence of the MgF2 window (Burrows et al. 1995).
[For further details of the HDF, see http://www.stsci.edu/ftp/science/hdf/hdf.html.]
We apply corrections to the version 2 HDF images to remove large-scale gradients that
are apparent after object-masking and smoothing of the remaining sky. Comparison of
these features with ratio images formed from the v2 flat field calibrations and more recent
calibration files (the latter are the calibration files currently used in the WPFC2 pipeline)
indicate that the gradients are caused by errors in the v2 calibrations. To remove these
features we smooth the flat field ratio images (to remove small-scale noise) and multiply
them into the HDF images. These gradients are quite small, typically 0.5% across an
entire WF image, and are not expected to affect photometry of individual objects, because
algorithms such as FOCAS estimate the sky level near each object. We remove these
gradients because we are interested in detecting r.m.s. fluctuations in the sky at the level
of 0.1%. In section we discuss the possible contribution of residual flat field uncertainty to
the apparent sky fluctuations.
To prepare an object-masked image, we use the HDF version 2 FOCAS catalog to
identify sources with total magnitudes brighter than I814 = 30 (Williams et al. estimate the
-- 6 --
HDF to be 80% complete to I814 = 29), which includes virtually every object in the catalog.
For comparison, we also derive catalogs from the HDF using the SExtractor package (Bertin
& Arnouts 1996); we find the mask structure to be relatively insensitive to the choice
of catalog and choose the available v2 FOCAS catalog to allow others to reproduce our
results. The same mask, derived from detections in the summed V606 + I814 image, is used
for all filters. For each object we flag the pixels that lie within an ellipse that is twice as
large as the area used by FOCAS to compute the "total magnitude." The FOCAS "total
magnitude area" is always at least twice the area within the detection isophote, so the
masked area is at least four times as large as the isophotal detection region. Smoothing of
the masked images on a range of scales does not reveal residual "doughnuts" around the
masked objects. This masking procedure removes 30% of the pixels in each WF image.
Therefore it does not appear that the sky is confusion-limited to the isophotal detection
threshold of ∼ 27 mag/arcsec2 that was used for the FOCAS catalog.
2.2. Correlation Analysis
The field of view of the WF cameras is well-matched to probing EBL fluctuations
caused by the correlation structure of galaxy profiles. The WF cameras each have a 80′′
field of view of and 0.1′′ pixel scale. At redshift z = 1, 1′′ corresponds to 8.5kpc (Ω = 1 and
H0 = 50km/s). On these angular scales, the clustering amplitude of I814 < 29 galaxies is
ω(1′′) ∼< 0.1 (Colley et al. 1996; Villumsen et al. 1997). This weak clustering implies that
small-scale fluctuations in the EBL will be dominated by the surface brightness profiles of
faint objects rather than by the clustering among them.
For observed surface brightness µ(x), we define the autocorrelation function
C(θ) = hµ(x)µ(x + θ)i − µ2,
(1)
where µ = hµ(x)i. We often refer to the mean surface brightness µ as the mean "sky"
brightness (after masking detected objects, µ corresponds to the usual definition of a "sky
level"). We use Cartesian coordinates x, with θ = x − x′, because the relative angles in the
HDF satisfy θ ≪ 1. The surface brightness autocorrelation may be written as the product
of a dimensionless autocorrelation function and the square of the mean sky brightness
C(θ) = µ2ω(θ).
(2)
To prepare an image for correlation analysis, we compute the mean value of the
unmasked pixels and subtract this mean from the image, set pixels within the object ellipses
to zero, and taper the edges of the image with a cosine bell function. To estimate the
-- 7 --
autocorrelation function, we compute the Fourier transform of the image, square the moduli
of the Fourier coefficients (for cross correlations, we compute the modulus of the products
of the images' Fourier coefficients) and correct these amplitudes for the effective area of the
image after masking and weighting. We then inverse Fourier transform this two-dimensional
power spectrum to produce the two-dimensional autocorrelation function. Averaging over
all angles yields the function C(θ). Computation of C(θ) by direct summation of the
products of the pixel values yields identical results (but uses significantly more CPU cycles).
Analysis of test images with known power spectra confirm that we recover the true signal,
after correcting for the integral constraint bias.
The integral constraint bias arises because we estimate the mean surface brightness µ
from the image itself. The implicit assumption that the ensemble average of µ is identical
to the mean within the image is equivalent to assuming that there are no fluctuations of the
mean surface brightness between different images. This bias causes us to underestimate the
autocorrelation by the integral average of the true autocorrelation function over the area Ω
of the image,
∆ =
(3)
1
Ω2 Z d2xZ d2x′ ωtrue(x − x′),
where ωtrue is the true dimensionless autocorrelation function.
Figure 2 shows the dimensionless autocorrelation function of the object-masked HDF
"sky" in each of the three bandpasses. In each case, we average the results for the WF2, 3,
and 4 fields. Representative error bars are attached to the V606 curve only and are estimated
from the variation among the three fields and within bins of width δ(log10 θ) = 0.4. Near
zero lag, the signal is dominated by photon shot noise over the scale of the 0.1′′ WF pixels.
On scales from 0.15′′ to 8′′, these autocorrelation functions are well fitted by a power law
C(θ) = µ2ω(1′′)(θ/1′′)γ,
(4)
with γ ∼ −0.6. Table 1 summarizes the power law fits and their uncertainties. At 1′′, the
typical fluctuation is ω(1′′) ∼ 10−6, i.e., an r.m.s. correlated fluctuation of ∼ 0.1% of the
mean sky brightness, or ∼ 30.5 mag/arcsec2 in V606.
The turnover of ω(θ) on scales ∼> 10′′ is consistent with the integral constraint bias that
we expect for a power law correlation function over a field of this size. Using the power law
fits in Table 1, we integrate equation (3) over a 41′′ × 41′′ field and find that the correlation
functions in Figure 2 are underestimated by approximately ∆ ∼ 4 × 10−7 (note heavy solid
bar in Figure 2).
What could account for the power law shape of these correlations? In section 3 we
discuss several instrumental systematics and foregrounds that might contribute to the
-- 8 --
measured correlations. We note that the unmasked wings of the point spread function
might cause correlations with similar shape to those plotted in Figure 2. To explain the
shape of the correlation function with extragalactic sources, one can imagine an unclustered
population of sources of different scale size such that the sum of their autocorrelations
yields a power law. Alternatively, the similarity of the power law index, γ ∼ −0.6, to that
observed for galaxy clustering, γ ∼ −0.8, suggests that the clustering of sources causes the
power law shape. However, because we have reason to believe that systematics affect this
measurement, it is premature to fit such models.
If we only mask galaxies brighter than I = 26mag, rather than I < 30mag, we find
autocorrelations at θ < 0.5′′ that are an order of magnitude above the curves in Figure 2.
This test indicates that fluctuation analysis of shallower imaging data would be dominated
by the profiles of galaxies detected by the HDF. In ground-based imaging, this small-scale
correlation structure would be smeared by seeing and the correlation signal from the profiles
of galaxies in the magnitude range 26 < I814 < 29 would be overwhelmingly larger than
the signal measured in the HDF. In other words, the HDF is the first data set in which we
could have detected a signal as weak as that shown in Figure 2. One might worry that, just
as correlations in shallower imaging would be dominated by galaxies below the detection
threshold, galaxies just fainter than the HDF limit dominate the observed correlations; in
section 4 we show that this is not the case.
Cross correlations of images in different filters and comparison of these results with the
respective autocorrelations provide a critical test of the nature of the observed fluctuations.
Dot-dashed curves in Figure 2 show cross correlations of B450 vs. V606 and V606 vs. I814. The
cross correlations are shown as ω(θ) = C1×2(θ)/(µ1µ2), where µ1 and µ2 are the mean sky
levels. The photon count shot noise spike near zero lag is absent in the cross correlations
because this source of noise is uncorrelated between the different images. Absence of this
spike shows that the power law behavior of the correlations extends down to the smallest
scale that we observe. This smallest scale is 0.04′′ rather than the 0.1′′ scale of the WF
pixels, thanks to the sub-pixel resolution recovered by drizzling (see Fruchter & Hook 1997).
The close agreement between auto and cross correlations indicates that similar
structure is present in different filters, in exposures that were obtained at different times.
The relative amplitudes are consistent with a common origin for most of the fluctuations
in all three filters, with some extra signal in the B450 and I814 images that is not present
in V606 (V606 also has the highest signal-to-noise ratio). The dimensionless correlations in
Figure 2 represent fluctuations relative to the mean sky level, thus the good match between
correlations in different filters implies that the color of the fluctuating component must be
close to that of the mean sky. It is important to remember that, because the mean sky is
-- 9 --
much brighter than any possible EBL component, an instrumental systematic that varies
slowly with wavelength could cause correlated structure with the same color as the mean
sky.
We have also computed the autocorrelation function of the color of the unresolved
sky, i.e., the ratio of surface brightness in different bandpasses (note that this is the
ratio of measured surface brightness before subtracting the mean sky). Color correlations
of the B450/V606 and V606/I814 ratio images, expressed as ω(θ) = C1/2(θ)/(µ1/µ2)2,
have slightly smaller amplitude and steeper slope than the auto and cross correlations.
Color correlations should be relatively less affected by flatfield uncertainties because any
wavelength-independent flatfield structure is divided out in forming the ratio image. We
will report in detail on these results in Paper II (Vogeley 1998). An important point is
that we expect the amplitude of color correlations to be smaller than the individual filter
autocorrelations by a factor that strongly depends on the relative color of the sources of the
fluctuations and the mean sky. Here we note only that the detection of color correlation
of the same order as the auto and cross correlations argues against a dominant systematic
contribution from flatfield errors.
3.
INSTRUMENTAL SYSTEMATICS AND FOREGROUND SOURCES
Several instrumental systematics and foreground sources may contribute to the
clustering signal plotted in Figure 2. These effects include errors in calibration of the
instrument sensitivity, smearing by the point spread function (PSF) of the telescope,
wide-angle scattered light, and fluctuations in the Zodiacal and Galactic foregrounds.
Because these systematics and foregrounds only add to the measured clustering signal,
we may use the observed correlations as a firm upper limit on the true EBL fluctuations;
this is the approach that we follow in Section 4 below. If we identify an instrumental
effect that contributes to the fluctuations and subtract this contribution, then we obtain an
even stronger constraint on additional sources of the EBL. For example, we have already
been able to remove some signal by correction for obvious flatfield errors, and preliminary
analysis clearly indicates a contribution from scattering by the telescope PSF. In Paper II
we provide a detailed analysis of these possible contributions, with the goal of improving
these upper limits.
-- 10 --
3.1. HST/WFPC2 Systematics
Errors in calibration of spatial variations in the instrument sensitivity will cause
apparent fluctuations in the sky. As noted above, smoothing of the object-masked sky
revealed large-scale features in the v2 HDF images, which we removed using better flat field
calibrations. Color correlation analysis removes signal caused by wavelength-independent
flatfield errors, but this does not rule out a wavelength-dependent component. Another test
is to compare the two-dimensional autocorrelation structure of the flatfields themselves with
the detected signal. If the uncertainties in the flatfields are proportional to the flatfields
themselves, then features in their two-dimensional autocorrelations will show up in the sky
correlations. In addition to cross-correlating the flat field images with the data directly, we
can compare the multipole moments of their two-dimensional autocorrelation functions.
Preliminary analyses do not reveal cross correlations or anisotropies that are characteristic
of flat field errors.
Measurement of the multipole moments can also reveal problems caused by charge
transfer efficiency or other defects aligned with the CCD rows or columns, which cause a
quadrapole signal. Scattered light from the HST secondary mirror support spider creates
an "X" pattern that would yield a large hexadecapole moment.
The point spread function of the WFPC2 has extended wings, which probably arise
from scattering inside the cameras (Krist & Burrows 1994; Krist 1995; Burrows et al. 1995).
This PSF is anisotropic and varies with position within each camera. For our purpose, the
most important effect of the PSF wings is to scatter some flux well beyond the mask region
of an object. A crude test for this effect is to make an image in which we place a model
of the PSF at the position of each detected object, with amplitude proportional to the
measured magnitude, then mask this image as we do for the data and cross-correlate with
the masked HDF image. Preliminary results from this and similar tests (such as varying
the mask size) indicate that this effect produces correlations with shape that is similar to
the HDF correlations and could contribute as much as 10 − 40% of the observed correlation
amplitude. This may be the dominant systematic effect on our measurement. However, it
is difficult to cleanly separate this PSF effect. Similar effects would arise if the profiles of
detected galaxies extend far beyond the masked regions (i.e., if the sky is confusion-limited
in these galaxies at some very faint isophote), or if lower surface brightness objects are
clustered with the detected galaxies.
The dark count rate in the WFPC2 varies both spatially and temporally (Burrows et
al. 1995). These variations are probably caused by fluorescence of the MgF2 window on the
camera. At fixed epoch, the dark count rate is constant within the central region of each
field, but declines near the edges of the CCD. The amplitude of the dark count rate, and
-- 11 --
therefore the steepness of the roll-off, varies in time with the cosmic ray flux. To minimize
sensitivity to this effect, we restrict our analysis to the central one-quarter of the area of
each CCD.
The drizzling procedure that was used to combine the HDF exposures includes
corrections for the geometric field distortion in each camera. Restricting our analysis to the
central area of each field also minimizes sensitivity to residuals from this distortion.
3.2. Foreground Sources
The motion of the Earth relative to the source of the Zodiacal light causes fluctuations
in this foreground to be decorrelated between WFPC2 exposures. The Zodiacal light
is Solar flux that is backscattered from a layer of dust that extends to roughly 3 A.U.
(Dermott et al. 1996; Reach et al. 1996). In the time it takes for HST to orbit the Earth
(96 minutes), the Earth's motion around the Sun changes the HST's line of sight through
this dust layer. During the month of December (when the HDF was observed), the line of
sight through a screen at a distance of 3 A.U. towards a fixed extrasolar target at 12h37m
+62◦13′ (the position of the HDF) changes by ∼ 2′. The field of view of a WF camera
is ∼ 1.25′, therefore sky brightness fluctuations due to the Zodiacal foreground become
decorrelated if we examine fluctuations by cross correlating WFPC2 exposures that are
separated in time by at least the duration of one orbit. Exposures in different filters were
separated by many orbits, thus cross correlations of images in different filters should be
unaffected by ZL fluctuations. The good match of auto and cross correlations indicates that
ZL fluctuations do not strongly affect any of our measurements.
The next brightest foreground is Galactic cirrus, which is presumably the reflection
of starlight from high-latitude dust. The HDF field was chosen to lie at a minimum in
the IRAS 100µm maps, thus we expect this Galactic foreground to be much smaller than
average. Guhatakurta & Tyson (1989) measured the color of the Galactic cirrus and found
that it is ∼ 1m redder in BJ − R than either the faint blue galaxies or the structure with
surface brightness fainter than ∼ 30 mag/arcsec2 seen on scales > 6′′ by Tyson (1988). This
color is similarly too red to match the color implied by the auto and cross correlations of
the HDF sky (within a few tenths of a magnitude of the Zodiacal light color). However,
because little is know about correlations of this foreground on scales of a few arcseconds,
we cannot exclude this possibility.
-- 12 --
4. CONSTRAINTS ON SOURCES OF THE EBL
4.1. Galaxy Populations
Is there a population of sources that (1) would not have been directly detected in
the HDF, (2) would contribute significant surface brightness to the EBL, and (3) would
not overproduce fluctuations in the EBL? If we specify the distribution of fluxes, surface
brightness profiles, and angular clustering of a proposed undetected population, then we
can compute the effect of these sources on fluctuations in the EBL. Using the measured
correlations of the object-masked sky in the HDF as an upper limit on the true EBL
fluctuations, we then test whether the predicted fluctuations are allowed by the measured
sky correlations.
The autocorrelation of the surface brightness from a population of sources is the sum of
contributions from the correlation structure within the profiles and from clustering among
the sources. For a population of sources with identical apparent surface brightness profiles
µgal(θ), the autocorrelation function is
Cgalaxies(θ) = ncgal(θ) + n2 Z d2x cgal(θ − x)ωclust(x).
The first term is from the correlations of flux within the object profiles; cgal is the
convolution of a galaxy profile with itself,
cgal(θ) = Z d2x µgal(x)µgal(x + θ).
(5)
(6)
The second term is from clustering among the galaxies, with angular two-point correlation
ωclust(θ).
In this subsection we assume that ωclust = 0, so that we can constrain possible EBL
sources on the basis of their profiles alone. The addition of source clustering would increase
the predicted EBL fluctuations and strengthen the constraints on these models. In the next
subsection we examine the other extreme, a clustered diffuse EBL source.
Because it does not account for absorption of flux from one galaxy by dust in another,
equation (5) is exact only in the limit where the sources are transparent or sparsely
distributed. If galaxies have finite optical depth then we require more sources to produce
fixed optical EBL. Another effect of absorption is that clustering of sources causes a smaller
increase in the EBL fluctuations because nearby galaxies tend to shield one another. A
complete accounting for these effects requires specification of the redshift distribution,
spectral energy distribution, and extinction law for the sources. We will examine such
detailed models in Paper III.
-- 13 --
Here we consider only simple phenomenological source models, in which unclustered
sources have exponential surface brightness profiles with identical apparent central surface
brightness µ0 and apparent angular scale length h. Such a population contributes total
surface brightness
µpop = nf = Eµdetect,
(7)
where n is the projected number density, f = 2πµ0h2 is the total flux of each source, and E
is the ratio of this extra surface brightness to the surface brightness in detected galaxies to
the HDF limit, µdetect (here we define µdetect, which is the same as µgalaxies defined in section
1, to avoid confusion with the additional source population).
To compute the EBL fluctuations from this source population, we first compute the
convolution of an exponential disk with itself,
cexp(θ) = cexp(0)p(θ/h) = (cid:18)π
2
0h2(cid:19)(cid:18) 2
µ2
π Z d2u e−ue−u+θ/h(cid:19) ,
(8)
which defines a dimensionless profile function p(θ/h) with p(0) = 1. The autocorrelation of
the population is
Cpop(θ) = n
π
2
µ2
0h2p(θ/h) =
nf 2
8πh2 p(θ/h).
(9)
The power law fits to the HDF correlations (Table 1) yield a constraint on the EBL
fluctuations caused by this population,
Cpop(θ) < Cobs(1′′)θγ
(10)
for all θ. Combining the equations above, we derive a constraint on the the V606 band
central surface brightness µ0 and angular scale length h of these sources,
µ0 < 6.3 × 10−22(h/1′′)−0.6E−1 erg/s/cm2/sr/Hz.
(11)
If we define an effective area πh2 for each source, then the covering factor (the average
number of objects along a random line of sight) is of order
χ = nπh2 =
Eµdetect
2µ0
.
(12)
Figure 3 shows the resulting constraints on the V606 band central surface brightness µ0
and scale size h of a population of identical unclustered exponential disks that contributes
mean surface brightness equal to that of detected galaxies (E = 1). The region below
the solid line satisfies the constraint of equation (11). Above this line the sources'
-- 14 --
autocorrelation function would exceed this upper limit for some range of θ. The right-hand
axis indicates the covering factor χ for different µ0. Above the dashed line at the top of this
figure lie objects directly detected in the HDF using FOCAS with an isophotal threshold of
∼ 27 mag/arcsec2. For a smaller contribution to the EBL, i.e., for E < 1, we relax the µ0
limit by 2.5 log E and change the covering factor by a factor E−1. In the limit E ∼< 0.1, the
allowed region abuts the detected region.
Two examples of populations that contribute µpop = µdetect and marginally satisfy
the correlation constraint illustrate how very different such a population would be from
the HDF detections. For h = 0.1′′ a population of extremely faint, total magnitude
V606 = 32.4, sources with µ0 = 29.4 mag/arcsec2 is marginally allowed. The number of
such objects, n = 1.3 × 109deg−2, is ∼ 600 times the total number per deg2 of detected
galaxies. For h = 1.0′′, the correlations allow a population of V606 = 28.9mag sources,
with µ0 = 30.9 mag/arcsec2. For comparison, the typical scale size of V606 = 29mag
detected galaxies is ∼< 0.2′′. The projected number density of the additional V606 = 28.9mag
population would be 20 times the total number of detected objects. On the righthand axis
of this plot we note the covering factor that corresponds to each choice of µ0 and h. For
the former example, this covering factor is χ ∼ 3, for the latter χ ∼ 13. Such large covering
factors imply that these objects might be detected as QSO absorption line systems and/or
from reddening of background objects.
For comparison to these models, note that the most extended low surface brightness
galaxy seen to date, Malin 1, has an extrapolated disk central surface brightness of
µ0 = 26 mag/arcsec2 in V and scale size h = 82h−1
75 kpc (Bothun et al. 1987). At z = 0.5
this would be dimmed to roughly 28 mag/arcsec2 in I and would have an apparent scale
size h = 17′′. We obtain relatively poor constraints on sources with h ∼> 10′′ because the
integral constraint bias is comparable to the correlation function on these scales (see Table
1); the 41′′ × 41′′ field that we analyze is too small to accurately measure larger scale
fluctuations. However, for E = 1, objects like Malin 1 clearly would be ruled out in Fig. 3
for a reasonable extrapolation of the measured correlation function.
Thus, to make a large contribution to the EBL, the correlation constraint requires that
this additional population and the detected galaxies form an extremely bimodal distribution
in surface brightness. The undetected sources must have very low central surface brightness
and be confusion-limited on the sky. Some galaxies certainly lie within the region between
the FOCAS detections and the "allowed" region for extra sources of the EBL. As shown
in Section 1, extrapolation of the number counts would place some objects here but they
would not contribute very much surface brightness to the EBL.
Extrapolation of the detected galaxy counts adds little to the EBL, but would these
-- 15 --
fainter galaxies cause detectable correlations? No. Following equation (9), a population of
galaxies fainter than mlim with number per magnitude distribution N(m) = dn/dm and
identical angular scale size h has zero-lag autocorrelation signal
Cpop(0) =
1
8πh2 Z ∞
mlim
dm N(m)h10−0.8(m+48.60)i (erg/s/cm2/sr/Hz)2.
(13)
If we extrapolate the HDF V-band counts from V = 29.5mag to zero flux with logarithmic
count slope α = 0.2 and scale size h = 0.2′′, then C(0)pop = 2.2 × 10−43(erg/s/cm2/sr/Hz)2,
which translates to a dimensionless correlation ωpop(0) = Cpop(0)/µ2 = 2.2 × 10−7.
Comparison to Figure 2 shows that this signal is two orders of magnitude below the
measured correlations. Thus, although galaxies brighter than the HDF detection limit
would dominate the clustering signal (see section 2), galaxies just fainter than this limit
have little effect on the observed correlations.
4.2. Weakly Clustered Diffuse Light
So far, we have ignored clustering among the undetected objects. If the undetected
sources are clustered, then the solid line in Figure 3 moves down towards even lower allowed
central surface brightness. In the limit of point-like objects, the autocorrelation function of
a population (eq.[5]) reduces to
Cpop(θ) = nf 2δ(θ) + n2f 2ωclust(θ).
(14)
This must be convolved with the PSF to determine the observed autocorrelation. If we set
nf = Eµdetect then we obtain a simple constraint on the 1′′ angular clustering amplitude of
the sources,
ωclust(1′′) <
Cobs(1′′)
(Eµdetect)2 .
(15)
This constraint applies for all scales if we make the additional assumption that these objects
have power law angular correlations, ωpop(θ) ∝ θγ with γ ∼ −0.6, similar to that fit to the
HDF sky correlations.
On scales larger than the PSF size, the inequality in equation (15) is equivalent to a
constraint on a clustered diffuse EBL component with µdif f use = Eµdetect. Figure 4 shows
the constraints on combinations of µdif f use and ωdif f use(1′′) from the HDF correlations. For
E = 1, the upper bound on angular clustering of diffuse light is ωdif f use(1′′) < 8 × 10−3.
This 1′′ clustering amplitude is an order of magnitude smaller than the clustering measured
for galaxies as faint as I < 29 in the HDF (Colley et al. 1996; Villumsen et al. 1997). Thus,
to contribute as much surface brightness as the detected galaxies, these undetected objects
-- 16 --
must be not only extremely faint and numerous, but also uniformly distributed to exquisite
precision.
Apart from imagined sources of additional EBL, several observations of galaxy clusters
detect diffuse light or intergalactic stars that would contribute to the amplitude and
clustering of the EBL if similar surface brightness is associated with most galaxies. From R
band imaging of Abell 2029, Uson, Boughn, & Kuhn (1991) infer that 10% of the light in
this cluster is in a diffuse component. Using WFPC2 imaging of the Virgo cluster, Ferguson,
Tanvir, & Von Hippel (1996) find a population of intergalactic stars that could contribute
total flux equal to 10% of that in galaxies. Theuns & Warren (1996) detect candidate
planetary nebulae in Fornax and infer an intergalactic stellar population that contributes as
much as 40% of the cluster light. M´endez et al. (1997) also find candidates for intergalactic
planetary nebulae in Virgo. Tyson, Kochanski, & Dell'Antonio (1997) detect diffuse light
in CL0024+1654 that comprises 15% of the light in the central 100kpc of this cluster.
If diffuse intergalactic light makes a fractional contribution to the EBL (relative to
µdetect from galaxy counts) that is similar to the intergalactic surface brightness detected
in clusters, then it could cause EBL fluctuations close to those measured in the HDF
sky. The constraints on diffuse intergalactic light also follow from equation (15) if the
diffuse light is clustered in similar fashion to faint galaxies, with a power law correlation
function Cdif f (θ) ∼ µ2
dif f θ−0.6. Let us suppose that the clustering amplitude of the diffuse
light is comparable to that of faint galaxies. The small arrow in Figure 4 indicates the
level of diffuse light that contributes an additional µdif f use = 0.1µdetect (i.e., similar to
the 10% contribution of intergalactic stars in Virgo). The EBL correlations from this
source would match the HDF correlations if this light has the same clustering amplitude
as R < 23mag galaxies, ω(1′′) ∼ 1 (Couch, Jurcevic, & Boyle 1993). Because 75% of the
light from detected galaxies comes from I814 < 23 galaxies, it seems plausible that much of
the diffuse light would be associated with galaxies in this same magnitude interval, and be
gravitationally clustered with similar amplitude. Note that the contribution of any such
a source to the EBL fluctuations would leave even less room in the upper limits on sky
correlations for a discrete source population.
4.3.
Is the HDF Typical?
A remaining question is whether the HDF is a typical field for the purpose of probing
fluctuations in the EBL. The answer depends on the redshift and angular size of the
hypothesized sources of undetected EBL. The 80′′ × 80′′ field of view of each WF camera
limits the utility of the HDF for studying fluctuations on scales much larger than 10′′.
-- 17 --
This angular coverage also limits the range of apparent magnitude of the galaxies that
the field includes. Because this observation was designed to be an unbiased probe of the
high-redshift universe, the field was deliberately chosen to avoid bright foreground galaxies
(V < 22mag) which would fill much of the WFPC2 field of view.
This avoidance of bright galaxies raises the concern that the HDF is biased against
detection of EBL fluctuations from previously undetected sources at redshift z ∼< 0.3 that
are clustered with bright galaxies. [A similar bias arises in the direct EBL detection method
of Bernstein et al., because they use WFPC2 imaging to measure the total sky brightness.]
However, it seems extremely contrived to envisage a population of sources that is not
detected at z > 0.3 through fluctuation analysis, is not directly detected at z < 0.3 in deep
ground-based imaging, and yet contributes a large fraction of the optical background.
Another issue is variation of EBL fluctuations from field to field. The total imaged area
of the three WF cameras is 5.3 arcmin2, of which we analyze only the central 1.3 arcmin2. If
additional sources of EBL are distributed similarly to the detected galaxies, then expected
fluctuations in galaxy counts provides a rough estimate of uncertainty in the mean EBL
over this area. The Poisson fluctuation in counts of galaxies with magnitude 26 < V < 29.5
would be 5% over a 1.3 arcmin2 area. For the magnitude range 23 < V < 26, the expected
fluctuation rises to 12%. Galaxy clustering does not significantly increase this lower bound
on the field-to-field fluctuations; although the field of view spans only a few hundred kpc
in angle, these magnitude intervals include galaxies over many hundreds of comoving Mpc
in distance. We conclude that the field-to-field variation in the amplitude of the HDF
correlations is likely to be caused by systematic effects rather than true fluctuations in the
EBL.
5. CONCLUSIONS
Fluctuations in the object-masked sky in the HDF provide a strong, albeit indirect,
constraint on the mean optical EBL. If the HDF provides a fair probe of the extragalactic
sky between detected galaxies, then we conclude that this observation and other deep
imaging surveys have already detected the sources of the majority of the optical EBL.
The mean optical EBL is, at most, a few tens of percent above the mean surface
brightness from detected galaxies. At λ = 8100A, a generous allowance for up to 50%
additional EBL suggests 7.8 × 10−6 ≤ νIν < 1.2 × 10−5erg/s/cm2/sr. This mean level
lies at the lower end of the uncertainty range of the measurement by Bernstein et al.
(νIν = 2.1 ± 1.2 × 10−5erg/s/cm2/sr, after including V < 23mag galaxies) and well below
previous upper limits.
-- 18 --
We infer this mean EBL level because no plausible sources of additional extragalactic
surface brightness can satisfy the multiple constraints that these sources (1) have not
already been detected, (2) contribute total surface brightness comparable to the surface
brightness contributed by detected galaxies, and (3) do not produce EBL fluctuations
larger than the upper limits set by correlations in the object-masked HDF. Figure 3 shows
that these constraints admit only extremely low surface brightness objects that must be so
extended and numerous as to be confusion-limited on the sky. The parameters of simple
phenomenological models that meet these constraints would make such a population disjoint
from the parameter space of all detected objects. Extrapolation of the HDF galaxy number
counts to infinitely faint limits would add at most a few percent to the mean EBL. Figure
4 shows that a diffuse component that is as large as the surface brightness from detected
galaxies would require a V -band angular clustering amplitude of ω(1′′) < 8 × 10−3, an order
of magnitude less clustered than the faintest observed galaxies. Fluctuations in the EBL
from intergalactic stars as seen in galaxy clusters could be as large as the upper limits on
the fluctuations in the HDF, but would contribute only incrementally to the mean EBL.
In addition to constraining the plausible level of the optical EBL, these upper limits
on small-scale sky fluctuations in the HDF provide a test for any proposed model for faint
galaxy populations. We plan to test galaxy population models that are more detailed than
the simple phenomenological model described in Section 4 (Paper III). We invite others to
suggest possible models, compute the predicted autocorrelation function of sources below
the detection limits, and compare with these limits from the HDF. These correlations are
also important for attempts to detect weak lensing signals in deep optical imaging (Van
Waerbeke et al. 1997; Refregier et al. 1997).
As we emphasize throughout this paper, the measured correlations of the object-masked
sky in the HDF should be treated as upper limits on the true EBL fluctuations. The
wings of the WFPC2 point spread function almost certainly contribute to the measured
fluctuations. Although there are reasons to exclude flat fielding errors or Galactic light as
the principal sources of the fluctuations, we cannot rule out some contribution from these
effects. To improve these upper limits, we are examining these systematics and foregrounds
in further detail (Paper II).
The HDF provides the best constraint to date on small-scale EBL fluctuations
because its superior resolution allows detection of faint galaxy populations that would be
confusion-limited in ground-based imaging. In section 2, we discuss how fluctuation analysis
of shallower HST or ground-based images would be dominated on small scales by the
profiles of galaxies that only the HDF detects. However, ground-based imaging is required
to study fluctuations on angular scales much larger than several arcseconds, where access
-- 19 --
to larger collecting area and accurate control of systematics outweigh the HST's advantages
of high resolution and lower sky background. If we measure object-masked sky correlations
in deep ground-based images, we can now use the HDF galaxy counts and profiles to model
and subtract the surface brightness correlations that are caused by galaxies that are fainter
than the ground-based detection limits, but that were detected by the HDF. Fluctuation
analysis of deep ground-based imaging will allow statistical tests for diffuse intergalactic
light on larger angular scales and for low surface brightness companions to galaxies that
were too bright to be included in the HDF (see section 4c).
Future HST observations will allow measurement of EBL fluctuations in independent
fields and at fainter surface brightness levels than the HDF. The planned HDF South
(October 1998) will reach to similar depth and allow us to test if the HDF North is typical.
The Advanced Camera for Surveys (scheduled for installation on HST in 1999) will have
superior sensitivity and better-controlled systematic uncertainties than WFPC2, as well as
a 200′′ × 200′′ contiguous field of view, thus allowing measurement of surface brightness
fluctuations at fainter levels and somewhat larger scales. The proposed GTO program for
ACS focuses on deep imaging of galaxy clusters. Fluctuation analysis of the ACS cluster
fields will be an important test for sources of diffuse intergalactic surface brightness. In the
2 − 10µm infrared, the proposed Next Generation Space Telescope might detect statistical
fluctuations in the EBL from the first generation of stars (e.g., Haiman & Loeb 1997) even
if such objects evade direct detection.
We thank R. Williams and the entire HDF team, whose vision and effort made this
project possible. We acknowledge discussions with and many helpful comments from A.
Babul, R. Bernstein, S. Casertano, A. Connolly, H. Ferguson, A. Fruchter, M. Geller,
E. Groth, R. Lupton, P. Madau, J. Peebles, A. Refregier, P. Stockman, M. Strauss, A.
Szalay, T. Tyson, and D. Wilkinson. We also thank R. Bernstein and L. Pozzetti for
providing results in advance of publication. This paper was initiated during a visit to the
Aspen Center for Physics. Support for this work was provided by NASA through grants
HF-01078.01-94A and AR-05812.01-94A from the Space Telescope Science Institute, which
is operated by AURA, Inc. under NASA contract NAS5-26555,
REFERENCES
Barcons, X., & Fabian, A. C. 1989, MNRAS, 237, 119
Bernstein, R. A. 1997, Ph.D. thesis, California Institute of Technology
Bernstein, R. A., Freedman, W. L., & Madore, B. F. 1998, in preparation
-- 20 --
Bertin, E., & Arnouts, S. 1996, A&A Suppl., 117, 393
Bothun, G. D., Impey, C. D., Malin, D. F., & Mould, J. R. 1987, AJ, 94, 23
Boughn, S. P., & Kuhn, J. R. 1986, ApJ, 309, 33
Boughn, S. P., Saulson, P. R., & Uson, J. M. 1986, ApJ, 301, 17
Burrows, C. J., et al. 1995, Wide Field and Planetary Camera 2 Instrument Handbook,
STScI publication
Cole, S., Treyer, M., & Silk, J. 1992, ApJ, 385, 9
Colley, W. N., Rhoads, J. E., Ostriker, J. P., & Spergel, D. N. 1996, ApJ, 43, 63
Condon, J. J. 1974, ApJ, 188, 279
Connolly, A. J., Szalay, A. S., Dickinson, M., Subbarao, M. U., & Brunner, R. J. 1997, ApJ,
486, L11
Couch, W. J., Jurcevic, J. S., & Boyle, B. J. 1993, MNRAS, 260, 241
Dalcanton, J. J. 1997, ApJ, in press
Dermott, S. F., Jayaraman, S., Xu, Y. L., Grogan, K., & Gustafson, A. S. 1996, in Unveiling
the Cosmic Infrared Background, ed. E. Dwek, AIP Conf. Proc. 348, 25
Disney, M. J. 1976, Nature, 263, 573
Djorgovski, S., Soifer, B. T., Pahre, M. A., Larkin, J. E., Smith, J. D., Neugebauer, G.,
Smail, I. Matthews, K., Hogg, D. W., Blandford, R. D., Cohen, J., Harrison, W., &
Nelson, J. 1995, ApJ, 438, L13
Dube, R. R., Wickes, W. C., & Wilkinson, D. T. 1977, ApJ, 215, L51
Dube, R. R., Wickes, W. C., & Wilkinson, D. T. 1979, ApJ, 232, 333
Fall, S. M., Charlot, S. & Pei, Y. C. 1996, ApJ, 464, L1
Ferguson, H., Tanvir, N. & Von Hippel, T. 1996, BAAS, 189, 80.06
Fomalont, E. B., Kellerman, K. I., Anderson, M. C., Weistrop, D., Wall, J. V., Windhorst,
R. A., & Kristian, J. A. 1988, AJ, 96, 1187
Fruchter, A. S., & Hook, R. N. 1997, in Applications of Digital Image Processing XX, ed.
A. Tescher, Proc. S.P.I.E. vol 3164, in press (preprint astro-ph/9708242)
Guhathakurta, P., & Tyson, J. A. 1989, ApJ, 346, 773
Gunn, J. E. 1965, Ph.D. thesis, California Institute of Technology
Haiman, Z., & Loeb, A. 1997, ApJ, 483, 21
-- 21 --
Hasinger, G., Burg, R., Giacconi, R., Hartner, G., Schmidt, M., Truemper, J., & Zamorani,
G. 1993, A&A, 275, 1; erratum: 291, 348
Hauser, M. G. 1996, in Unveiling the Cosmic Infrared Background, ed. E. Dwek, AIP Conf.
Proc. 348, 11
Hewish, A. 1961, MNRAS, 123, 167
Kashlinsky, A., Mather, J. C., & Odenwald, S. 1997, ApJ, 473, L9
Krist, J. E. 1995, in Calibrating the Hubble Space Telescope: Post Servicing Mission, eds.
A. Koratkar & C. Leitherer, 311
Krist, J. E., & Burrows, C. J. 1994, WFPC2 Instrument Science Report 94-01, STScI
Lillie, C. F. 1968, Ph.D. thesis, University of Wisconsin
Madau, P., Ferguson, H. C., Dickinson, M. E., Giavalisco, M., Steidel, C. C., & Fruchter,
A. 1996, MNRAS, 283, 1388
Madau, P., Pozzetti, L., & Dickinson, M. 1997, ApJ, submitted (preprint astro-ph/9708218)
Martin, C., & Bowyer, S. 1989, ApJ, 338, 677
Mattila, K. 1990, in The Galactic and Extragalactic Background Radiation, Proc. IAU
Symp. 139, eds. S. Bowyer & C. Leinert (Dordrecht: Kluwer), 257
McGaugh, S. S., Bothun, G. D., & Schombert, J. M. 1995, AJ, 110, 573
M´endez, R. H., Guerrero, M. A., Freeman, K. C., Arnaboldi, M., Kudritzki, R. P., Hopp,
U., Capaccioli, M., & Ford, H. 1997, ApJ, in press (preprint astro-ph/9710179)
Milliard, B., Donas, J., Laget, M., Armand, C., & Vuillemin, A. 1992, A&A, 257, 24
Oke, J. B., & Gunn, J. E. 1983, ApJ, 266, 713
Paresce, F. 1990, in The Galactic and Extragalactic Background Radiation, Proc. IAU
Symp. 139, eds. S. Bowyer & C. Leinert (Dordrecht: Kluwer), 307
Partridge, B., & Peebles, P. J. E. 1967a, ApJ, 147, 868
Partridge, B., & Peebles, P. J. E. 1967b, ApJ, 148, 377
Persic, M., de Zotti, G., Boldt, E. A., Marshall, F. E., Danese, L., Francheschini, A., &
Palumbo, G. G. C. 1989, ApJ, 336, L47
Pozzetti, L., Madau, P., Zamorani, G., Ferguson, H. C., & Bruzual, G. A. 1997, MNRAS,
submitted
Reach, W. T., Franz, B. A., Kelsall, T., & Weiland, J. L. 1996, in Unveiling the Cosmic
Infrared Background, ed. E. Dwek, AIP Conf. Proc. 348, 37
Refregier, A., et al. 1997, in preparation
-- 22 --
Roach, F. E., & Smith, L. L. 1968, Geophys. J., 15, 227
Scheuer, P. A. G. 1957, Proc. Camb. Phil. Soc., 53, 764
Shectman, S. A. 1973, ApJ, 179, 681
Shectman, S. A. 1974, ApJ, 188, 233
Spinrad, H., & Stone, R. P. S. 1978, ApJ, 226, 609
Toller, G. N. 1983, ApJ, 266, L79
Theuns, T. & Warren, S. J. 1996, MNRAS, 284, 11
Tyson, J. A. 1988, AJ, 96, 1
Tyson, J. A. 1995, in Extragalactic Background Radiation, eds. D. Calzetti, M. Livio, & P.
Madau (Cambridge University Press), 103
Tyson, J. A., Kochanski, & Dell'Antonio, I. 1997, ApJ, submitted
Uson, J. M., Boughn, S. P., & Kuhn, J. R. 1991, ApJ, 369, 46
Van Waerbeke, L., Mellier, Y., Schneider, P., Fort, B., & Mathez, G. 1997, A&A, 317, 303
Vaisanen, P. 1996, A&A, 315, 21
Villumsen, J. V., Freudling, W., & da Costa, L. N. 1997, ApJ, 481, 578
Vogeley, M. S. 1998, in preparation (Papers II, III)
Whitrow, G. J., & Yallop, B. D. 1965, MNRAS, 130, 31
Williams, R. E., Blacker, B., Dickinson, M., Dixon, W. V. D., Ferguson, H. C., Fruchter, A.
S., Giavalisco, M., Gilliland, R. L., Heyer, I., Katsanis, R., Levay, Z., Lucas, R. A.,
McElroy, D. B., Petro, L., Postman, M., Adorf, H. M., & Hook, R. N. 1996, AJ, 112,
1335
This preprint was prepared with the AAS LATEX macros v4.0.
-- 23 --
Table 1: Autocorrelation and Cross Correlation Power Law Fits a
Filter
B450
V606
I814
B450 × V606
V606 × I814
ω(1′′)
γ
3.58 (±0.69) × 10−6 −0.80 (±0.16)
1.99 (±0.38) × 10−6 −0.61 (±0.14)
2.99 (±0.59) × 10−6 −0.59 (±0.13)
1.96 (±0.39) × 10−6 −0.62 (±0.15)
2.14 (±0.46) × 10−6 −0.53 (±0.14)
µ b
6.62 × 10−19
1.00 × 10−18
1.39 × 10−18
...
...
∆ c
4.14 × 10−7
3.67 × 10−7
5.84 × 10−7
3.53 × 10−7
4.90 × 10−7
aThe observed correlation function over 0.15′′ < θ < 8′′ is fit by C(θ) = µ2ω(1′′)(θ/1′′)γ.
bMean of object-masked sky in erg/s/cm2/sr/Hz. Note that 10−18 erg/s/cm2/sr/Hz = 22.94 AB mag/arcsec2.
cIntegral constraint bias for a 41′′ × 41′′ field, estimated using these fits (See eq.[3]).
-- 24 --
0.4
0.6
0.8
0
Fig. 1. -- Observational constraints on the mean EBL. Inset box replots the dotted region
(optical limits only) using linear axes. Solid symbols are lower limits from the average
surface brightness contributed by galaxies detected in deep imaging surveys: the HDF, other
optical surveys, and K-band (solid circles; Pozzetti et al. 1997), UV (square; Milliard et
al. 1992). Arrows indicate upper limits on the extragalactic contribution to the night sky
brightness in the UV from Paresce (1990), in the optical from Toller (1983) and Dube et al.
(1977, 1979), and in the IR from Hauser (1996) (left to right, respectively). Open circles
indicate a detection of the mean EBL (Bernstein 1997) with 1σ uncertainties. The detection
level includes a small adjustment for flux from bright galaxies (V < 23mag) that were
excluded from this measurement. The surface brightness in AB magnitudes that corresponds
to νIν = 10−5 erg/s/cm2/sr at λ = 5500A is 27.3 mag/arcsec2.
-- 25 --
Fig. 2. -- Autocorrelations of "sky" pixels in the central 41′′ × 41′′ of the HDF images in
the B450, V606, and I814 bandpasses, and cross correlations of B450 × V606 and V606 × I814.
These curves are averages over the WF2, 3, and 4 fields. Error bars on the V606 curve are
from the variance among the fields (which dominates this uncertainty) and within bins of
width δ(log10 θ) = 0.4. Other curves have similar uncertainty. Shot noise dominates the
autocorrelations for θ < 0.1′′. Both auto and cross correlations are well-fitted by power laws
from 0.15′′ to 8′′ (see Table 1 for fitting parameters). The drop-off at large θ is an artifact
of the integral constraint bias on the correlation function for a finite area of sky. The heavy
solid bar at 4 × 10−7 marks the typical amount by which we underestimate the correlations
due to this effect. Cross correlation removes the shot noise and confirms that roughly the
same fluctuations are seen in different filters. Several sources of instrumental systematics
and astronomical foregrounds may contribute to this measured correlation signal. Thus, we
treat this measurement as an upper limit on small-scale fluctuations in the EBL.
-- 26 --
Fig. 3. -- Correlation constraints on a population of identical exponential disks, assuming no
clustering among the objects. Using the power-law correlation signal in Figure 2 as an upper
limit on the true fluctuations, here we explore the allowed range of h and µ0 in the V606 band.
The surface density of objects is set so that the EBL contributed by these sources is equal
to that in detected galaxies. A population with central surface brightness µ0 and scale size
h that lies above the solid line would produce correlations larger than observed. Below this
line, the correlation constraint is satisfied, but the covering factor on the sky (right-hand
axis) is quite large; any population with µ0 > 29.5mag/arcsec2 would completely cover the
sky > 3 times. Indicated are the total magnitudes and surface density of two marginally-
allowed populations. The region above the dashed line is the locus of galaxies detected to the
limits of the HDF. Some galaxies may exist between the dashed box and solid line without
causing excessive EBL correlations; they simply cannot contribute very much total surface
brightness to the EBL.
-- 27 --
Fig. 4. -- Constraints on clustered diffuse light with mean surface brightness µ and 1′′
clustering amplitude ω(1′′), assuming that the angular correlation function of the diffuse light
has the same slope as the measured ω(θ). The allowed region in the µ−ω(1′′) plane lies below
the solid line. To contribute surface brightness equal to that in detected galaxies (dashed
line), the diffuse light would have to be nearly uniform. A small fractional contribution to
the EBL, similar to the fraction of diffuse light in clusters, might be allowed if this light
has angular clustering similar to faint galaxies (note arrows at bottom of this figure). The
observed correlations marginally allow diffuse light with surface brightness that is ∼ 10%
that of detected galaxies if the correlation amplitude is similar to R < 23mag galaxies. Note
that 75% of the resolved EBL comes from galaxies with I ∼< 23mag. Such an EBL component
would account for the observed fluctuations but would add only incrementally to the total
mean EBL.
|
0803.0078 | 1 | 0803 | 2008-03-01T19:23:07 | Cluster Multi-spacecraft Determination of AKR Angular Beaming | [
"astro-ph"
] | Simultaneous observations of AKR emission using the four-spacecraft Cluster array were used to make the first direct measurements of the angular beaming patterns of individual bursts. By comparing the spacecraft locations and AKR burst locations, the angular beaming pattern was found to be narrowly confined to a plane containing the magnetic field vector at the source and tangent to a circle of constant latitude. Most rays paths are confined within 15 deg of this tangent plane, consistent with numerical simulations of AKR k-vector orientation at maximum growth rate. The emission is also strongly directed upward in the tangent plane, which we interpret as refraction of the rays as they leave the auroral cavity. The narrow beaming pattern implies that an observer located above the polar cap can detect AKR emission only from a small fraction of the auroral oval at a given location. This has important consequences for interpreting AKR visibility at a given location. It also helps re-interpret previously published Cluster VLBI studies of AKR source locations, which are now seen to be only a subset of all possible source locations. These observations are inconsistent with either filled or hollow cone beaming models. | astro-ph | astro-ph |
GEOPHYSICAL RESEARCH LETTERS, VOL. ???, XXXX, DOI:10.1029/,
Cluster Multi-spacecraft Determination of AKR Angular
Beaming
R. L. Mutel1 I. W. Christopher1 J .S. Pickett1
Abstract. Simultaneous observations of AKR emission
using the four-spacecraft Cluster array were used to make
the first direct measurements of the angular beaming pat-
terns of individual bursts. By comparing the spacecraft loca-
tions and AKR burst locations, the angular beaming pattern
was found to be narrowly confined to a plane containing the
magnetic field vector at the source and tangent to a circle
of constant latitude. Most rays paths are confined within
15◦ of this tangent plane, consistent with numerical simula-
tions of AKR k-vector orientation at maximum growth rate.
The emission is also strongly directed upward in the tan-
gent plane, which we interpret as refraction of the rays as
they leave the auroral cavity. The narrow beaming pattern
implies that an observer located above the polar cap can de-
tect AKR emission only from a small fraction of the auroral
oval at a given location. This has important consequences
for interpreting AKR visibility at a given location. It also
helps re-interpret previously published Cluster VLBI stud-
ies of AKR source locations, which are now seen to be only
a subset of all possible source locations. These observations
are inconsistent with either filled or hollow cone beaming
models.
1. Introduction
Determining the angular beaming characteristics of ter-
restrial auroral kilometric radiation (AKR) has been an im-
portant focus of AKR investigations since its discovery more
than forty years ago. Gurnett [1974] made the first analysis
of AKR beaming using one year of Imp-6 and Imp-8 satellite
observations of AKR emission. He found that the statisti-
cal power pattern formed a 'distinct cone-shaped bound-
ary' at a large angle to the magnetic field direction. Sub-
sequent statistical analyses of AKR radiation patterns were
fitted to either frequency-dependent filled cone [Green et al.,
1977; Green and Gallagher , 1985] or hollow cone [Calvert,
1981b, 1987] beaming models, the latter suggested by stud-
ies of Jovian decametric emission [e.g., Dulk , 1970], whose
properties are consistent with a very thin (∼ 1◦ ) hollow
cone. More recently, Kasaba et al. [1997] analyzed 38 months
of AKR bursts recorded on the GEOTAIL spacecraft. They
could not distinguish between hollow and filled cone models,
but found that the AKR illumination pattern was systemat-
ically modified both by geomagnetic activity and by season.
These studies provide statistical descriptions of the over-
all sky pattern illuminated by AKR bursts, but they do
not address the angular beaming pattern of
individual
1Dept. Physics and Astronomy, University of Iowa, Iowa
City IA 52242, USA.
Copyright 2018 by the American Geophysical Union.
0094-8276/18/$5.00
1
AKR bursts. AKR emission is thought to consist of a
large number of 'elementary radiation sources' with nar-
row frequency-time structures [Gurnett and Anderson, 1981;
Pottelette et al., 2001; Mutel et al., 2006]. Hence,
in-
struments with inadequate frequency and/or time resolu-
tion (e.g.
swept-frequency receivers) receive the sum of
contributions from many spatially separated elementary
sources whose frequency-time structures lie within the time-
frequency resolution window of the instrument. Also, since
most beaming studies have utilized observations from sin-
gle satellites, the observed power pattern on the sky is a
statistical ensemble of beams from many individual radiat-
ing sources rather than the angular beaming pattern of an
isolated source.
The wideband (WBD) instrument on the four Cluster
spacecraft provides a unique opportunity to determine the
angular power pattern of individual AKR bursts for the first
time. The WBD system records the received waveform di-
rectly, so that it has adequate time and frequency resolu-
tion to isolate emission from individual elementary AKR
sources. In addition, since the Cluster constellation forms a
2-dimensional array on the sky, it can simultaneously sample
the burst over a range of solid angles. Finally, by measur-
ing the differential delays between pairs of spacecraft, the
locations of individual AKR bursts can also be determined
[Mutel et al., 2003]. This provides a spatial filter which iso-
lates radiation from a single region and allow the array to
sample individual burst power patterns.
2. AKR Beaming: Model predictions
The electron cyclotron maser instability (CMI) is widely
believed to be the mechanism responsible for auroral kilo-
metric radio emission [e.g., Treumann, 2006]. AKR sources
are found in thin, low density cavities in the upward current
region above the auroral zones [Calvert, 1981a; Ergun et al.,
1998] . The cavities are generally oriented tangent to the
auroral oval and aligned with the magnetic field. They have
small latitudinal widths (10 km -100 km) compared with
their longitudinal extent, which can extend several thousand
km.
The radiation is generated most efficiently in the internal
extraordinary (X) mode close to the electron gyro-frequency.
The growth rate is strongly peaked for propagation nearly
perpendicular to the magnetic field, as seen in both in situ
observations [Roux et al., 1993; Ergun et al., 1998] and in
model calculations [Pritchett et al., 2002; Mutel et al., 2007]
. The internal X-mode cannot propagate directly, but can be
converted to external right circularly polarized X-mode radi-
ation after upward propagation along the tangent plane, as
first noted by Louarn and Le Qu´eau [1996]. Pritchett et al.
[2002], using a 2-dimensional simulation of AKR emission in
a thin cavity, also found that the emission is highly beamed
in the 'along track' or longitudinal direction. This prediction
is strengthened by FAST in situ observations in the upward
current region which found that the tangential power is 100
times stronger than the perpendicular spectrum [Pritchett
et al., 2002].
X - 2
MUTEL ET AL.: ANGULAR BEAMING OF AKR
Figure 1. (a) Geometry of the tangent plane beaming
model. The tangent plane cavity, shown in blue, is tan-
gent to a circle of constant latitude at the source. The
green lines are escaping ray paths, the black line is the
magnetic field, and yellow arrows are the xyz coordinates
system described in the text. (b) Refraction angle ver-
sus horizontal distance for rays at 125 kHz (red), 250
kHz (green), and 500 (blue) kHz initially perpendicular
to the magnetic field. Ray paths are shown in the lower
right inset.
3. Geometry of beamed AKR emission
We used the wideband plasma wave instrument (WBD)
[Gurnett et al., 2001] on the four Cluster spacecraft in VLBI
mode to determine the locations of individual AKR bursts
for more than 12,000 individual AKR bursts over 39 epochs.
The VLBI technique solves for AKR burst locations using
measurements of differential delays between all six pairs of
spacecraft [Mutel et al., 2003]. The position uncertainty
is typically 500-1000 km in the plane perpendicular to the
source-observer line of sight.
Once a source position is determined, we calculate the
tangent plane coordinates of each spacecraft as seen from
the source location. This is done by constructing an or-
thonormal 3-dimensional coordinate system with origin at
the source, (cid:126)x aligned outward along the magnetic field di-
rection, (cid:126)z in the meridian plane pointed toward the local
magnetic pole, and (cid:126)y = (cid:126)z × (cid:126)x. The tangent plane latitude
is the angle between the spacecraft vector and the tangent
(xy) plane, while the longitude is the angle between the
spacecraft vector projected onto the the tangent plane and
(cid:126)x. The coordinate geometry is illustrated in Fig. 1a.
Figure 2. (a) Cluster spacecraft SC4 tangent plane lati-
tude distribution at 125 kHz (red), 250 kHz (green), and
500 kHz (blue) determined from 12,000 AKR burst lo-
cations. (b) Same as (a), but tangent plane co-longitude
distribution.
3.1. Confinement to tangent plane
Once a source location is determined, we calculate the
tangent plane coordinates of the Cluster spacecraft at the
time of reception of the individual burst. Histograms of the
spacecraft tangent plane latitudes and longitudes for 12,000
individual AKR bursts are shown in Fig. 2. The latitude
distribution is strongly peaked near 0◦ , with more than
98% of solutions within 20◦ of the origin, and appears to
be frequency-independent. The longitude distribution has
been folded about the origin, since the tangent plane model
is symmetric to reflection about the (cid:126)x axis. The histogram
for all three frequencies is peaked near the origin, indicating
that the ray paths are strongly refracted.
3.2. Refractive effects
As radiation leaves the auroral cavity, it is subject to re-
fraction by the denser magnetospheric plasma. Detailed cal-
culations of AKR propagation in the Earth's magnetosphere
[Calvert, 1987; Gaelzer et al., 1994; Green, 1988; Schreiber
et al., 2002; Burinskaya and Rauch, 2007; Xiao et al., 2007]
predict upward refraction and/or reflection from the plasma-
pause. Hence, the 'tangent plane' model suggests that AKR
emission is confined to a plane containing the magnetic field
vector at the source, and is refracted upward (Fig. 1b).
In this paper we estimate the expected refraction outside
the auroral cavity with a simple ray tracing calculation using
the cold plasma dispersion equation. Rays at 125, 250, and
500 kHz are launched perpendicular to the magnetic field
at magnetic latitude 70◦ from heights corresponding to the
0 10 20 30 40 50 60 70 80 90 0 200 400 600 800 1000Refraction angle (deg)Horizontal distance (perpendicular to field) [km] 0 200 400 600 800 1000 0 200 400 600 800 1000Distance along field [km]Horizontal Distance (perpendicular to field) [km](b)(a)(a)(b) 0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4-80-60-40-20 0 20 40 60 80FractionTangent plane latitude 0 0.05 0.1 0.15 0.2 0.25 0 10 20 30 40 50 60 70 80 90FractionTangent plane co-longitude (refraction angle)MUTEL ET AL.: ANGULAR BEAMING OF AKR
X - 3
Figure 3. U pper panel : Locations of individual AKR bursts at 125 kHz (red), 250 kHz (green) and 500 kHz (blue) at
epochs (a) 22 January 2003, (b) 23 Jul 2005, and (c) 28 July 2005. Mutual visibility maps for the Cluster array of spacecraft
at each epoch are shown in yellow. Lower panel : Cluster spacecraft tangent plane coordinates as determined from each
AKR burst location.
22 January 200300:52:30 to 02:35:2112:0018:0006:0000:0050.070.023 July 200502:21:18 to 02:59:5912:0018:0006:0000:00-50.0-70.028 July 200512:37:15 to 13:20:2812:0018:0006:0000:00-50.0-70.050-50100-100LatitudeLongitude125 kHz250 kHz500 kHz50-500028 July 200550-50100-100LatitudeLongitude125 kHz250 kHz500 kHz50-500023 July 200550-50100-100LatitudeLongitude125 kHz250 kHz500 kHz50-500022 January 2003(a)(b)(c)X - 4
MUTEL ET AL.: ANGULAR BEAMING OF AKR
local electron gyro-frequency. We assume a dipole magnetic
field, and an average electron density profile
−3(cid:16) r
(cid:17)−2
(1)
ne(r) = 50 cm
2re
where re is the radius of the Earth. Although the electron
density varies significantly with both magnetic latitude and
geomagnetic activity, this profile represents an approximate
long-term average [cf. Laakso et al., 2002, Fig. 3a]. The
resulting refraction angles versus coordinate position, along
with the calculated ray paths, are shown in Fig. 1b. The ray
tracing solution shows that the expected refraction is very
strong, with average refraction angle between 70◦ and 80◦ ,
consistent with observed ray paths. More realistic magneto-
spheric refraction models, including possible reflection from
the plasmasphere, result in similar total refraction angles
(e.g. Xiao et al. [2007]).
3.3. Visibility maps
AKR emission that is beamed in a tangent plane and re-
fracted upward will illuminate two symmetrically displaced
azimuthal slices on a great circle belt that is the extension
of the tangent plane on the sky. The belt will have a char-
acteristic width given by the effective opening angle of the
tangent plane and a longitudinal extent given by the (up-
ward) refracted ray paths determined by the ambient mag-
netospheric plasma outside the cavity. For a given spacecraft
location, only a fraction of all source locations on the polar
cap are able to illuminate the spacecraft. Hence, for a given
tangent plane opening angle and refraction angle range, we
can calculate the locus of all hypothetical AKR source loca-
tions on the polar cap which can illuminate the spacecraft.
We denote this locus of points as the visibility map for that
spacecraft location.
The upper panel of Fig. 3 shows VLBI maps of AKR
source positions for three epochs (22 Jan 2003 at 00:52 -
02:35 UT, 23 July 2005 at 02:21-03:00 UT, and 28 July
Figure 4. Comparison of visibility maps for tangent
plane beaming model (0◦ - 45◦ longitude, 0◦±15◦ lati-
tude, yellow) with 20◦ wide hollow cone models [open-
ing angles 20◦ (red) and 40◦ (green)] for the 4-spacecraft
Cluster array on Nov 10, 2005 at 11:30 UT. The locations
of AKR burst source determined from differential delay
solutions are shown as black x's (125 kHz), and squares
(250 kHz).
2005 at 12:37-13:20 UT) mapped onto a CGM coordinate
grid. Joint visibility maps of the Cluster array (i.e., the
visibility area common to all four spacecraft) are shown in
yellow. We computed the visibility maps assuming a tangent
plane beaming pattern with a -15◦ to 15◦ latitudinal range
and 0◦ to 45◦ longitude range. Note that for each epoch the
AKR burst locations lie almost entirely within the visibility
map. The lower panel shows the tangent plane coordinates
of each spacecraft as viewed from each AKR source location.
This figure provides clear confirmation that the spacecraft
are located close to the tangent plane equator of each source
as expected from the tangent plane beaming model. We can
also see that the longitudinal positions are close to the source
zenith, i.e., the ray paths suffer substantial refraction.
3.4. Comparison with hollow cone models
Could the observations discussed in this paper also be
compatible with hollow cone beaming models? In Fig. 4
we compare visibility maps calculated for a tangent plane
beaming model (yellow) with hollow cone models having 20◦
opening angle (red) and 40◦ opening angle (green). The tan-
gent plane model has a width of 30◦ while the hollow cone
models have widths of 20◦. The visibility maps were com-
puted for Cluster spacecraft positions on Nov 10, 2005 at
11:30 UT. The black x's (125 kHz)and squares (250 kHz)
show AKR source locations determined by VLBI. It is clear
that hollow cone models cannot fit the observed source lo-
cations, while the tangent plane model provides an excellent
fit.
4. Implications of
Beaming Model
the Tangent Plane
The observations reported in this paper demonstrate that
individual AKR bursts do not radiate in either filled or hol-
low cones, as previously suggested, but rather are confined
to a narrow plane tangent to the source's magnetic latitude
circle and containing the local magnetic field vector. The
rays are also directed upward, consistent with expected re-
fraction as rays leave the auroral cavity. This geometry con-
firms the numerical models of Louarn and Le Qu´eau [1996]
and Pritchett et al. [2002] predicting longitudinal propaga-
tion.
It also implies that AKR observations from remote
locations sample only a small part of the auroral oval from
any given location. For example, the maps of AKR emission
published by Mutel et al. [2003, 2004] likely represent only
a small fraction of the total extent of AKR emission on the
auroral oval.
Acknowledgments. The University of Iowa acknowledges
the support of NASA Goddard Space Flight Center under Grant
NNX07AI24G. We are grateful to Roman Schreiber and Jan
Hanasz for several fruitful discussions.
References
Burinskaya, T. M., and J. L. Rauch (2007), Waveguide regime
of cyclotron maser instability in plasma regions of de-
pressed density, Plasma Physics Reports, 33, 28–37, doi:
10.1134/S1063780X07010047.
Calvert, W. (1981a), The stimulation of auroral kilometric radi-
ation by type III solar radio bursts, Geophys. Res. Lett., , 8,
1091–1094.
Calvert, W. (1981b), The AKR emission cone at low frequencies,
Geophys. Res. Lett., , 8, 1159–1162.
Calvert, W. (1987), A Jupiter Data Analysis Program (JDAP)
research grant on wave accessibility and attributes, Tech. rep.
Dulk, G. A. (1970), Characteristics of Jupiter's Decametric Radio
Source Measured with Arc-Second Resolution, Astrophys. J.,
, 159, 671–+.
MUTEL ET AL.: ANGULAR BEAMING OF AKR
X - 5
Ergun, R. E., et al. (1998), FAST satellite wave observations in
the AKR source region, Geophys. Res. Lett., , 25, 2061–2064,
doi:10.1029/98GL00570.
Gaelzer, R., F. L. Ziebell, and R. S. Schneider (1994), Ray tracing
studies on auroral kilometric radiation in finite width auroral
cavities, J. Geophys. Res., , 99, 8905–8916.
Green, J. L. (1988), Ray tracing planetary radio emissions, in
Planetary Radio Emissions II, edited by H. O. Rucker, S. J.
Bauer, and B. M. Pedersen, pp. 355–379.
Green, J. L., and D. L. Gallagher (1985), The detailed intensity
distribution of the AKR emission cone, J. Geophys. Res., , 90,
9641–9649.
Green, J. L., D. A. Gurnett, and S. D. Shawhan (1977), The an-
gular distribution of auroral kilometric radiation., J. Geophys.
Res., , 82, 1825–1838.
Gurnett, D., and R. Anderson (1981), The kilometric radio emis-
sion spectrum: Relatonship to auroral acceleration processes,
Geophysicsal Monogrpah Series, 25, 341–350.
Gurnett, D. A. (1974), The earth as a radio source: terrestrial
kilometric radiation., J. Geophys. Res., , 79, 4227–4238.
Gurnett, D. A., et al. (2001), First results from the Cluster wide-
band plasma wave investigation, Annales Geophysicae, 19,
1259–1272.
Kasaba, Y., H. Matsumoto, K. Hashimoto, and R. R. Anderson
(1997), Angular distribution of auroral kilometric radiation ob-
served by the GEOTAIL spacecraft, Geophys. Res. Lett., , 24,
2483–+.
Laakso, H., R. Pfaff, and P. Janhunen (2002), Polar observations
of electron density distribution in the Earth's magnetosphere.
1. Statistical results, Annales Geophysicae, 20, 1711–1724.
Louarn, P., and D. Le Qu´eau (1996), Generation of the Auro-
ral Kilometric Radiation in plasma cavities-II. The cyclotron
maser instability in small size sources, Plan. Sp. Sci., 44, 211–
224.
Mutel, R., D. Gurnett, and I. Christopher (2004), Spatial and
temporal properties of AKR burst emission derived from Clus-
ter WBD VLBI studies, Annales Geophysicae, 22, 2625–2632.
Mutel, R. L., D. A. Gurnett, I. W. Christopher, J. S. Pickett,
and M. Schlax (2003), Locations of auroral kilometric radi-
ation bursts inferred from multispacecraft wideband Cluster
VLBI observations. 1: Description of technique and initial re-
sults, Journal of Geophysical Research (Space Physics), 108,
8–1, doi:10.1029/2003JA010011.
Mutel, R. L., J. D. Menietti, I. W. Christopher, D. A. Gurnett,
and J. M. Cook (2006), Striated auroral kilometric radiation
emission: A remote tracer of ion solitary structures, Journal
of Geophysical Research (Space Physics), 111, 10,203–+, doi:
10.1029/2006JA011660.
Mutel, R. L., W. M. Peterson, T. R. Jaeger, and J. D. Scudder
(2007), Dependence of cyclotron maser instability growth rates
on electron velocity distributions and perturbation by solitary
waves, Journal of Geophysical Research (Space Physics), 112,
7211–+, doi:10.1029/2007JA012442.
Pottelette, R., R. A. Treumann, and M. Berthomier (2001), Au-
roral plasma turbulence and the cause of auroral kilometric
radiation fine structure, J. Geophys. Res., , 106, 8465–8476,
doi:10.1029/2000JA000098.
Pritchett, P. L., R. J. Strangeway, R. E. Ergun, and C. W. Carlson
(2002), Generation and propagation of cyclotron maser emis-
sions in the finite auroral kilometric radiation source cavity,
Journal of Geophysical Research (Space Physics), 107, 13–1,
doi:10.1029/2002JA009403.
Roux, A., et al. (1993), Auroral kilometric radiation sources - In
situ and remote observations from Viking, J. Geophys. Res., ,
98, 11,657–+.
Schreiber, R., O. Santolik, M. Parrot, F. Lefeuvre, J. Hanasz,
M. Brittnacher, and G. Parks (2002), Auroral kilometric radi-
ation source characteristics using ray tracing techniques, Jour-
nal of Geophysical Research (Space Physics), 107, 20–1, doi:
10.1029/2001JA009061.
Treumann, R. A. (2006), The electron cyclotron maser for as-
trophysical application, Astron. Astrophys. Rev., 13, 229–315,
doi:10.1007/s00159-006-0001-y.
Xiao, F., L. Chen, H. Zheng, and S. Wang (2007), A paramet-
ric ray tracing study of superluminous auroral kilometric ra-
diation wave modes, Journal of Geophysical Research (Space
Physics), 112, 10,214–+, doi:10.1029/2006JA012178.
R. L. Mutel, Dept. Physics and Astronomy, University of Iowa,
Iowa City IA 52242 ([email protected])
I. W. Christopher, Dept. Physics and Astronomy, University
of Iowa, Iowa City IA 52242 ([email protected])
|
astro-ph/0109313 | 1 | 0109 | 2001-09-19T13:59:38 | High-redshift QSOs in the FIRST survey | [
"astro-ph"
] | In a pilot search for high-redshift radio QSOs, we have obtained spectra of 55 FIRST sources (S(1.4GHz) > 1 mJy) with very red (O-E > 3) starlike optical identifications. 10 of the candidates are QSOs with redshifts 3.6 < z < 4.4 (4 previously known), six with z > 4. The remaining 45 candidates comprise: a z = 2.6 BAL QSO; 3 low-redshift galaxies with narrow emission lines; 18 probable radio galaxies; and 23 M stars (mainly misidentifications). The success rate (high-redshift QSOs / spectroscopically-observed candidates) for this search is 1/2 for S(1.4GHz) > 10 mJy, and 1/9 for S(1.4GHz) > 1 mJy. With an effective search area of 4030 deg^2, the surface density of high-redshift (z > 4) QSOs discovered with this technique is 0.0015 deg^-2. | astro-ph | astro-ph |
Mon. Not. R. Astron. Soc. 000, 000 -- 000 (0000)
Printed 25 October 2018
(MN LATEX style file v1.4)
High-redshift QSOs in the FIRST survey
C.R. Benn,1 ⋆ M. Vigotti,2 M. Pedani,3 J. Holt,1,4 K.-H. Mack,6,2,7 R. Curran,1,5
S.F. S´anchez1
1Isaac Newton Group, Apartado 321, 38700 Santa Cruz de La Palma, Spain
2Istituto di Radioastronomia, CNR, via Gobetti 101, 40129 Bologna, Italy
3Centro Galileo Galilei, 38700 Santa Cruz de La Palma, Spain
4Department of Physics & Astronomy, University of Sheffield, Hicks Building, Sheffield S3 7RH, UK
5Department of Physical Sciences, University of Hertfordshire, College Lane, Hatfield, Herts, AL10 9AB, UK
6ASTRON/NFRA, Postbus 2, NL-7990 AA Dwingeloo, the Netherlands
7Radioastronomisches Institut der Universitat Bonn, Auf dem Hugel 71, D-53121 Bonn, Germany
25 October 2018
ABSTRACT
In a pilot search for high-redshift radio QSOs, we have obtained spectra of 55 FIRST
sources (S1.4GHz > 1 mJy) with very red (O-E > 3) starlike optical identifications. 10
of the candidates are QSOs with redshifts 3.6 < z < 4.4 (4 previously known), six with
z > 4. The remaining 45 candidates comprise: a z = 2.6 BAL QSO; 3 low-redshift
galaxies with narrow emission lines; 18 probable radio galaxies; and 23 M stars (mainly
misidentifications). The success rate (high-redshift QSOs / spectroscopically-observed
candidates) for this search is 1/2 for S1.4GHz > 10 mJy, and 1/9 for S1.4GHz > 1 mJy.
With an effective search area of 4030 deg2, the surface density of high-redshift (z >
4) QSOs discovered with this technique is 0.0015 deg−2.
Key words: quasars: general - quasars: emission lines - radio continuum: galaxies -
early Universe
1
INTRODUCTION
Observations of high-redshift (z > 4) QSOs strongly con-
strain models of: the formation and evolution of galaxies
and their central black holes (Kauffmann & Haehnelt 2000);
the contribution of QSOs to the ionisation of the intergalac-
tic medium at high redshifts (Steidel, Pettini, Adelberger
2001); and, through studies of the Lyman absorption forest,
the chemical evolution of the intergalactic medium along the
line of sight to the QSO (Rauch 1998, Hamann & Ferland
1999).
Most of the ∼ 300 z > 4 QSOs now known have been
discovered as a result of searches for objects with unusually
red optical colours (Fig. 1). A summary of recent searches is
given in Table 1. Kennefick et al (1995a, b) searched for ob-
jects with unusually red g, r, i colours in 340 deg2 covered by
the second Palomar Sky Survey (POSS-II) of the northern
hemisphere and discovered 10 z > 4 QSOs (see also Djor-
govski 2001). Similarly, Sloan Digital Sky Survey (SDSS)
commissioning images have been searched for objects with
red u, g, r, i, z colours. More than 100 z > 3.5 QSOs have
been found so far (Fan et al 2000, 2001; Schneider et al 2001;
Zheng et al 2001; Anderson et al 2001, Richards et al 2001),
⋆ Email: [email protected]
c(cid:13) 0000 RAS
and this number will rise rapidly as SDSS nears completion
(∼ 10000 deg2) over the next few years. In the south, Storrie-
Lombardi et al (2001) searched UKST plates scanned with
the Automated Plate Measuring facility (APM) in Cam-
bridge for objects with BJ − R > 2.5, and found 49 QSOs
with z > 4.
A disadvantage of the purely optical searches is that
complex colour criteria are needed to exclude the much
larger number of red stars, which makes it difficult to correct
for incompleteness. Alternatively, simple one-colour criteria
can be used, but then a large fraction of the high-redshift
QSOs is missed (compare in Table 1 the surface densities at-
tained for different types of optical selection). Starting with
a sample of radio QSOs allows one to reduce the number
of candidates without resorting to complex colour criteria,
and is less likely to bias the selection against dusty objects.
E.g. Hook et al (1998) sought red objects identified with
flat-spectrum QSOs with S5GH z > 25 mJy over 1600 deg2,
and found 6 with z > 3 (none with z > 4). Snellen et al
(2001) sought red objects identified with flat-spectrum radio
sources S5GH z > 30 mJy in the Cosmic Lens All-Sky Sur-
vey (CLASS, Myers et al 2001) over 6400 deg2, and found 4
QSOs with z > 4, i.e. 1 z > 4 QSO per 1600 deg2.
Higher surface densities can be attained by observing
the counterparts of fainter radio sources. In this paper we
2
C.R. Benn et al
Figure 1. The variation of POSSI O − E colour with redshift,
for QSOs from the FIRST bright quasar survey (crosses, White et
al 2000), the SDSS and POSSII searches for high-redshift QSOs
(spots), and a selection of radio QSOs with 3.4 < z < 4.0 from
the NASA Extragalactic Database (circles). The QSOs in these
samples were selected in part on the basis of colour, but nev-
ertheless trace the typical reddening of a QSO with increasing z
(prediction for typical QSO shown as solid line) as the continuum
redward of the Lyα line moves out of the observed O band (red
limit ≈ 5000 A). As redshift increases above 4.5, the predicted
brightness of a QSO in E band drops sharply, so few are likely to
be detected. With the criteria O − E > 3, E < 18.6 used here,
our QSO search is probably complete for z > 4, but incomplete
for 3 < z < 4.
report on a pilot search for high-redshift QSOs amongst very
red, O − E > 3 (see Fig. 1), optical identifications of FIRST
radio sources S1.4GH z > 1 mJy (with no selection on radio
spectral index).
There is no overlap between this search and the FIRST
bright quasar survey (White et al 2000, Becker et al 2001),
whose selection criterion O − E < 2 preclude detection of
most high-redshift QSOs (none with z > 3.7 were cata-
logued), as do the colour selection criteria of most large QSO
surveys, e.g. the 2dF QSO survey (Croom et al 2001).
2 SAMPLE
The FIRST radio catalogue (White et al 1997) currently in-
cludes 722354 sources detected with peak S1.4GH Z > 1 mJy
over 7988 deg2 (≈ 90 sources deg−2), mainly 7 < RA < 17h,
-5 < Dec < 57o. We sought optical identifications of these
sources with objects catalogued by the APM (Irwin et al
1994) on the POSS-I O (blue) and E (red) plates. We did
not seek optical identifications at the mid-points of likely
double radio sources (∼ 10% of the catalogue), since this
would have required a substantial enlargement of the search
area. The colour-magnitude distribution of the star-like opti-
cal identifications lying within 1.5 arcsec of a FIRST source
is shown in Fig. 2. We selected as candidate z > 4 QSOs
Figure 2. Distribution in colour and magnitude of starlike
(POSSI/APM) optical identifications of FIRST radio sources.
Objects with no O magnitude have been plotted with O − E =
21.7 - E, and the E magnitudes have been randomised ± 0.1 so
that the density of points in this region can be seen. Most of the
objects in the lower part of the diagram are true QSOs. Most of
those clustered at upper right are misclassified radio galaxies (the
dashed line shows the expected colour-magnitude relationship for
a redshifted giant elliptical galaxy), and many of the other red
objects are misidentifications with galactic M stars. Some may
also be BAL QSOs, which tend to be red. The candidate high-
redshift QSOs observed here are drawn from the region enclosed
by the dotted line.
all starlike optical identifications with O − E > 3 (see Fig.
1) and E < 18.6 (the POSS blue limit is O ≈ 21.7). Trial
spectroscopy showed that ∼ 80% of these are actually radio
galaxies, which greatly outnumber QSOs at E ∼ 18, O−E ∼
3 (Fig. 2). It was possible to filter out most of these by in-
specting images from the digitised POSS-II survey, using the
SExtractor program (Bertin & Arnouts 1996) to characterise
the relationship between FWHM and intensity for stars (in
each image), and thus to distinguish between stellar and ex-
tended images. In addition, we filtered out the ≈ 25% of
starlike images for which APM recorded no blue magnitude
(i.e. implied magnitude fainter than the blue plate limit),
but which were easily visible on the Minnesota APS scans
of POSS-I (Pennington et al 1993); these images were missed
by APM due to the coarser scanning resolution and greater
susceptibility to confusion with nearby images. This selec-
tion procedure yielded a total of 109 candidate high-redshift
QSOs.
3 OBSERVATIONS AND REDUCTION
Spectra were obtained of a random sample of 55 of the 109
candidates with the Isaac Newton Telescope's IDS spectro-
graph on 2001 Mar 14, and with the Isaac Newton (IDS),
William Herschel (ISIS) and Calar Alto 2.2-m (CAFOS)
spectrographs on service nights between 2000 March 19 and
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
Table 1. Recent searches for high-redshift QSOs
Search programme
(1)
DPOSS-II
SDSS
APM/UKST
INT wide-angle surv.
CLASS flat-spectrum
GB/FIRST flat-spec.
"
FIRST
"
Colour
criterion
(2)
optical
limit
(3)
r < 19.6
i < 20
gri
ugriz
BJ − R >2.5 R < 19.3
griz
r < 24
Slim
mJy
(4)
--
--
--
--
O − E > 2
O − E > 1.2
O − E > 1.2
O − E > 3
O − E > 3
E <19
E <19.5
E <19.5
E <18.6
E <18.6
S5 > 30
S5 >25
S5 >25
S1.4 >1
S1.4 >10
High-redshift QSOs in FIRST
3
Area Cands
deg2
(5)
(6)
340
180
8000
12.5
6400
1100
1100
4030
4030
85
27
50
50
55
9
z >
N σ
(7)
4.0
4.0
4.0
4.0
4.0
4.0
3.0
4.0
4.0
(8)
10
18
49
3
4
0
6
6
4
deg−2
(9)
0.029
0.10
0.0061
0.24
0.00063
0.0
0.005
0.0015
0.0010
Hiz /
cands
(10)
0.11
Reference
(11)
Kennefick et al 1995a,b
Fan et al 2001
Storrie-L. et al 2001
Sharp et al 2001
0.15
0.00
0.12
0.11
0.44
Snellen et al 2001
Hook et al 1998
"
This paper
"
Column 4 gives, where appropriate, the radio flux-density limit Sν , for frequency ν in GHz. Column 9 gives the surface density
of high-redshift QSOs discovered. Column 10 gives the ratio between the number of high-redshift QSOs found (column 8) and the
number of candidates for spectroscopy (column 6).
The images were debiased and flat-fielded, cosmic
rays were eliminated, and the spectra were extracted,
wavelength-calibrated and intensity-calibrated in the usual
way, using the IRAF package. A summary of the observa-
tions and results is given in Table 2.
4 RESULTS
Spectra showing clear emission or absorption features were
classified 'Q' (QSO, broad emission lines), 'S' (Sy2 or star-
burst type spectrum, narrow emission lines) or '∗' (spectral
features of M star). The remainder have been classified 'G'
(probable radio galaxy, i.e. giant elliptical), where the char-
acteristic 4000-A break is detected, otherwise '?'. Most of
the '?' objects will be radio galaxies, rather than misiden-
tifications with stars, judging from the number of M stars
detected, the typical distribution of spectral types at this
magnitude, and the magnitude distributions and the distri-
butions on the sky of the '?' and '∗' objects. One of the '?'
objects, 1036+42, has a flat SED inconsistent with it be-
ing a radio galaxy; it may be a low-redshift QSO (z < 2.2).
Of the other '?' objects, none has an SED consistent with
it being a high-redshift QSO, except perhaps 1120+34 and
1349+38. The redshifts of the high-redshift QSOs were mea-
sured from emission lines other than Lyα (because of the
asymmetrical absorption), except in the case of 0839+51.
The colour-magnitude diagram of the candidates is shown
in Fig. 3. The spectra of the high-redshift QSOs are pre-
sented in Figs. 4, 5.
We find 10 high-redshift QSOs (z > 3.6) out of 55 candi-
dates. These candidates were selected from 109 for the whole
FIRST catalogue, covering 7988 deg2, so the effective area
covered by this search is ∼ 4030 deg2, i.e. we find 0.0025
high-redshift QSOs deg−2 with z > 3.6, or 0.0015 deg−2
with z > 4.0. For S1.4GH z > 1 mJy, E < 18.6, the z > 4
search will be near-complete (Fig. 1). The surface density
on the sky is compared with that from other searches for
high-redshift QSOs in Table 1. The surface density of z > 4
radio QSOs discovered here is twice as high as that from the
Snellen et al search, because of the lower flux-density limit.
The efficiency of the search (number of high-redshift QSOs /
number of candidates for spectroscopy) is 0.44 for S1.4GH z >
10 mJy, higher than for previous searches, probably because
of the careful filtering of the candidates.
Figure 3. Colour-magnitude distribution for the candidate high-
redshift QSOs. The symbols for spectral type are as used in Ta-
ble 2 (Q = QSO, S = narrow-emission-line galaxy, G = radio
galaxy, ∗ = star, ? = probable radio galaxy). Large font indicates
S1.4GH z > 10 mJy; small font S1.4GH z < 10 mJy. A nominal
POSSI blue plate limit O = 21.7 is indicated by the dashed line.
Objects with no O magnitude have been plotted with O − E =
21.7 - E, and the E magnitudes have been randomised ± 0.2 mag
to reduce crowding. The arrows indicate upper or lower limits.
2001 Jan 5 (dates given in Table 2). Apart from 2000 Apr 5,
May 12 and 2001 Jan 5, the nights were photometric. The
INT IDS spectrograph was used with the R150V grating,
yielding spectra with dispersion 6.5 A/pixel, usually cen-
tred at 6500 A. The WHT ISIS dual-arm spectrograph was
used with the R158R and R158B gratings in the red and
blue arms respectively (with a dichroic separating the light
blue and red of 6100 A). In the red arm, with a TEK CCD
detector, this yielded spectra with dispersion 2.9 A, usually
centred at 7400 A. In the blue arm, with an EEV CCD, the
dispersion was 1.6 A/pixel, and the spectra were usually
centred at 4800 A. The CAFOS spectrum has dispersion 8.5
A/pixel.
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
4
C.R. Benn et al
Figure 4. Spectra of 8 of the 10 high-redshift QSOs (see Fig 5 for the remainder). Spectral features are labelled at wavelengths
corresponding to the quoted redshift, assuming rest-frame wavelengths in A of 1216 (Lyα), 1240 (NV), 1302 (OI/SiII blend), 1400
(SiIV/OIV] blend), 1549 (CIV). For 0941+51, the metal-line absorption systems at z = 3.80, 3.83 are also indicated. The spectra have
not been corrected for the terrestrial atmospheric absorption bands, notably at 7594 and 6867 A (A and B bands).
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
Table 2. High-redshift QSO candidates
RA
J2000
(1)
Dec
J2000
(2)
High-redshift QSOs:
S1.4
mJy
(3)
R-O
′′
(4)
E O − E
Tel
(5)
(6)
(7)
Date
obs.
(8)
0.6
072518.29
0.8
074711.16
0.7
083141.68
0.8
083946.11
1.3
094119.36
0.5
105756.27
0.4
122027.93
1.1
130940.61
0.9
142308.19
0.9
151601.55
0.5
163950.51
Low-redshift QSOs and galaxies:
370517.9
273903.6
524517.1
511203.0
511932.3
455553.0
261903.6
573309.1
224157.5
430931.3
434003.3
26.6
1.5
0.9
41.6
2.7
1.4
35.0
11.5
35.4
1.5
23.8
091441.31
093255.83
101058.24
103549.91
103647.80
104622.91
104801.61
104958.45
105719.80
105750.00
111411.21
111557.91
111707.05
112028.10
115230.18
121532.16
124450.93
130604.97
134951.93
140816.68
143749.26
M stars:
072359.95
072801.49
074144.39
075127.64
084407.26
085700.54
090556.33
091953.76
100125.11
105102.41
111212.69
112518.16
113043.07
120849.72
132903.23
133833.44
135357.42
150859.95
150938.97
161150.75
163535.34
164217.03
164401.60
295621.4
353439.1
283247.0
300732.1
420852.7
345436.0
260032.3
411043.4
342807.0
353300.4
293244.4
352757.1
411736.1
345829.4
271808.4
250956.6
353906.7
352603.9
382334.9
352205.8
325917.8
523949.4
380344.2
333549.3
272736.4
280740.4
275540.0
224419.1
340906.7
370629.4
314914.5
431015.5
495158.8
514134.1
392907.1
323031.5
531642.1
343659.7
271431.0
434649.8
434412.4
352415.9
402230.8
420047.4
4.8
1.9
2.3
9.0
2.2
1.0
11.2
1.3
5.4
2.3
1.1
3.4
0.7
6.5
4.2
1.0
1.5
1.1
1.7
9.6
1.3
0.8
1.6
136.8
1.2
6.2
0.8
2.4
1.2
1.1
9.0
1.4
3.0
4.2
0.8
5.8
0.9
98.0
1.0
1.9
1.0
5.2
6.4
4.2
0.3
0.9
0.4
0.3
0.4
0.7
0.5
0.5
0.0
0.7
1.0
1.4
0.8
0.8
0.4
0.7
0.2
0.9
0.4
1.1
0.7
0.8
0.7
0.6
1.2
0.5
0.8
1.5
0.8
0.8
0.4
0.9
1.1
0.5
1.2
1.4
0.7
1.2
0.4
0.8
1.0
0.6
0.2
1.2
18.3
17.2
15.3
18.5
17.6
16.5
17.9
18.1
18.5
18.4
17.1
18.6
18.2
18.6
18.4
18.6
18.6
18.6
18.0
18.1
18.5
18.0
18.5
18.5
18.2
17.0
18.0
17.1
18.4
18.4
18.4
18.6
17.6
17.3
18.4
18.0
17.5
18.5
17.5
17.1
16.8
17.8
17.2
17.8
17.7
15.3
18.1
17.9
17.8
17.9
18.4
17.8
18.1
17.4
17.6
I
I
I
I
I
>3.4 W 010105
010314
>4.5
010314
3.9
>3.2
010314
>4.1 W 000319
000405
010314
>3.6 W 000319
010314
>3.2
010314
>3.3
3.9 W 000718
3.9
3.3
I
I
>3.1
3.1
>3.1
3.3
>3.1
>3.1
3.0
3.2
>3.6
>3.2
3.8
>3.2
>3.2
>3.5
3.0
3.7
3.5
3.0
>3.3
>3.4
>3.1
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
010314
010314
010314
010314
010314
010314
010314
010314
010314
010314
010314
010314
010314
010314
010314
010314
000405
010314
010314
010314
010314
I
I
>4.1
3.7
010314
010314
>3.3 W 010104
>3.7 W 010104
3.0 W 010105
>3.2 W 010105
>4.2
010314
I
010314
3.0
I
010314
3.2
I
I
>3.9
010314
C 000512
3.7
3.4
I
010314
010314
I
3.1
010314
I
3.2
010314
I
3.4
010314
I
3.3
3.9
I
010314
010314
I
3.4
010314
I
>3.4
3.4
I
010314
3.4 W 000718
3.1 W 000718
3.5 W 000718
High-redshift QSOs in FIRST
5
ID z
σz
Notes
(10)
(11)
(12)
(13)
Irwin et al (1998), z = 3.91
Snellen et al (2000), z = 4.41
Stern et al (2000), z = 4.12
Q 4.33
Q 4.17
Q 3.87
Q 4.42
Q 3.85
Q 4.11
Q 3.694
Q 4.274
Q 4.316
Q 2.590
Q 3.99
.01
.02
.03
.02
.01
.02
.005
.005
.005
.005
.015
0.587
.002
3727, 5007?
blue SED
0.279
.001
3727 4861 4959 5007 6563
possible high-redshift QSO
0.282
.001
3727 4861 4959 5007
possible high-redshift QSO
G
?
?
S
?
G
?
?
G
?
S
?
G
?
?
S
?
?
?
G
G
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
T
sec
(9)
900
200
200
300
1200
300
1200
300
1200
300
1000
300
300
100
300
200
300
300
300
300
1300
300
300
300
300
300
300
300
300
300
300
300
300
300
400
400
1200
400
200
200
200
300
900
300
300
300
300
300
300
300
300
300
300
300
300
The columns give: (1-2) right ascension and declination, (3) FIRST 1.4-GHz (integrated, sometimes < peak) flux density, (4) radio
- optical displacement, (5-6) APM POSSI E magnitude and O − E colour, (7) telescope with which spectrum obtained (W =
William Herschel 4.2-m, I = Isaac Newton 2.5-m, C = Calar Alto 2.2-m), (8) date of observation, (9) exposure time, (10) spectral
type (Section 4), (11-12) redshift and rms error on redshift. For low-redshift emission-line objects, the rest-frame wavelengths of
detected lines are given in A in the last column: 3727 OII, 4861 Hβ, 4959/5007 OIII, 6563 Hα. For high-redshift QSOs, see Figs.
4, 5.
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
6
C.R. Benn et al
Figure 5. Spectra of the two remaining high-redshift QSOs (see caption of Fig. 4); and of 4 other emission-line objects detected in this
search (see text). 1516+43 was undetected on the POSS blue plate, but the field is optically confused and its true colour may be bluer
than O − E = 3.
Four of
the 10 high-redshift QSOs were previ-
ously known: 0747+27 (M. Irwin, private communication),
0831+52 (Irwin et al 1998), 0839+51 (Snellen et al 2001) and
1057+45 (Stern et al 2000). 0831+52 is unusually bright (E
= 15) and the Lyα forest been studied in detail (Ellison et
al 1999). 327-MHz flux densities are available for three of
the high-redshift QSOs from the WENSS survey (Rengelink
et al 1997). The 0.3 - 1.4-GHz spectral indices α (Sν ∝ ν α)
are all flat: -0.1 for 0725+37, 0.0 for 0839+51, and -0.4 for
1639+43. The other 7 QSOs are too faint to be detected in
WENSS, or lie below the WENSS declination limit.
The SEDs of the high-redshift radio QSOs reported here
are as varied as those found in optical searches. At least
0941+51 and 0831+52 show metal-line absorption redward
of Lyα, probably associated with saturated Lyα lines in
the forest. One, 1309+57, shows very narrow Lyα emission,
FWHM 20 A. 1639+43 has almost no Lyα emission, which
is unusual (only ∼ 1/60 of SDSS QSOs), and it is similar in
this respect to the z = 4.2 radio QSO, 0918+06, discovered
by Snellen et al. Snellen et al suggested that the unusually
strong absorption of the Lyα line might be due to the host
galaxy (in both cases this is probably a giant elliptical). A
few of the QSOs have probable damped Lyα systems (DLAs,
seen in ∼ 20% of high-redshift QSOs). 0941+51 has a DLA
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
at z = 3.52, and this has now been observed at high spectral
resolution with WHT ISIS (Centurion et al, in preparation).
The other 45 candidates are a mixed bag (see Maglioc-
chetti et al 2000 for the typical distribution of identification
types of all colours at 1 mJy). 1516+43 is a z = 2.59 broad-
absorption-line (BAL) QSO with broad, deep SiIV and CIV
absorption troughs blueward of the emission features (Fig.
5). 1035+30 (probably), 1114+29 and 1215+25 are low-
redshift galaxies exhibiting narrow emission lines (Fig. 5).
18 have SEDs consistent with those of giant elliptical galax-
ies at z ∼ 0.3. 23 have spectra of late-type M stars. This
number is consistent with that expected for chance coinci-
dences, but a few may be true radio stars (Helfand et al 1999
describe a search for counterparts of bright stars in FIRST).
5 CONCLUSIONS
We have conducted a pilot search for high-redshift radio
QSOs, obtaining optical spectra of 55 FIRST radio sources
(S1.4GH z > 1 mJy) with red (O-E > 3) starlike optical iden-
tifications. 10 of the candidates are QSOs with redshifts 3.6
< z < 4.4 (4 previously known). Six have z > 4. The re-
maining 45 candidates are a mixture of low-redshift galaxies
(misclassified as stellar) and M stars (mainly misidentifica-
tions). The success rate (high-redshift QSOs / candidates)
for this search is 1/2 for S1.4GH z > 10 mJy, and 1/9 for
S1.4GH z > 1 mJy. With an effective search area of 4030
deg2, the surface density of z > 4 QSOs discovered with
this technique is 0.0015 deg−2.
High-redshift QSOs in FIRST
7
Fan X. et al, 2001, AJ, 121, in press, astroph/0008122
Hamann F., Ferland G., 1999, ARA&A, 37, 487
Helfand D.J., Schnee S., Becker R.H., White R.L., McMahon
R.G., 1999, AJ, 117, 1568
Hook I.M., Becker R.H., McMahon R.G., White R.L., 1998, MN-
RAS, 297, 1115
Irwin M.J., Ibata R.A., Lewis G.F., Totten E.J., 1998, ApJ, 505,
529
Irwin M.J., McMahon R.G., Maddox S., 1994, Spectrum no. 2.,
p. 14
Kauffmann G., Haehnelt M., 2000, MNRAS, 311, 576
Kennefick J.D., de Carvalho R.R., Djorgovski S.G., Wilber M.M.,
Dickson E.S., Weir N., 1995a, AJ, 110, 78
Kennefick J.D., Djorgovski S.G., de Carvalho R.R., 1995b, AJ,
110, 2553
Maglioccchetti M., Maddox S.J., Wall J.V., Benn C.R., Cotter
G., 2000, MNRAS, 318, 1047
Myers S.T. et al, 2001, AJ, in press
Pennington R.L., Humphreys, R.M., Odewahn, S.C., Zumach,
W., Thurmes, P.M., 1993, PASP, 105, 521.
Rauch M., 1998, ARA&A, 36, 267
Rengelink R.B., Yang T., de Bruyn A.G., Miley G.K., Bremer
M.N., Rottgering H.J.A., Bremer M.A.R., 1997, A&AS, 124,
259
Richards G.T. et al, 2001, AJ, 121, 2308
Schneider D.P. et al, 2001, AJ, 121, in press, astroph/0012083
Sharp R.G., McMahon R.G., Irwin M.J., Hodgkin S.T., 2001,
MNRAS, astro-ph/0103079
Snellen I.A.G, McMahon R.G., Dennett-Thorpe J., Jackson N.,
Mack K.-H., Xanthopoulos E., 2001, MNRAS, in press, astro-
ph/0103291
Steidel C.C., Pettini M., Adelberger K.L., 2001, ApJ, astro-
ph/0008283
Stern D., Djorgovski S.G., Perley R.A., de Carvalho R.R., Wall
Acknowledgments
J.V., 2000, AJ, 119, 1526
Storrie-Lombardi L.J., Irwin M.J., McMahon R.G., Hook I.M.,
2001, MNRAS, astroph/0012446
White R.L., Becker R.H., Helfand D.J., Gregg M.D., 1997, ApJ,
475, 479
White R.L., Becker R.H., Gregg M.D., Laurent-Muehleisen S.A.,
2000, ApJS, 126, 133
Zheng W. et al, 2000, AJ, 120, 1607
We are grateful to Alessandro Caccianiga (Observatorio
de Lisboa) for taking spectra with the Calar Alto 2.2-
m telescope. Rachel Curran and Joanna Holt were 1-year
placement students at ING in 1999-2000 and 2000-2001
respectively. KHM was supported by the European Com-
mission, TMR Programme, Research Network Contract
ERBFMRXCT96-0034 "CERES". The William Herschel
and Isaac Newton Telescopes are operated on the island of
La Palma by the Royal Greenwich Observatory in the Span-
ish Observatorio del Roque de los Muchachos of the Instituto
de Astrofisica de Canarias. The Calar Alto observatory is op-
erated by the Max-Planck-Institute for Astronomy, Heidel-
berg, jointly with the Spanish National Commission for As-
tronomy. The APS Catalog of POSS I (http://aps.umn.edu)
is supported by NASA and the University of Minnesota. The
NASA/IPAC Extragalactic Database (NED) is operated by
the Jet Propulsion Laboratory, California Institute of Tech-
nology, under contract with NASA.
REFERENCES
Anderson S.F. et al, 2001, MNRAS, astro-ph/0103228
Becker R.H. et al, 2001, ApJS, astro-ph/0104279
Bertin E., Arnouts S., 1996, A&AS, 117, 393
Croom S.M., Smith R.J., Boyle B.J., Shanks T., Loaring N.S.,
Miller L., Lewis I.J., 2001, MNRAS, astro-ph/0104095
Djorgovski S.G., 2001, http://astro.caltech.ed/∼george/z4.qsos
Ellison S.L., Lewis G.F., Pettini M., Sargent W.L.W., Chaffee
F.H., Foltz C.B., Rauch M., Irwin M.J., 1999, PASP, 111, 946
Fan X. et al, 1999, AJ, 118, 1
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
|
astro-ph/0506351 | 1 | 0506 | 2005-06-15T17:14:30 | The masses, radii and luminosities of the components of U Geminorum | [
"astro-ph"
] | We present a phase-resolved spectroscopic study of the secondary star in the cataclysmic variable U Gem. We use our data to measure the radial velocity semi-amplitude, systemic velocity and rotational velocity of the secondary star. Combining this with literature data allows us to determine masses and radii for both the secondary star and white dwarf which are independent of any assumptions about their structure. We use these to compare their properties with those of field stars and find that both components follow field mass-radius relationships. The secondary star has the mass, radius, luminosity and photometric temperature of an M2 star, but a spectroscopic temperature of M4. The latter may well be due to a high metallicity. There is a troubling inconsistency between the radius of the white dwarf inferred from its gravitational redshift and inclination and that inferred from its temperature, flux, and astrometric distance.
We find that there are two fundamental limits to the accuracy of the parameters we can derive. First the radial velocity curve of the secondary star deviates from a sinusoid, in part because of its asphericity (which can be modelled) and in part because the line flux is not evenly distributed over its surface. Second we cannot be certain which spectral type is the best match for the lines of the secondary star, and the derived rotational velocity is a function of the spectral type of the template star used. | astro-ph | astro-ph |
Mon. Not. R. Astron. Soc. 000, 1 -- 13 (2005)
Printed 4 July 2019
(MN LATEX style file v2.2)
The masses, radii and luminosities of the components of U
Geminorum
T. Naylor1 A. Allan1 and K.S. Long2
1School of Physics, University of Exeter, Stocker Road, Exeter EX4 4QL
2Space Telescope Science Institute, 3700 San Martin Drive, Baltimore, MD 21218, USA
ABSTRACT
We present a phase-resolved spectroscopic study of the secondary star in the cat-
aclysmic variable U Gem. We use our data to measure the radial velocity semi-
amplitude, systemic velocity and rotational velocity of the secondary star. Combining
this with literature data allows us to determine masses and radii for both the sec-
ondary star and white dwarf which are independent of any assumptions about their
structure. We use these to compare their properties with those of field stars and find
that both components follow field mass-radius relationships. The secondary star has
the mass, radius, luminosity and photometric temperature of an M2 star, but a spec-
troscopic temperature of M4. The latter may well be due to a high metallicity. There
is a troubling inconsistency between the radius of the white dwarf inferred from its
gravitational redshift and inclination and that inferred from its temperature, flux, and
astrometric distance.
We find that there are two fundamental limits to the accuracy of the parameters
we can derive. First the radial velocity curve of the secondary star deviates from a
sinusoid, in part because of its asphericity (which can be modelled) and in part because
the line flux is not evenly distributed over its surface. Second we cannot be certain
which spectral type is the best match for the lines of the secondary star, and the
derived rotational velocity is a function of the spectral type of the template star used.
Key words: binaries: eclipsing -- white dwarfs -- stars: late-type -- stars: fundamental
parameters -- stars: individual: U Gem -- novae, cataclysmic variables.
1
INTRODUCTION
Precise radii and masses for the white dwarfs and late-type
secondary stars in cataclysmic variables (CVs) are critical
tests of our understanding of their evolution. For the sec-
ondary stars there has been a long debate as to whether they
follow main-sequence mass-radius and mass-temperature re-
lationships. These issues are reviewed in Smith & Dhillon
(1998), who concluded that, although the CV data are con-
sistent with the main sequence relationships, there is such
a large scatter in the CV properties that the main-sequence
relationships cannot be used to predict one quantity from
another. Amongst the CVs with M-type secondaries, a large
part of that scatter was driven by just two systems, IP Peg
and U Gem. However, these are systems whose derived pa-
rameters should be amongst the most precise since they are
eclipsing systems.
In an earlier paper we re-visited IP Peg (Beekman et al.
2000) and brought its parameters into line with those of
the other CVs of similar orbital period, making U Gem a
natural next target. However, there was another, pressing
reason for re-visiting the parameters of this system, in that
Long & Gilliland (1999) obtained a radial velocity curve for
the white dwarf. This is the first radial velocity curve of a
white dwarf obtained for any CV. This not only allows an al-
most assumption-free mass ratio q(= M2/M1) to be derived,
but also allows one to measure the gravitational redshift to
the white dwarf, which given the white dwarf mass yields its
radius. In principle, therefore, one could obtain independent
measurements of both the mass and radius for each of the
white dwarf and secondary star. There were three obstacles
to such a complete analysis.
• There was no measurement of the rotational velocity of
the secondary star (v2sin(i)) from which its radius could be
derived assuming only that it is phase locked.
• To measure the gravitational redshift of the white dwarf
requires a measurement of the mean radial velocity of the
red dwarf (γ2). There were two measurements of this which
conflicted.
• The radial velocity curve of the secondary star showed
an apparent eccentricity. The best explanation for this would
be that the line flux was not distributed evenly across the
surface of the star, but without more data to examine the
2
T. Naylor et al
residuals in detail, any corrections that should be applied to
account for this were uncertain.
We therefore re-observed U Gem to solve these problems.
In doing so we acquired a very high quality dataset, which
illuminates various problems in parameter determination.
This was part of a programme to obtain phase resolved red
spectroscopy of cataclysmic variables and low-mass X-ray
binaries with precise correction for the telluric absorption,
so that one could study the TiO bands (Webb et al. 2002;
Beekman et al. 2000; Webb et al. 2000). The observational
techniques used in this paper should be seen as a further
improvement on the ones described in those papers.
This paper is set out as follows. In Sections 2 and 3
we describe the data acquisition and reduction. The almost
complete absence of interlocking assumptions then allows us
to lay out the determination of the dynamical parameters
in a very formal manner, discussing each parameter and its
uncertainties in turn (Section 4). Our dataset is such that
we discover that the uncertainties in both K2 and v2sin(i)
are dominated by systematics. We then use these to obtain
the physical parameters of interest in Section 5, before com-
paring them to field stars in our discussion (Section 6). We
finally draw our conclusions in Section 7.
2 OBSERVATIONS
Spectra were obtained on the nights beginning January 4,
5, 6 and 7 2001 and January 17 2000 using the 2.5m Isaac
Newton Telescope with the 235mm camera of the Interme-
diate Dispersion Spectrograph using a 1.5 arcsec slit. The
slit width corresponds to approximately two pixels at the
detector, and was a compromise. On the one hand narrow-
ing it further would risk undersampling in wavelength space,
which would compromise the precision with which we could
measure velocities. Conversely, widening it would mean that
velocity shifts would be introduced if the star moved across
the width of the slit. The largest possible value of these shifts
corresponds to the slit width, or ±15km s−1. Since the γ ve-
locity of the white dwarf is only known to this precision, and
in practice the seeing would ensure the shifts were smaller
than this, the slit width was a reasonable compromise. In
fact, as we shall discuss later, we found that these shifts
could be measured and corrected, considerably improving
the quality of our data.
The slit was orientated such that both U Gem and the
V ≃ 15 star at α =07 55 07.08 δ =+21 59 19.7 (J2000) fell in
the slit. For the majority of our spectra we used the R1200R
grating which gave a resolution of 1.55A and a coverage of
7480-8320A at 0.83A/pixel. This allowed us to cover the TiO
band and KI lines at 7700A and the NaI doublet at 8190A.
On the last night we used the R1800V grating to cover 7855-
8335A at 0.48A/pixel with a resolution of 0.76A, covering
just the NaI doublet. We used an exposure of 420s for all our
U Gem spectra. Since we required accurate radial velocities
we took an arc spectrum after every two spectra of U Gem.
We were also concerned to ensure that the signal-to-noise in
our final mean spectra (which is around 350 for the lower
resolution data) was not dominated by fringing on the CCD,
so at the same time as we acquired an arc, we also obtained
a tungsten lamp flatfield.
Since the primary aim of our observing procedure was
to ensure excellent removal of the telluric absorption, we
interspersed our observations of U Gem with observations
of the rapidly rotating A0 star SAO78799 This star is rel-
atively featureless in the range of interest, and so could be
used to correct the telluric bands (see Section 3). After each
observation of the A0 star we took an arc spectrum and a
flatfield before moving the telescope.
We also required spectra of late-type stars for cross cor-
relation and to determine an accurate spectral type for U
Gem. These were obtained using the same instrumental se-
tups as above. Again each observation was accompanied by
a flatfield and arc, along with a spectrum of a nearby A0
star, which also had its own flatfield and arc.
3 SPECTRUM EXTRACTION
We treated all our observations of U Gem and the late-type
stars in an identical fashion. Mean bias frames were sub-
tracted from each image. We then created flatfields by calcu-
lating a correction with respect to a two dimensional polyno-
mial fitted through each flatfield frame, and these were used
to correct the arcs and star images. Spectra were then ex-
tracted from each stellar observation using an optimal tech-
nique (e.g. Horne 1986). For each such spectrum an arc was
extracted from the appropriate exposure, from the same re-
gion of the detector as the star was extracted from. It was
used to derive a polynomial relationship between pixel num-
ber and wavelength, which was used to wavelength calibrate
the stellar spectra. Each target spectrum was then divided
by the appropriate observation of an A0 star to remove the
atmospheric absorption bands.
In an initial reduction of the data, we followed the
telluric correction procedure developed in Beekman et al.
(2000), where no interpolation of the telluric correction star
onto the wavelength scale of the target was attempted as this
can tend to smooth the spectra. Instead each target spec-
trum pixel is simply divided by the nearest pixel in wave-
length from the A0 star spectrum.
In some cases this produced a rather poor telluric cor-
rection, and this was traced to different positions of the tar-
get star across the width of the slit, leading to small differ-
ences in the wavelength calibration. We therefore cross cor-
related all our target spectra and telluric correction spectra
with one A-star spectrum, in the region 7624-7690A which
is dominated by telluric bandheads. We then applied the de-
rived shifts to the wavelength values of each data point in
all our target and telluric correction spectra. Before dividing
we interpolated the telluric correction stars onto the same
wavelength points as their respective targets using quadratic
interpolation. We also normalised the correction spectrum
such that it was one in the middle of its wavelength range
to preserve the overall number of counts correctly for the
target. This entire procedure improved some of the correc-
tions dramatically. Equally importantly, though, it corrects
for the drift of U Gem across the width of the slit, a point
we shall return to in Section 4.1. We could not carry out
this procedure for the higher resolution data, as there is not
a region of such strong telluric absorption.
Each spectrum was corrected to the velocity of the so-
lar system barycentre by allowing for the Earth's motion.
Throughout the reduction process we propagated the un-
certainties from the optimal extraction. This allowed us to
identify four spectra as of significantly lower signal-to-noise
than the others (probably due to cloud), which were removed
from the subsequent analysis. A further spectrum was re-
moved due to a charged particle event in the region used for
the velocity correction procedure. This left us with 118 low
resolution and 59 high resolution spectra. Figure 1 shows
our final summed spectra, shifted into the reference frame
of the secondary star.
U Gem 3
The
118
lower
mean
resolution
late-type
spectra,
corresponding
tra
http://www.astro.ex.ac.uk/people/timn/UGem/description.html.
available
star
are
via
the
CDS,
or
and
spec-
from
4 EXTRACTING THE DYNAMICAL
PARAMETERS
There are three parameters that we were interested in ex-
tracting from our spectra. The observed radial velocity semi-
amplitude of the secondary star, before correction for incli-
nation (K2); the rotational velocity of the secondary star,
again before allowing for inclination (v2sin(i)); and the mean
velocity of the secondary star (γ2). In principle we have
to measure all three parameters simultaneously. For exam-
ple, we need a broadened spectral type template to cross-
correlate against our individual U Gem spectra to obtain
K2, but can only obtain that by first velocity shifting all the
individual spectra into the rest frame of the secondary star
and summing them. Of course we cannot do this without
first knowing K2. We therefore adopted an iterative proce-
dure outlined in the next two paragraphs, with each stage
described in detail in the subsections below.
We began by analysing the lower resolution spectra. An
inital radial velocity curve was extracted by cross correlat-
ing the NaI region of the U Gem spectra against a spectral
type template. A fit to this gave an initial radial velocity
semi-amplitude which we used to remove the orbital motion
of the secondary star by shifting each spectrum in velocity.
However, it was clear that the radial velocity curve was sig-
nificantly non-sinusoidal between phases 0.25 and 0.75 (see
Section 4.1.2 and Figures 3 and 4), and so we restricted
these sums to the data outside this phase range. We then
measured the velocity shift between the summed spectrum
and the spectral type template, and then shifted the stan-
dard, and broadened it so its apparent rotational velocity
matched that of the secondary star. We next re-cross cor-
related each U Gem spectrum with the broadened template
to create a new radial velocity curve, from which we could
again extract radial velocity curve parameters. We iterated
around this procedure until the change in the velocity shift
we applied to the spectral type template converged to within
one km s−1. Having obtained an initial radial velocity solu-
tion from the NaI region using one spectral type template
we then used this as the starting parameters for using each
spectral type template for both the NaI and KI/TiO regions.
As there are a smaller number of high resolution spec-
tra, and they could not be corrected in velocity for the slit
drift (see Section 3) we have used them only to measure
v2sin(i). Thus we coadded them using the K2 found from
the lower resolution data, and then fitted broadened tem-
plates, again using spectra from only phases 0.75 to 0.25.
Figure 1. The mean of all spectra, summed into the rest frame
of the secondary star using the parameters derived from the NaI
lines using GJ213 as the spectral type template. The upper spec-
trum is the sum of the lower resolution data, the lower spectrum
the higher resolution data. The higher resolution data have been
multiplied by 1.8 (the difference in pixel size in wavelength space).
Figure 2. The mean of the low resolution spectra between phases
0.75 and 0.25 (histogram) with the best fit velocity broadened
spectrum of GJ213 (line) overlaid. The wavelength regions are
those over which the fitting was performed.
Aside from the dynamical parameters, the final output
from this procedure is a spectrum in the rest frame of the
secondary star. We show these spectra for both the high
and low resolution data in Figure 1. In addition, Figure 2
shows the comparison between a broadened template star,
and the sum of the spectra between phase 0.75 and 0.25
for the regions we used for the cross-correlation and v2sin(i)
fitting.
4
T. Naylor et al
Table 1. Measured radial velocities from the NaI feature in the
lower resolution spectra. The columns give the barycentric Julian
Date for the middle of the exposure, the calculated phase for the
exposure, the image number, radial velocity and uncertainty. The
uncertainties are scaled to give a χ2
ν of one for a circular fit. All
exposures are of 420s duration.
BJD
Phase
Image
Number
RV
(km s−1)
Uncertainty
(km s−1)
2 451 561.43994
2 451 561.45081
2 451 561.45612
2 451 914.41589
2 451 914.42090
2 451 914.42700
2 451 914.43201
2 451 914.44177
2 451 914.44672
2 451 914.45264
2 451 914.45758
2 451 914.46594
2 451 914.47095
2 451 914.47687
2 451 914.48181
2 451 914.49042
2 451 914.49542
2 451 914.50122
2 451 914.50616
2 451 914.51453
2 451 914.51947
2 451 914.52527
2 451 914.53027
2 451 914.53937
2 451 914.54437
2 451 914.55017
2 451 914.55518
2 451 914.56470
2 451 914.56964
2 451 914.57550
2 451 914.58044
2 451 914.58881
2 451 914.59375
2 451 914.59961
2 451 914.60455
2 451 914.61292
2 451 914.61786
2 451 914.62372
2 451 914.62866
2 451 914.63660
2 451 914.64160
2 451 914.64758
2 451 914.65253
2 451 914.65753
2 451 914.66583
2 451 914.67084
2 451 914.67657
2 451 914.68158
2 451 914.68988
2 451 914.69482
2 451 914.70062
2 451 914.70557
2 451 914.77521
2 451 914.78015
2 451 914.79144
2 451 914.79639
2 451 915.44385
2 451 915.44879
2 451 915.45428
0.5544
0.6158
0.6458
0.8262
0.8545
0.8890
0.9173
0.9725
0.0005
0.0339
0.0619
0.1091
0.1374
0.1709
0.1988
0.2475
0.2758
0.3085
0.3365
0.3838
0.4117
0.4445
0.4728
0.5242
0.5525
0.5853
0.6135
0.6674
0.6953
0.7284
0.7564
0.8036
0.8316
0.8647
0.8927
0.9399
0.9679
0.0010
0.0289
0.0738
0.1021
0.1359
0.1638
0.1921
0.2391
0.2673
0.2998
0.3281
0.3750
0.4029
0.4357
0.4637
0.8573
0.8853
0.9491
0.9770
0.6369
0.6649
0.6959
206090
206094
206095
243111
243112
243115
243116
243123
243124
243127
243128
243135
243136
243139
243140
243146
243147
243150
243151
243157
243158
243161
243162
243168
243169
243172
243173
243180
243181
243184
243185
243191
243192
243195
243196
243202
243203
243206
243207
243213
243214
243217
243218
243219
243225
243226
243229
243230
243236
243237
243240
243241
243272
243273
243280
243281
243399
243400
243403
-122
-219
-247
-253
-224
-185
-141
-49
4
63
121
193
239
262
279
299
293
280
268
210
160
106
53
-65
-125
-175
-218
-266
-293
-302
-282
-283
-251
-217
-188
-118
-73
3
60
131
181
224
254
274
282
285
287
269
224
182
128
83
-227
-192
-91
-36
-233
-255
-292
34
32
29
11
10
11
11
10
11
10
9
9
9
9
8
9
9
8
8
11
10
10
12
9
8
9
8
10
8
8
8
8
8
8
8
8
8
9
12
11
11
13
10
9
9
11
9
8
8
8
8
8
10
11
12
12
9
9
9
BJD
Phase
Image
Number
RV
(km s−1)
Uncertainty
(km s−1)
2 451 915.45923
2 451 915.46637
2 451 915.47131
2 451 915.47687
2 451 915.48181
2 451 915.48926
2 451 915.49420
2 451 915.49976
2 451 915.50470
2 451 915.51202
2 451 915.51697
2 451 915.52252
2 451 915.52747
2 451 915.53467
2 451 915.53961
2 451 915.55011
2 451 915.56897
2 451 915.57391
2 451 915.57941
2 451 915.58429
2 451 916.43011
2 451 916.43506
2 451 916.44061
2 451 916.44550
2 451 916.45367
2 451 916.45862
2 451 916.46417
2 451 916.50348
2 451 916.50842
2 451 916.51398
2 451 916.51892
2 451 916.52625
2 451 916.53119
2 451 916.53674
2 451 916.54169
2 451 916.54938
2 451 916.55432
2 451 916.55981
2 451 916.56482
2 451 916.57239
2 451 916.57733
2 451 916.58282
2 451 916.58777
2 451 916.59698
2 451 916.60193
2 451 916.60748
2 451 916.61243
2 451 916.63782
2 451 916.64276
2 451 916.64838
2 451 916.65338
2 451 916.66052
2 451 916.66553
2 451 916.67114
2 451 916.67615
2 451 916.68915
2 451 916.69476
2 451 916.69977
2 451 916.70721
0.7239
0.7643
0.7922
0.8236
0.8515
0.8936
0.9216
0.9530
0.9809
0.0223
0.0503
0.0817
0.1096
0.1503
0.1783
0.2376
0.3442
0.3722
0.4032
0.4308
0.2120
0.2400
0.2714
0.2990
0.3452
0.3732
0.4046
0.6267
0.6547
0.6861
0.7140
0.7554
0.7834
0.8148
0.8427
0.8862
0.9141
0.9452
0.9735
0.0163
0.0442
0.0753
0.1032
0.1553
0.1832
0.2146
0.2426
0.3861
0.4141
0.4458
0.4741
0.5145
0.5428
0.5745
0.6028
0.6763
0.7080
0.7363
0.7784
243404
243408
243409
243412
243413
243417
243418
243421
243422
243426
243427
243430
243431
243435
243436
243440
243445
243446
243449
243450
243541
243542
243545
243546
243550
243551
243554
243559
243560
243563
243564
243568
243569
243572
243573
243577
243578
243581
243582
243586
243587
243590
243591
243595
243596
243599
243600
243632
243633
243636
243637
243641
243642
243645
243646
243651
243654
243655
243659
-292
-289
-287
-262
-235
-184
-137
-97
-33
43
63
147
208
251
284
311
253
226
192
135
287
293
290
291
248
220
201
-231
-256
-279
-297
-291
-286
-271
-247
-200
-159
-101
-55
20
63
141
174
242
277
289
286
209
162
105
53
-28
-96
-152
-193
-267
-275
-306
-269
8
8
8
8
8
8
8
9
11
18
39
19
20
21
23
17
14
12
13
12
9
9
9
9
9
9
24
9
8
8
8
11
9
8
8
9
9
10
9
10
9
8
8
8
8
8
8
8
9
9
9
10
14
12
16
11
11
20
25
U Gem 5
Figure 3. The radial velocity curve extracted from the low res-
olution data by cross-correlation of the NaI lines with the M4.0
star GJ213, corrected to the solar system barycentre. The curve
is the best (circular) fit to the data.
4.1 K2
4.1.1 The radial velocity curve
We cross-correlated the KI/TiO (7550-7750A) and NaI
(8160-8230A) regions of each spectrum with that of each
spectral type template. We chose the upper limit of the TiO
region to avoid the feature at 7773A, which is probably OI
(Friend et al. 1988). We then fitted a sine curve to the re-
sulting radial velocity curves, with the period fixed at that of
Marsh et al. (1990). Each velocity was weighted according to
the calculated signal-to-noise in the spectrum from which it
was derived, using the region 7800-8000A for the low resolu-
tion data and 8140-8250A for the high resolution data. Since
we only have relative weights for the data points, we cannot
assign absolute values of χ2, but the relative change in χ2
between the different fits is meaningful. We found that all
the NaI fits have χ2 values within ±2 percent of each other,
but the KI/TiO region gives values 20 to 80 percent larger.
We therefore decided that NaI provides a better solution,
and in Figure 3 we show the radial velocity curve and the
best fit and in Figure 4 the residuals about that fit. Table 1
gives the individual radial velocity points.
We note in passing that the RMS of the fits fell consider-
ably between the initial and final radial velocity curves. This
is presumably because the broadened templates match the
data better, and so yield better cross-correlation functions.
This leads us to believe that the reason we obtained better
results in Beekman et al. (2000) from cross-correlating with
a summed spectrum of IP Peg rather than the template stars
was that the latter were not broadened.
4.1.2 The nature of the deviations
It is immediately obvious from Figures 3 and 4 that there
are systematic deviations from the best fit. There are two
obvious hypotheses which might explain these deviations.
The first is a region of reduced NaI absorption flux around
the inner Lagrangian point. There is observational backing
Figure 4. The residuals about the fit shown in Figure 3. The
error bars are the relative uncertainties scaled to give a χ2
ν of one.
for such an hypothesis, since surface mapping of the sec-
ondary star using the radial velocity data available prior
to the current work, maps the observed apparent eccentric-
ity into a region of reduced NaI absorption (Davey & Smith
1992). Such a region would lead to a lack of NaI absorption
on the hemisphere of the secondary star rotating towards
the observer at phases around 0.4, and away from the ob-
server at 0.6. Since removing flux from the blueward wing of
the line will result in an apparent redshift, this would, ap-
parently, explain the residuals. There are a range of possible
physical causes for such a region, but it is unclear whether
the region needs to be hotter or cooler than the immaculate
photosphere to cause a deficit in NaI absorption. Heating
will cause the overall continuum flux to rise, but will also
decrease the equivalent width of the NaI line. Which effect
will dominate the change in flux is unclear, though the ir-
radiation calculations of Brett & Smith (1993) do yield the
required decrease in NaI flux. Cool spots on the other hand
may be caused by by magnetic activity (see the observations
of Webb et al. 2002, and references therein), or by a com-
bination of limb and gravity darkening. Again the sign of
the effect on the NaI flux is unclear as in the former case
there is is also a gravity effect, and in the latter the line limb
and gravity darkening co-efficients are essentially unknown.
Nevertheless, it appears that a spot resulting in a decrease
in the NaI (and KI) flux at the stellar surface could explain
the observations.
The second obvious candidate is absorption of light from
the secondary star as it passes through the accretion disc.
Littlefair et al. (2001) have demonstrated the importance of
such "mirror eclipses" in the IR, and finding their counter-
part in optical data would clearly be of importance. A mirror
eclipse would create further absorption in the spectrum. At
phase 0.4, the star would be moving behind material moving
away from the observer, and at phase 0.6 material moving
towards them. This excess absorption would again create
the observed effect.
There is a very obvious way to choose between these
models, by comparing the spectra at phases 0.4 and 0.6 with
our summed spectrum taken from phases 0.75 to 0.25. The
6
T. Naylor et al
Figure 5. The mean of the spectra between phase 0.75 and 0.25
(solid line) compared with the mean of spectra in the phase range
0.37 to 0.48 (left) and 0.52 to 0.63 (right).
dark spot model predicts, at phase 0.4, that there will be
a lack of flux on the blue wings of the line, whilst a mir-
ror eclipse would predict extra absorption on the red wing.
Figure 5 shows these spectra, scaled so that the continuum
values are similar to the phase 0.75 to 0.25 sum, which has
been overlaid on the same plots. It is reassuring that with
such a simple scaling the depths of the NaI lines match so
well; it implies that the majority of the variability and of
the flux originates from the secondary star. If this was not
the case the equivalent width of the NaI lines would not be
so constant.
The case at phase 0.4 seems clear-cut. The red wings
of the lines are co-incident, but the blue wing of the phase
0.4 spectrum is consistently redder by between a half and
one pixel (15-30 km s−1). This lack of blue absorption at
this phase supports the idea of a spot. At phase 0.6 the sit-
uation is more ambiguous, with the whole line being shifted
bluewards by between a half and one and a half pixels (15-45
km s−1).
To assess the credibility of the spot model further, we
modelled the radial velocity curve of a Roche-lobe filling
star using a modified version of the code first presented in
Shahbaz et al. (1993). We assigned a radial velocity to each
point on the grid which covers the surface of the secondary
star and used the fluxes from each point to produce a line
profile at that phase. As is usual in modelling rotational
profiles, we normalised the velocity scale for the profile in
terms of v2sin(i). Throughout the modelling we used a mass
ratio q = M2/M1 =1/3, and inclination of 67 degrees. We
calculated the radial velocity deviations the model would
produce from a circular orbit as the first moment of the
profile. As a baseline, we first calculated a model with no
limb or gravity darkening. Such a model is equivalent to as-
suming that the NaI flux is uniform across the surface of the
star, and does not depend on the angle an element is viewed
at. Figure 6 shows the result, with deviations of the correct
order (0.04v2sin(i) ≃12km s−1), but at the wrong phases
(0.25 and 0.75). These deviations are due to the aspheric-
ity of the secondary star. The inner Lagrangian point moves
Figure 6. The radial velocity shifts expected from two differ-
ent models of the secondary star. The solid line is for a uniform
distribution of line flux over a Roche-lobe filling secondary star
for a q of 1/3. The dashed line shows the result of adding gravity
darkening.
faster than the mean surface of the secondary, and so when
on the limb of the secondary moving towards the observer
(as at phase 0.25) it produces a small blue shift with respect
to the expected radial velocity curve. This effect can be ob-
served in our residuals in Figure 4 but is not the dominating
"problem".
We next placed a spot on the inner Lagrangian point.
Whilst we could have achieved this by an arbitrary dimming
of the flux, we chose instead to include gravity darkening
in our model, by using a "standard" gravity darkening co-
efficient for a convective envelope of 0.08 (Lucy 1967). We
did this not with the idea of establishing that gravity dark-
ening is the cause of the dark spot, but simply to establish
whether it has the correct order-of-magnitude. The result of
this is again shown in Figure 6, where it can be seen that it
results in two more peaks in the radial velocity curve resid-
ual, of roughly the same size as those due to the asphericity
of the secondary star, but now at phase 0.4 and 0.6.
Before comparing this model with our residuals we sub-
tracted from it the best fitting sinusoid with a fixed phase
zero to simulate the effect of our fitting process. This has an
amplitude of 0.01v2sin(i). The result is shown in Figure 7,
where we have assumed a v2sin(i) of 115 km s−1 (see Section
4.3). The deviations due to the asphericity of the secondary
star are well modelled, but it is clear we need a somewhat
darker spot to reproduce the data. Furthermore, the dark
spot deviations are obviously asymmetric, which cannot be
explained by this simple model. However, further modelling
lies beyond the scope of this work, our aim is simply to
establish how far these deviations from a perfect sinusoid
affect our derived parameters.
4.1.3 Deciding a value and an uncertainty
The important result from Section 4.1.2 is that the result
of fitting a sinusoid to the radial velocity deviations due to
both the asphericity of the secondary star, and the inner
U Gem 7
ondary star with previous work, it seems best to use the raw
K2, i.e. without the corrections suggested either by ourselves
or Friend et al. (1990). Our value of 302km s−1 should there-
fore be compared with 309 ± 3km s−1 (Friend et al. 1990)
and 283 ± 15 (Wade 1981). Given the uncertainty we have
about our errors, these values are in agreement.
4.2 Spectroscopic Phase Zero
Although this is not important for our work here, we can
also use our radial velocity curve to derive a measurement
of spectroscopic phase zero. We obtain a time of TDB 2 451
915.8618 or HJD 2 451 915.8611. The uncertainty is about
0.0002 day. This lies about 0.0006 day or about 1.5σ away
from the extrapolation of the Marsh et al. (1990) ephemeris,
after taking into account the uncertainties in the ephemeris
itself.
4.3 V2sin(i)
4.3.1 Method
Examination of Figure 4 shows that there are significant
deviations from the radial velocity curve between phases
0.25 and 0.75, as one might expect as a result of an in-
ner Lagrangian dark spot. This implies that the dark spot
significantly distorts the line profile, and so if we are to de-
termine the rotational velocity of the secondary star it is
sensible to omit these phases. Therefore we co-added the
spectra between phases 0.75 and 0.25 using the derived semi-
amplitude, and then fitted them in the same wavelength
regions used to derive K2, to a broadened version of the
spectral type templates, along with a smoothly varying con-
tinuum. We broadened the spectral type template using the
analytical function appropriate for a spherical star.
Each of our spectra is a time average, during which the
secondary star will move along its orbit, adding a further
broadening to the line profiles. To check that we had suc-
ceeded in making our observations short enough to ensure
this effect was negligible we calculated the maximum ex-
tra broadening any spectrum should have (30 km s−1). We
then pre-broadened the spectrum of the spectral-type stan-
dard by this amount, and repeated our fitting procedure.
This changed the derived v2sin(i) by < 4 km s−1, and so
the effects of velocity smearing can be neglected.
As discussed in Section 4.1.2, the non-sphericity of the
secondary star has a significant effect on the radial velocity
curve. Shahbaz (1998) shows that we might also expect it
to affect the line profile broadening, making it significantly
different from the spherical star broadening assumed in our
model. To test this we used our Roche model to create a
series of line profiles from phase 0.25 to 0.75, and averaged
them. We then fitted this profile with our spherical model,
and found that the difference in derived v2sin(i) was less
than 2 percent. This is far smaller than other uncertainties
in our analysis and so can be safely ignored.
4.3.2 Deciding a value and an uncertainty
The statistical uncertainties in our values can be estimated
by creating a χ2 grid for the two parameters, which are the
Figure 7. The solid line is the model for the residuals which
includes both the Roche geometry and gravity darkening (the
dashed line of Figure 6). We have subtracted from this model
the best fitting sine wave with the phase fixed, and then scaled
it using a v2sin(i) of 115 km s−1. The circles are the residuals
presented in Figure 5.
Lagrangian dark spot is about 0.01v2sin(i), or about 1.2 km
s−1. Clearly this is not the entire story, it looks as though the
dark spot is darker than our predictions, and the asymme-
try remains unexplained. We therefore apply our predicted
correction, since the effects modelled in it should be present,
but we should also be aware that there are probably further
corrections of a similar order. If we simply take a mean of
the NaI values in Table 2 and make this correction our final
answer is 300 km s−1. Deciding an uncertainty for this value
is not trivial. We can derive a statistical uncertainty for a
the sine fit by scaling the relative error bars to yield a χ2
ν of
1, which yields around 1 km s−1. Of course, if we included
the effects of asphericity and the dark spot in our model the
size of error bar required to reach a χ2
ν of one would decline,
along with our final uncertainty estimate. Conversely, it is
clear that our symmetric model will never fit both the devi-
ations at phase 0.4 and 0.6 to better than 5km s−1. Even if
we used that as an uncertainty, it would never dominate the
uncertainties in our final parameters for U Gem. Therefore
we will ignore the uncertainty in K2 throughout this paper,
and leave any future users of this measurement to use the
above discussion to choose an appropriate uncertainty for
their use.
4.1.4 Comparison with previous values
We can now, finally, understand the apparent eccentricities
found by Friend et al. (1990) and Wade (1981) of 0.027 and
0.086 respectively. If we fit our own data with an elliptical
orbit we also obtain an eccentricity of about 0.024, with a
significant decrease in the residuals caused by the inner La-
grangian dark spot. So, it seems clear that the measured
eccentricities in the orbit of U Gem are caused by these ef-
fects, but that the correction for this, derived in the previous
section seems to be small, ∼1 km s−1. Therefore, to compare
our value for the radial velocity semi-amplitude of the sec-
8
T. Naylor et al
fraction of the light from the secondary star and the rota-
tional broadening. At each point in the grid we calculated
the value of χ2, and then re-normalised the grid so that the
best fitting model gave a χ2
ν of one. Our 1σ uncertainty is
then simply the range of values of v2sin(i) which lie within
a χ2 (not χ2
ν ) of 2.3 of the best fit. This procedure is con-
servative in the sense that the calculated uncertainty would
be smaller if we did not rescale χ2. This gives uncertainties
around 10 km s−1 for either of the two regions used at low
resolution, or the high resolution data.
In addition to the statistical uncertainty, we must also
consider the fact that there is a very obvious correlation be-
tween the values of v2sin(i) given in Table 2 and the spectral
type of the template used for the cross-correlation. Later
spectral types gave smaller values of v2sin(i), presumably
because the lines in these stars are closer to saturation, and
so need less broadening to match the width of the lines of U
Gem. As we shall discuss in Section 6.1 we cannot use the
spectral type of the secondary star as a guide, and so we
are forced to look at the values of χ2
ν to decide which fits
are the most appropriate. The best NaI fits cover the range
120-130km s−1, whilst KI/TiO cover 105-130 km s−1. We
therefore adopt a value of 115±15km s−1 for v2sin(i), where
the uncertainty is driven by not knowing the correct tem-
plate to use, as this is larger than the uncertainty derived
from the χ2 analysis.
4.4 γ2
4.4.1 Method
To obtain γ2 we calculated the difference in velocity between
each spectral type template and the mean, velocity shifted U
Gem spectrum it produced, and then added the barycentric
radial velocity of the template obtained from the literature.
The obvious course is to take a simple mean of all the derived
γ velocities, but before doing so we will examine the values
obtained as a function of both lines measured, and spectral
type standard used.
The most obviously discrepant values are those given
by GJ1111. As indicated by the values of χ2
ν in Table 2, the
broadened template for this star is a very bad fit to the data.
The next most discrepant values are those for GL490B. Here
the discrepancy is almost certainly due to the γ velocity as-
sumed for GL490B. Most of the template star velocities are
from Gizis et al. (2002), and have a accuracy of better than
1.5 km s−1, but two objects (GJ463 and GL490B) are from
Reid et al. (1995). Although Reid et al. (1995) estimate the
accuracy of their velocity measurements at 10km s−1, we cal-
culated the differences between a given star's measured ve-
locity in Reid et al. (1995) and that in Marcy et al. (1987).
The velocities in the latter have uncertainties of less than
1km s−1. This gave an RMS of 14km s−1 for the 58 stars.
Using this uncertainty, the measurements for GL490B are in
agreement with the other measurements, excepting GJ1111.
To check that different spectral features did not give
discrepant values for γ2 we could have taken a simple mean
of the velocities for the five remaining objects (four in the
case of the higher resolution data) for each wavelength range.
However, if we are to exclude GL490B on the grounds it has
a poor determination of its systemic velocity, we must also
exclude GJ463. After removing this star we see from the
RMS (which is, admittedly derived from a small number
of values) that the KI/TiO measurements are in agreement
with both the high and low resolution NaI velocities.
4.4.2 Deciding a value and an uncertainty
Given the above discussion, it seems we can take a very
direct approach and simply take the mean of all the mea-
surements for the four good radial velocity standards. This
gives a final measurement of 29±6 km s−1, where the uncer-
tainty is simply the standard error. In principle the errors in
the measurements of the velocities of the standards will add
correlations to our 11 measurements, but this is small com-
pared with our final uncertainty. We may also have added
a little noise by including the high resolution data (which
originate from fewer spectra), but as we shall see later, the
final uncertainty in the gravitational redshift of the white
dwarf is governed not by this error, but by the uncertainty
in γ1.
4.4.3 Comparison with previous work
One of the most important aims of this work was to decide
between the two contradictory measurements for γ2 of 46±6
km s−1 Friend et al. (1990) and 84.9 ± 9.9 Wade (1981).
Our data clearly favour the Friend et al. (1990) result, the
difference between the two being at the 2σ level.
4.5 Final dynamical parameters
We summarise our values for the dynamical parameters
as follows. K2=300 km s−1, with an uncertainty which is
unclear, but small enough to have an insignificant impact
on what follows. v2sin(i)=115±15km s−1, where the uncer-
tainty in the appropriate spectral type to use dominates our
error budget. γ2=29±6 km s−1, where the signal-to-noise in
our spectra is probably responsible for the majority of the
uncertainty.
5 PHYSICAL PARAMETERS
We can now use our measurements of the dynamical param-
eters, to derive the physical parameters of interest.
5.1 The Mass Ratio
ratio M2/M1
simply the ratio K1/K2.
The mass
Long & Gilliland (1999) give K1 = 107.1±2.1 km s−1, which
when combined with our value of K2 yields 0.357±0.007.
is
5.2 The Inclination
The eclipse morphology of U Gem gives quite tight con-
straints on the inclination. The eclipse is thought to be that
of the bright spot and edge of the disc, not the white dwarf
(Warner & Nather 1971), an idea supported by the delay
between spectroscopic phase zero and mid-eclipse (Wade
1981). For a q of 0.357, the system would eclipse the white
dwarf if i >74 degrees. For our derived mass ratio we expect
the disc to be no larger than 0.66RL1 (Paczynski 1977), a
Table 2. Measured Parameters. Columns 1-3 give the spectral type, name and assumed γ velocity for each spectral type standard.
Columns 4-5 give the values of K2 derived from the NaI and KI/TiO features in the lower resolution spectra. Columns 6-11 give the
results for velocity broadening the spectral type templates to match the mean U Gem spectrum. For each parameter we give the values
derived from the NaI and KI/TiO features in the lower resolution spectra, and from the NaI feature in the higher resolution spectra.
Thus columns 6-8 give the best fitting values of v2sin(i) and (in brackets) the associated χ2
ν , and columns 9-11 the implied fraction of
the light originating from the secondary star. Columns 12-14 give the derived values of γ2 for U Gem, again from the lower resolution
NaI and KI/TiO range, and the higher resolution NaI range.
U Gem 9
% from
secondary
Hi
(8)
NaI KI/TiO
(9)
(10)
Hi
(11)
γ2
(km s−1)
NaI KI/TiO
(12)
(13)
Hi
(14)
Radial Velocity standard
Type Name
γ
K2
(km s−1)
(km s−1) NaI KI/TiO
(1)
(2)
GJ463
GL109
GL490B
GJ213
GL51
M3
M3
M4
M4
M5
M5.5 GL65A
M6
GJ1111
(3)
21
30
-23
106
-7
22
8
(4)
302
301
301
301
302
301
301
(5)
305
303
303
303
300
296
289
v2sin(i)
(km s−1)
KI/TiO
(7)
143 (2.8)
140 (2.4)
116 (2.3)
129 (2.3)
105 (2.2)
96 (2.7)
128 (3.4)
NaI
(6)
142 (3.1)
139 (2.9)
130 (2.3)
132 (2.7)
119 (2.6)
118 (2.9)
110 (3.1)
Mean (GL109, GJ213, GL51, GL65A)
Standard error
116
104
76
98
58
61
50
109
84
51
64
38
31
26
121 (1.0)
118 (1.1)
106 (1.1)
110 (1.1)
38
30
11
18
24
25
43
24
5
43
34
18
30
37
39
66
34
4
77
67
53
44
,
30
25
28
48
27
3
(4)
(5)
M1 =
K 3
2 P
2πG sin3(i) (cid:16)1 +
2
K1
K2(cid:17)
which yields
M1 =
0.91 ± 0.01M⊙
sin3(i)
.
The uncertainty here should be treated with some scepticism
as it includes just the uncertainty in K1, not that from K2,
for reasons discussed in Section 4.1.3. Given our range of
inclinations, this leads to a range for M1 of 1.02 to 1.32M⊙.
The radius of the white dwarf can now be determined
from the gravitational redshift (γ1 − γ2) as
R1 =
GM1
c(γ1 − γ2)
,
(6)
where c is the velocity of light. Long & Gilliland (1999) mea-
sured γ1 as 172±15km s−1. If we subtract from this our value
for γ2 of 29 km s−1 we derive a gravitational redshift for the
white dwarf of 143±15km s−1. We use this in Figure 8 to
summarise the constraints on M1.
6 DISCUSSION
6.1 The Secondary Star
The first question to settle is whether the star lies on the
main-sequence mass radius relationship. Although small, the
uncertainties cover all the mass-radius relationships given in
Figure 2 of Smith & Dhillon (1998). This is still true if we
tighten the mass constraint to that given by using the tighter
limits on the inclination given by Smak (2001).
Where the problem lies with the secondary star is the
spectral type. In Figure 9 we plot the spectral type of field
M-stars given in Reid et al. (1995) and Hawley et al. (1996)
which have masses in Delfosse et al. (2000) against mass.
Given its mass, we would expect the secondary in U Gem
to be an M2 star. Interestingly that is exactly the conclu-
sion reached by Harrison et al. (2000) on entirely different
radius which would just be eclipsed if i=62 degrees. The in-
clination must, therefore, lie in the range 68±6 degrees. Con-
veniently, the mid-point of this range is close to the 69±2
degrees given by the most recent of determinations of the
inclination from detailed bright spot analysis (Smak 2001,
and references therein).
5.3 The mass of the secondary star
From Kepler's Law it is easy to show that the mass of the
secondary star is
M2 =
K 3
1 P
2πG sin3(i) (cid:16)1 +
K2
K1(cid:17)2
.
Given our values derived above, this gives
M2 =
0.34 ± 0.02M⊙
sin3(i)
,
(1)
(2)
where the uncertainty originates from the uncertainty in K1.
If we wish to derive a value independent of sin(i), then taking
the range of values of the inclination from above, we find
that M2 = 0.44 ± 0.06, since we can ignore the uncertainty
in K1. It is worth emphasising that this is an absolute limit,
not a 1σ uncertainty.
5.4 The radius of the secondary star
The radius of the secondary is
R2 =
P vsin(i)
2πsin(i)
= 0.40 ± 0.05R⊙/sin(i).
(3)
In fact the uncertainty in sin(i) (4 percent) is smaller than
the uncertainty in our measurement of v2sin(i), so we take
the mid-point of the range in sin(i) to obtain R2 = 0.43 ±
0.06R⊙.
5.5 The mass and radius of the primary
The analogous formula to equation 1 for the primary is
10
T. Naylor et al
Figure 8. The constraints in the mass-radius plane for the white
dwarf. This is essentially an updated version of Figure 7 of Long &
Gilliland (1999) and Figure 4 of Long (2000). The gravitational
redshift and inclination constrain the white dwarf to lie within
the box delineated by the solid lines. The white dwarf mass is
primarily limited by the measurement of K2 and the inclination
limits of 62 to 74 degrees. (There is, in addition, a weaker depen-
dence on q.) Our measurement of the gravitational redshift to the
white dwarf restricts the white dwarf to lie between the two lines
marked γ2 − γ1. The Hamada-Salpeter mass-radius relationship
intersects the box, but is somewhat above the radius predicted
by the inclination of 69 degrees which originates from studies of
the bright spot. The photometric radius (see text) lies between
the two horizontal dotted lines, well above the dynamical limits.
Figure 9. The mass of M-type field stars as a function of spectral
type from Reid et al. (1995) and Hawley et al. (1996) The dotted
line is an unweighted linear fit.
grounds. They showed that the absolute magnitude of the
secondary star, and the optical/IR spectral energy distribu-
tion were well matched by an M2 star with a white dwarf
and a faint, power-law accretion continuum. The problem
arises because our best fits of spectral type standards to
the spectrum of U Gem imply a spectral type of M4 or M5
(all our spectral types are also from Reid et al. 1995). This
is perfectly consistent with the results obtained by similar
methods by other workers; M4+ (Friend et al. 1990) and
M4.5 (Wade 1981; Stauffer et al. 1979).
One may be able to solve this inconsistency if the sec-
ondary star in U Gem has a high metallicity. Figure 9
shows that there is a spread in spectral type at any given
mass. For example, there is an M3.5 star with a mass of
0.415M⊙, which if it were the analogue of the secondary
star in U Gem may come close to solving the discrepancy.
It is GJ2069A, actually a binary with a mean mass of
0.415M⊙ (Delfosse et al. 1999), and a spectral type of M3.5
(Reid et al. 1995). Delfosse et al. (2000) suggest that the
late spectral type is due to high metallicity, with [M/H]=0.5.
One could then explain our cross-correlation results as fol-
lows. If the star is of high metallicity, the lines will be closer
to saturation than one might expect for an M2 star. As such
the line shapes and relative depths of the bandheads will be
closer to that of a later type, solar metallicity star, which
has stronger lines. The χ2 fitting procedure, which attempts
to match line shapes and bandheads (but not the equivalent
widths because of the accretion continuum) will give a lower
χ2
ν for the later spectral type star. Further evidence that this
is the case comes from the fraction of the light which our fit-
ting procedure gives as originating from the secondary star
(see Table 2). The fraction increases with earlier spectral
types, already reaching 100 percent for M3, and presumably
would be even higher for an M2 spectral type standard. Such
behaviour is simply explained if the secondary is an M2 star
with line strengths enhanced by a high metallicity.
There is one potential problem with the high metallic-
ity explanation, in that it would also reduce the absolute
magnitude. Ribas (2003) does indeed find that the absolute
K-band magnitude of GJ2069A is 0.35 mag fainter than
stars of comparable mass. However, such a small change in
absolute magnitude corresponds to about a third of a spec-
tral sub-class, and so would not affect the conclusions of
Harrison et al. (2000), based on the absolute magnitude of
U Gem.
In summary, the observations are consistent with the
idea that the secondary star in U Gem is an M2 dwarf with
high metallicity. In terms of mass, radius and luminosity it
is indistinguishable from similar field M dwarfs.
6.2 The broader "cool secondary" problem
There have been suggestions for many years that the sec-
ondary stars in CVs may be too cool for their mass. If we
extrapolate from the observations of U Gem, then the prob-
lem is not that they are too cool, but simply have a spectral
type which is later than that for a solar metallicity star of the
same effective temperature. Smith & Dhillon (1998) appar-
ently showed that such claims were incorrect, by comparing
the mass-spectral-type relation for CV secondary stars with
that for field stars, derived from eclipsing binaries. Their
argument was that the spread in spectral types of CV sec-
ondaries at a given mass was matched by the spread in
field stars. For the field M-stars, the spread they referred
to was driven by the high mass for the probably-metal-rich
GJ2069A. Conversely, the spread in spectral types for the
CV secondary stars was driven by the high masses of two
objects. First U Gem, whose mass this work revises down-
U Gem 11
the Hamada-Salpeter mass-radius relationship. To make all
three consistent implies that the "preferred" inclination of
69 degrees (see Section 5.2) seems to be a little too high.
However, this conclusion should be tempered with the re-
alisation that field white dwarfs do not appear to follow
this relationship closely. Provencal et al. (1998) show that
a large number of white dwarfs apparently fall below this
relationship, for reasons which are not clear. The factor is
up to about 20 percent, which would be sufficient to solve
the discrepancy.
One can also estimate the radius from the FUV spec-
trum and the distance, as was discussed by Long & Gilliland
(1999) and Long (2000). Their results, scaled to re-
flect the new (and larger) astrometric distance of 100pc
(Harrison et al. 2004) are also plotted in Figure 8, and in-
dicate a white dwarf radius of 5.7 × 108cm, which we shall
refer to as the photometric radius. The uncertainties in this
value are small. The uncertainty in the parallax is about 5
percent, and hence contributes 5 percent to the uncertainty
in radius. The overall flux level of the HST spectra is uncer-
tain at the 10 percent level, which translates to a further 5
percent uncertainty in radius. This yields an overall uncer-
tainty of 7 percent, or a radius of 5.7 ± 0.4 × 108cm. This
figure is inconsistent with our value of the radius derived
from γ2 − γ1. This discrepancy is not straightforward to re-
solve. Lower inclinations give a larger white dwarf radius,
so if we take the lowest possible inclination, we still find
that γ2 − γ1 implies a radius of only 4.13 ± 0.44 × 108cm,
which differs from the photometric radius at the 2.5σ level.
Worse still this would place the white dwarf well above the
Hamada-Salpeter radius. If we wish to make the photomet-
ric radius consistent with the Hamada-Salpeter radius we
require γ2 − γ1=80km s−1, which is far outside our error
bars.
One possible solution is to posit another UV component
in the spectrum (Long et al. 1993). The UV flux in U Gem
declines slowly during quiescent intervals. However, as was
first noted by Kiplinger et al. (1991), the decline in flux is
not consistent with the temperature derived from modelling
the spectra in terms of a uniform temperature white dwarf.
When modelled as a single temperature white dwarf the tem-
perature of the white dwarf is hottest and the radius smallest
just after outburst. Long et al. (1993, 1995) suggested, based
on 850-1850A spectra obtained with the Hopkins Ultravio-
let Telescope, that the discrepancy might be resolved if the
outburst left behind a hot accretion belt on the surface of
the white dwarf that slowly decayed. Other possibilities for
the second component include emission caused by heating of
the white dwarf due to ongoing accretion and/or emission
from the inner disc. The inclusion of a second component
does sometimes reduce the estimated radius of the white
dwarf. However, it is difficult to make such a second com-
ponent quantitatively plausible. For example a white dwarf
with a temperature of 30,000 K and a radius of 4 × 108 cm
at a distance of 100 pc would produce only about half of
the observed UV flux. As a result, to reduce the inferred
radius from 5.7 × 108 to 4 × 108cm requires either that the
second component contributes about the same flux as the
first component, or that the second component causes the
temperature of the first component to be underestimated
by order 10,000K, or some combination thereof. Based on
our experience and the fact that the white dwarf in U Gem
Figure 10. Mass of M-type secondary stars in CVs as a function
of spectral type from Smith & Dhillon (1998) with revisions for
U Gem from this work and IP Peg from Beekman et al. (2000).
The dotted line is an unweighted linear fit to the field star data
of Figure 9.
wards by 30 percent, and secondly IP Peg, whose mass we
revised downwards by almost a factor two in Beekman et al.
(2000). We have plotted therefore, in Figure 10, the spectral
types for CVs with well determined secondary star masses
from Smith & Dhillon (1998) with the revisions for U Gem
and IP Peg.
Compared with the data available to Smith & Dhillon
(1998), there are now many more accurate field star masses,
which we have represented in Figure 10 by an unweighted
linear fit. The data suggest that, on average, CV secondaries
show a tendency to be over-massive for their spectral types.
For example, two of the best determinations, IP Peg and
U Gem lie at later spectral types than any field object of
similar mass. However, this tendency is small (up to about
2 sub-types) and we agree with Smith & Dhillon (1998) that
it is the spread in both the field and CV mass/spectral-type
relation, rather than any systematic shift, which precludes
the use of mass to spectral-type conversions.
There is current theoretical interest in explaining dif-
ferences between the observed spectral type and the spec-
tral type predicted by the mass of the star (e.g. Kolb et al.
2001). Our results suggest strongly that any such compari-
son should be carried out between the models and the pho-
tometric spectral type, not the spectroscopic one.
6.3 The White Dwarf
We can use our data to test whether the white dwarf prop-
erties are similar to those of field white dwarfs. From our
data we can test whether it follows a similar mass radius
relationship, and whether its luminosity is that expected for
its radius and temperature.
To begin with the mass and radius, we have placed on
Figure 8 the Hamada-Salpeter mass-radius relationship as
parameterised in Anderson (1988). The range of masses and
radii implied by the gravitational redshift in combination
with the estimates of the inclination are consistent with
12
T. Naylor et al
is by far the best studied white dwarf in a dwarf nova, we
are sceptical the errors in the existing spectral analyses are
large enough to easily reconcile the two ways of estimating
the radius of the white dwarf in U Gem.
In summary, the mass and radius of the white dwarf
deduced from the gravitational redshift and binary inclina-
tion are consistent with those for field stars, but inconsistent
with the radius deduced from the astrometric parallax, tem-
perature and observed UV flux.
ACKNOWLEDGEMENTS
The Isaac Newton Telescope is operated on the island of
La Palma by the Isaac Newton Group in the Spanish Ob-
servatorio del Roque de los Muchachos of the Instituto de
Astrofisica de Canarias. We thank Stuart Littlefair for useful
discussions, and the referee Robert Smith for a careful read-
ing and useful suggestions. Computing was performed on the
Exeter node of the Starlink network, funded by PPARC.
6.4 Determining q using v2sin(i)
A technique frequently used in studies of cataclysmic vari-
ables is to measure q using v2sin(i) and K2. This uses the
fact that
v2sin(i)
K2
= (1 + q)
R2
a
.
(7)
Since we have a measure of q which is independent of
v2sin(i) we can test the validity of this method. We find that
v2sin(i)/K2 = 0.38 ± 0.05. If we now use the approximation
for R2/a due to Eggleton (1983) we find that q = 0.41±0.03,
which is consistent with (though less accurate than) the
value determined using K1, giving us confidence in parame-
ters determined for this and other systems using v2sin(i).
7 CONCLUSIONS
Our primary conclusions are as follows.
• The mass and radius of the white dwarf in U Gem de-
termined from the gravitational redshift and inclination are
indistinguishable from a field white dwarf. However, the ra-
dius deduced from the UV spectrum and astrometric paral-
lax is inconsistent with this kinematic determination. Fur-
ther analysis and/or observations are required to understand
whether the kinematic and spectroscopic information can
be reconciled. This is important since U Gem is the proto-
typical dwarf nova.
• The secondary star in U Gem is indistinguishable in
mass and radius from a field M2 dwarf. However, the spec-
troscopic spectral type is later than might be expected, but
this can be explained if it is of higher than solar metallicity.
• The M-stars in cataclysmic variables seem, as a group to
be around 1 spectral sub-type cooler than might be expected
from their mass or radius. There is a scatter about the mean
mass/spectral type relationship of about a sub-type.
On our way to reaching these conclusions we have dis-
covered the following.
• We have tested the determination of the mass ratio us-
ing the rotational velocity of the secondary star, and shown
(at least in the case of U Gem) it gives answers consistent
with other data.
• However, the uncertainty in the derived rotational ve-
locity is dominated by the uncertainty in choosing which
spectral type star should used for the cross-correlation.
• The uncertainty in the radial velocity semi-amplitude
of the secondary star is dominated by effects at the few km
s−1 level which are due to the asphericity of the secondary
star and non-uniformities in the distribution of the line flux
over its surface.
REFERENCES
Anderson N., 1988, ApJ, 325, 266
Beekman G., Somers M., Naylor T., Hellier C., 2000, MN-
RAS, 318, 9
Brett J. M., Smith R. C., 1993, MNRAS, 264, 641
Davey S., Smith R. C., 1992, MNRAS, 257, 476
Delfosse X., Forveille T., Mayor M., Burnet M., Perrier C.,
1999, A&A, 341, L63
Delfosse X., Forveille T., S´egransan D., Beuzit J.-L., Udry
S., Perrier C., Mayor M., 2000, A&A, 364, 217
Eggleton P. P., 1983, ApJ, 268, 368
Friend M. T., Martin J. S., Smith R. C., Jones D. H. P.,
1990, MNRAS, 246, 637
Friend M. T., Smith R. C., Martin J. S., Jones D. H. P.,
1988, MNRAS, 233, 451
Gizis J. E., Reid I. N., Hawley S. L., 2002, AJ, 123, 3356
Harrison T. E., Johnson J. J., McArthur B. E., Benedict
G. F., Szkody P., Howell S. B., Gelino D. M., 2004, AJ,
127, 460
Harrison T. E., McNamara B. J., Szkody P., Gilliland R. L.,
2000, AJ, 120, 2649
Hawley S. L., Gizis J. E., Reid I. N., 1996, AJ, 112, 2799
Horne K., 1986, PASP, 98, 609
Kiplinger A. L., Sion E. M., Szkody P., 1991, ApJ, 366, 569
Kolb U., King A. R., Baraffe I., 2001, MNRAS, 321, 544
Littlefair S. P., Dhillon V. S., Marsh T. R., Harlaftis E. T.,
2001, MNRAS, 327, 475
Long K. S., 2000, New Astronomy Review, 44, 125
Long K. S., Blair W. P., Bowers C. W., Davidsen A. F.,
Kriss G. A., Sion E. M., Hubeny I., 1993, ApJ, 405, 327
Long K. S., Blair W. P., Raymond J. C., 1995, ApJ, 454,
L39+
Long K. S., Gilliland R. L., 1999, ApJ, 511, 916
Lucy L. B., 1967, Zeitschrift fur Astrophysics, 65, 89
Marcy G. W., Lindsay V., Wilson K., 1987, PASP, 99, 490
Marsh T. R., Horne K., Schlegel E. M., Honeycutt R. K.,
Kaitchuck R. H., 1990, ApJ, 364, 637
Paczynski B., 1977, ApJ, 216, 822
Provencal J. L., Shipman H. L., Hog E., Thejll P., 1998,
ApJ, 494, 759
Reid I. N., Hawley S. L., Gizis J. E., 1995, AJ, 110, 1838
Ribas I., 2003, A&A, 398, 239
Shahbaz T., 1998, MNRAS, 298, 153
Shahbaz T., Naylor T., Charles P. A., 1993, MNRAS, 265,
655
Smak J. I., 2001, Acta Astronomica, 51, 279
Smith D. A., Dhillon V. S., 1998, MNRAS, 301, 767
Stauffer J., Spinrad H., Thorstensen J., 1979, PASP, 91, 59
Wade R. A., 1981, ApJ, 246, 215
Warner B., Nather R. E., 1971, MNRAS, 152, 219
Webb N. A., Naylor T., Ioannou Z., Charles P. A., Shahbaz
T., 2000, MNRAS, 317, 528
Webb N. A., Naylor T., Jeffries R. D., 2002, ApJ, 568, L45
This paper has been typeset from a TEX/ LATEX file prepared
by the author.
U Gem 13
|
astro-ph/9511015 | 1 | 9511 | 1995-11-02T20:44:00 | Hubble Space Telescope Observations of the Lensing Cluster Abell 2218 | [
"astro-ph"
] | We present a striking new Hubble Space Telescope (HST) observation of the rich cluster Abell 2218 taken with the Wide-Field and Planetary Camera (WFPC2). HST's restored image quality reveals a sizeable number of gravitationally-lensed features in this cluster, significantly more than had been identified using ground-based telescopes. The brightest arcs are resolved by HST and show internal features enabling us to identify multiply-imaged examples, confirming and improving the mass models of the cluster determined from ground-based observations. Although weak lensing has been detected statistically in this and other clusters from ground-based data, the superlative resolution of HST enables us to individually identify weakly distorted images more reliably than hitherto, with important consequences for their redshift determination. Using an improved mass model for the cluster calibrated with available spectroscopy for the brightest arcs, we demonstrate how inversion of the lensing model can be used to yield the redshift distribution of $\sim$80 faint arclets to $R\simeq25$. We present a new formalism for estimating the uncertainties in this inversion method and review prospects for interpreting our results and verifying the predicted redshifts. | astro-ph | astro-ph | HUBBLE SPACE TELESCOPE OBSERVATIONS
OF THE LENSING CLUSTER ABELL
J.-P. Kneib, R.S. Ellis, I. Smail, W.J. Couch & R.M. Sharples
) Institute of Astronomy, Madingley Road, Cambridge CB HA, U.K.
) The Observatories of the Carnegie Institution of Washington, Santa Barbara St.,
Pasadena, CA -
) School of Physics, University of NSW, Sydney , NSW Australia
) Dept. of Physics, University of Durham, South Road, Durham DH LE, U.K.
Received
;
accepted
5
9
v
o
N
2
5
1
0
1
1
5
9
/
h
p
-
o
r
t
s
a
{ {
ABSTRACT
We present a striking new Hubble Space Telescope (HST) observation of the rich
cluster Abell taken with the Wide-Field and Planetary Camera (WFPC). HST's
restored image quality reveals a sizeable number of gravitationally-lensed features in
this cluster, signi(cid:12)cantly more than had been identi(cid:12)ed using ground-based telescopes.
The brightest arcs are resolved by HST and show internal features enabling us to
identify multiply-imaged examples, con(cid:12)rming and improving the mass models of the
cluster determined from ground-based observations. Although weak lensing has been
detected statistically in this and other clusters from ground-based data, the superlative
resolution of HST enables us to individually identify weakly distorted images more
reliably than hitherto, with important consequences for their redshift determination.
Using an improved mass model for the cluster calibrated with available spectroscopy
for the brightest arcs, we demonstrate how inversion of the lensing model can be
used to yield the redshift distribution of (cid:24) faint arclets to R ' . We present a
new formalism for estimating the uncertainties in this inversion method and review
prospects for interpreting our results and verifying the predicted redshifts.
Subject headings: cosmology: observations { galaxies: evolution { gravitational lensing
{ {
Introduction
.
The gravitational lensing of faint background galaxies by rich clusters is emerging as a very
promising method to constrain both the distribution of dark matter in clusters and the statistical
redshift distribution of galaxies beyond the reach of conventional spectrographs (Fort & Mellier
). The lensing distortion induced in the image of a typical distant galaxy by a foreground
rich cluster depends upon the product of a scale factor (involving the galaxy and cluster redshifts
and the adopted cosmological model) and the second derivatives of the pro jected cluster potential.
The ma jority of the faint lensed images are only weakly distorted and these are termed \arclets".
However, a small fraction are highly distorted \giant arcs" { images which lie near critical lines
and su(cid:11)er high ampli(cid:12)cation; these are particularly helpful in mass modelling since their relatively
bright magnitudes mean that they can often be studied spectroscopically. With redshifts for one or
more giant arcs in a cluster the absolute mass of the central regions can be accurately determined.
Multiply-imaged sources, even without redshifts, provide additional information on the geometrical
con(cid:12)guration of the potential well in the core regions (Mellier et al. , Smail et al. a).
Recent work has concentrated on clusters with arcs of known redshift and multiply-imaged sources.
In such cases, a robust model of the cluster mass can be constructed, allowing inversion of the
lens equations for the arclet population and yielding the redshift distribution of extremely faint
galaxies.For the well-studied cluster Abell (zcl = :), Kneib et al. ( a) demonstrated a (cid:12)rst
application of this inversion technique by identifying (cid:24) candidate arclets with axial ratios
a=b (cid:24)> : from ground-based images taken in superlative conditions. For each arclet, unlensed
magnitudes and probable redshifts to a limit of B ' were inferred from a detailed mass model
calibrated by the redshift of a giant arc and the properties of various multiple images (Kneib et
al. ). However, this new technique su(cid:11)ered from several uncertainties. Firstly, simple mass
models may ignore substructure in the cluster mass distribution leading to imprecise inversion.
Also, in the absence of spectroscopic or morphological data, some of the candidate multiply-imaged
ob jects used to model the form of the potential may be spuriously identi(cid:12)ed in ground-based data.
{ {
Finally, even in the best ground-based conditions, the limited angular resolution makes it di(cid:14)cult
to distinguish lensed arclets from intrinsically-elongated faint sources and to accurately measure
their shapes; such confusion may lead to contamination of the inverted redshift distribution by
cluster members, foreground spirals, and close galaxy pairs.
Even in its aberrated state the advantages of HST for lensing studies over the best
ground-based telescopes soon became evident (Smail et al. b). Here we illustrate that the
refurbished HST is even more powerful, allowing reliable identi(cid:12)cation of multiple images and faint
arclets. Considerable progress is thus possible with HST in the inversion method developed by
Kneib et al. ( a).
A plan of the paper follows. Section describes the observations and gives a qualitative
description of the HST images, including those lensed features which allow us to improve the
ground-based model of Kneib et al. ( ). Section describes the improved mass model we have
implemented. Starting from the mass model of a cluster, Section introduces the theory of the
lens inversion and discuss the probability distribution of the redshift of a sheared galaxy. The
sources of uncertainty in this inversion are also discussed in the context of observational data.
In Section we present our results on the faint (cid:12)eld galaxy redshift distribution and discusses
both the limitations of comparing such results with model predictions as well as the prospects
for verifying the inverted redshifts with further observations. Section summarises the overall
conclusions of the paper. Throughout this paper, we assume H = km s(cid:0) Mpc(cid:0) , (cid:10) = and
(cid:3) = .
. Observations and Analysis
.. Previous Observations of Abell
Abell (Fig. a), is one of the best-studied rich clusters at intermediate redshift
(zcl = :). Le Borgne et al. ( ) present a detailed photometric and spectroscopic survey of
the cluster and derive a rest-frame velocity dispersion of (cid:27)cl = + (cid:0) km s(cid:0) , indicative of a
{ {
deep potential well. This is supported by a high X-ray luminosity (Lx( .-. keV) = .(cid:2) ergs
s(cid:0) ) and a strong Sunyaev-Zeldovich decrement (Jones et al. , Birkinshaw & Hughes ).
The cluster contains a number of luminous giant arcs, discovered and extensively studied by Pell(cid:19)o
et al. ( , ). Several of the brighter arcs have been observed spectroscopically; redshifts
for these and ground-based colors for other lensed features provide the basic ingredients for the
recent mass model of Kneib et al. ( ). Using four systems of arcs and possible counter-arcs,
tentatively identi(cid:12)ed from ground-based colors (# , # -#, #-# and # in the
numbering scheme of Le Borgne et al. ( )), Kneib et al. ( ) determine a mass distribution
for the cluster core which is bimodal in form and concentrated around the two most luminous
cluster galaxies (# , #, Table ).
.. HST Observations and Photometric Catalogue
Abell was observed by the HST WFPC- camera on September , . Three exposures
totalling sec were taken through the F W (cid:12)lter. Each exposure was shifted relative to the
others by WFC pixels ( . arcsec) providing a partial overlap of the chip (cid:12)elds. After pipeline
processing, standard IRAF/STSDAS routines were employed to shift and combine the frames to
remove both cosmic rays and hot pixels. We discard the PC chip from our analysis because of
its brighter isophotal limit. The (cid:12)nal frame comprising the WFC chips (Fig. a & b) has an
e(cid:11)ective resolution of . arcsec and a (cid:27) detection limit per resolution element of R ' . We
convert our instrumental F W magnitudes into standard R using the synthetic zero point and
color corrections listed in Holtzman et al. ( ). For the color term we choose (V (cid:0) R) ' :
typical of the faint (cid:12)eld population (Smail et al. d). The color correction is + : mag, and
the typical photometric errors of our faintest ob jects, R < :, are (cid:14)R (cid:24) .{ ..
To produce a catalogue of faint arclets from our data we (cid:12)rst analysed the HST image using
the Sextractor package (Bertin , Bertin & Arnouts ). All ob jects with isophotal areas
above pixels ( . arcsec ) at the (cid:22)R = : mag arcsec(cid:0) isophote ((cid:27)/pixel) were selected.
A comparison of the di(cid:11)erential number counts of these images to deep ground-based R counts
{ {
(Smail et al. d) shows a marked excess of galaxies brighter than R (cid:24) : due to cluster
members (Figure ), and a steep roll-over in the observed counts beyond R (cid:24) arising from an
incompleteness which amounts to % in the R = { bin. We thus applied a magnitude limit
of R = giving a total of images over a . sq. arcmin area. A neural-network algorithm
(Bertin & Arnouts ) was used to separate stars and galaxies leading to the exclusion of
star-like ob jects from the catalogue.
From this list we selected a sample of candidate arclets, (cid:12)rst removing all galaxies with
R < : (probable cluster members) and ob jects lying in the halos of giant ellipticals and very
faint ob ject R > as their photometry and shapes are uncertain. The procedure reduced our
catalog to (cid:24) arclet candidates.
.. Multiply-Imaged Features
At this stage, it is useful to review the multiply-imaged features identi(cid:12)ed on the HST image
in the context of the ground-based predictions, prior to using them to improve the mass model of
Kneib et al. ( ).
Four bright arcs and counter-arcs were identi(cid:12)ed as matching images by Kneib et al. ( )
on the basis of their ground-based colours. Each of these is clearly resolved by HST with internal
structures that enable us to verify their multiply-imaged nature (see Fig. c of Smail et al. b).
We discuss each of these images here and summarise their photometric properties in Table .
# and # # is a most impressive arc system with an internally-symmetric pattern
of unresolved knots showing that this image is clearly formed from the merger of two images of
reversed parity. This enables the location of the critical line to be accurately identi(cid:12)ed. The
knots, which presumably represent HII regions in a blue star-forming galaxy, can also be seen in
the counter image #. A further feature of interest is the train-track-like morphology of the
source, also replicated in #.
# , #, # and # The red arc # has a spectroscopic redshift of z = : and
{ {
shows no internal structure; this is consistent with its identi(cid:12)cation as a background spheroidal
galaxy. It was naturally interpreted as a fold arc i.e. two merging images (Kneib et al. ) with
a single counter-image #. The absence of a strong discontinuity (even in the HST image) in
the surface brightness along the # arc can be explained if the surface brightness peak lies just
outside (or on) the caustic on the source plane. However, a detailed inspection of the HST image
demonstrates that this simple picture is unlikely to be correct as a faint extension of # is now
revealed which merges with #. From the ground-based data it was noticed that # & #
had similar colours to # but no simple model was able to explain such a con(cid:12)guration.
If # is indeed a counter-image of arc # , then we may consider whether # is also a
counter-image. In Section , we will show that, by incorporating individual cluster galaxies in the
mass model, it is straightforward to show that # is a fold-arc with #, # and # each
as counter-images.
# In contrast to # , the blue arc # with a spectroscopic redshift of z = : exhibits
a large amount of internal structure. Although the arc is luminous and therefore probably highly
magni(cid:12)ed, the bright southern end does not appear particularly strongly sheared and is apparently
not multiply-imaged. Close inspection of the northern section of this arc indicates that it extends
across the halo of the cluster galaxy #. The complex morphology can thus be explained via
a background galaxy straddling the caustic. The ma jority of the source lies outside the caustic
producing a single highly magni(cid:12)ed, weakly sheared image. The portion within the caustic is
multiply-imaged and produces the highly elongated tail across the halo of #.
A detailed examination of the HST image reveals several new potentially-important
multiply-imaged systems.
# This very faint thin arc was suggested as a possible lensed feature in the ground-based data
but is clearly veri(cid:12)ed as such by HST. The faintness makes it di(cid:14)cult to identify the individual
sub-components at this stage, although a number of bright knots are visible. Nevertheless, the
structure suggests a likely cusp arc as three components can be distinguished.
H{, H{ These are two impressive multiply-imaged systems which were unrecognised in the
{ {
ground-based studies (Fig. b). From the morphologies and positions, H{ appear to be three
images of a section of the disk of #, the remainder of the source being only singly-imaged. The
very faint features H{ ( R(cid:25)) are believed to represent a new very faint multiply-imaged pair.
Several candidates for the counter-image to this pair exist on the opposite side of the cluster.
# + H. # is a fold arc (two merging images) with H as a counter image.
In summary, the HST image not only allows us to con(cid:12)rm the lensed features that underpin
the ground-based mass model, but also provides additional information that enables us to re(cid:12)ne
the model. We have identi(cid:12)ed a total of multiply-imaged sources seen through the core of
Abell . This is a substantial improvement over the ground-based tally and signi(cid:12)cantly more
than the number known in any other cluster at this time. By analysing these features we can
thus hope for the most detailed view of the mass distribution within a cluster thus obtained. The
model re(cid:12)nements derived from these new multiply-imaged features are principally in the detailed
form of the mass model and lead to little change in the global cluster mass/light ratio. However,
they can have an e(cid:11)ect on the lensing inversion and we will explore this further in Section .
.. Arclets and Shear
Using our previously-de(cid:12)ned catalogue of faint sources ( x.), we now construct a \shear (or
deformation) map" de(cid:12)ned as the local average of the deformation vector (see x) of the lensed
galaxies:
< (cid:28) > (x; y ) = Z Z (cid:28) I (x ; y )! (x (cid:0) x ; y (cid:0) y )dx dy ;
()
where ! (x; y ) is a normalized weighting function. The weighting function chosen was a gaussian
of arcseconds FWHM. Fig. shows the deformation map within the (cid:12)eld of the WFC upon
which we have superimposed the location of some of the most luminous cluster members. This
map provides a view of the cluster potential with a resolution of arcsecs ((cid:24) kpc), allowing
us to detect any substructure in the cluster mass distribution on scales larger than the resolution.
We now use this information to assist in the construction of a re(cid:12)ned model for Abell taking
{ {
into account the new details on the multiple images discussed in x..
It is important to recognise that this map di(cid:11)ers from the true deformation for a number of
reasons including contamination by foreground and cluster galaxies and also the dependency of
the strength of the image deformation on redshift. The former e(cid:11)ect has already been minimised
by removing the brightest galaxies from our list. Furthermore, we can assume that the bulk of the
contaminating galaxies are randomly orientated and thus only dilute the modulus of the shear,
without a(cid:11)ecting its form. The earlier ground-based mass model was bimodal in form and centred
on the cD (# ) and #. The shear map in these regions suggests contributions from #
and # should now be included.
Although the shear map is statistical in nature, the HST resolution has encouraged us to
de(cid:12)ne a visually-selected sample of the brighter and larger arclets, intermediate to the bright arcs
reviewed in the previous section. These sources are sheared su(cid:14)ciently that their identi(cid:12)cation
as lensed features is in little doubt and furthermore most of them are within spectroscopic reach.
Their properties are summarised in Table and Fig. compares their distribution with the
\arclets" identi(cid:12)ed from ground-based data (Pell(cid:19)o et al. ). The total number of HST arclets
is now signi(cid:12)cantly increased. Furthermore, the HST data suggests that as many as a third of the
ground-based arclets are close galaxy pairs or misidenti(cid:12)ed edge-on disk galaxies.
In summary, the improved resolution of the repaired HST allows considerable progress to be
made in the identi(cid:12)cation and understanding of lensed features in Abell . The resolution of
the brighter arcs con(cid:12)rms several of the multiply-imaged features suggested from the ground-based
studies. In particular, the fold arc # is now identi(cid:12)ed as a -image con(cid:12)guration, and a
number of new multiply-imaged candidates are revealed. Similarly, the HST image allows us to
identify weakly-lensed features (arclets) with greater reliability, both on an individual basis and
statistically. In both respects, we are better placed to re(cid:12)ne the mass model developed on the
basis of ground-based imaging and to identify arclets for redshift determination.
{ {
. Mass Modelling
The mass modelling method we use is based on the precepts developed by Kneib ( ) which
has now been successfully applied to describe many di(cid:11)erent cluster lenses including MS
(Mellier et al. ), A (Kneib et al. ), Cl (Kneib et al. b), Abell (Kneib et
al. ) and Cl (Smail et al. b).
The basic approach is to use multiply-imaged systems and the mean orientation of the arclets
to constrain an analytical representation of the total mass based upon components associated with
likely centres of mass, i.e. massive cluster galaxies. Each component is described by a minimal
set of parameters: position, ellipticity, orientation, core size and central velocity dispersion. The
associated mass distribution should be approximately isothermal if the central mass is relaxed.
The particular analytical expression used is based on the pseudo-isothermal elliptical mass
distribution (PIEMD) with ellipticity e = (a (cid:0) b)=(a + b) derived by Kassiola & Kovner ( ):
(cid:27)
rcprc + (cid:26) =
Gprc + (cid:26) ;
(cid:6)(x; y ) = (cid:6)
()
with
x
y
(cid:26) =
( + e) +
( (cid:0) e) :
()
This expression has the advantage of describing mass distributions with arbitrarily large
ellipticities. For each component used, we smoothly truncate the elliptical mass distributions (c.f.
appendix of Kassiola & Kovner, ) using a linear combination of PIEMD components:
qrcut + (cid:26) A ;
rcut (cid:0) rc @
(cid:6)(x; y ) = (cid:6) rcrcut
prc + (cid:26) (cid:0)
()
where rcut is the truncation radius (the surface mass density falls as r(cid:0) for r >> rcut ). The total
mass of such a truncated mass distribution is (cid:12)nite and for r >> rcut in the limit where e ! :
Mtot = (cid:25)(cid:6) rcrcut = (cid:25)G (cid:27) rcut :
()
The ground-based mass model for Abell (Kneib et al. ) was based on two ma jor
components associated with galaxies # and #. As discussed in Section ., the detailed
{ {
information now available from the multiple-images (particularly # and its counter-images)
and the (cid:12)ne structure visible in the shear map (Fig. ), encourages us to improve on this model
by incorporating the e(cid:11)ect of halos associated with components around # and # and
other individual cluster galaxies. For each component the center, ellipticity and orientation are
matched to those observed for the associated light distribution (as is the case for those associated
with # and #). However, the dynamical parameters rc ; rcut and (cid:27) for these four main
components are kept as free parameters.
When including galaxy-scale components into our model it is clear that such a re(cid:12)nement
could, in principle, be continued inde(cid:12)nitely. In practice, we included all galaxies with R< . (as
the magnitude increases the mass of each galaxy become small and their lensing e(cid:11)ects become
negligeable). In total, we incorporate halos associated with luminous cluster galaxies into the
mass model. For each halo the ellipticity and orientation match those observed for the galaxy light
distribution. The other mass parameters are scaled according to the galaxy luminosity following
Brainerd, Blandford & Smail ( ).
(cid:27) = (cid:27) (cid:3)(L=L(cid:3))=
()
and
rcut = r(cid:3)cut (L=L(cid:3))=
()
where (cid:27) (cid:3) and r(cid:3)cut are free parameters in the minimization procedure. Furthermore to have a
pro(cid:12)le that is identical from one galaxy to another we scale the core radius r in the same way as
rcut :
r = r(cid:3) (L=L(cid:3))=
()
The mass of individual galaxy scale as the luminosity with:
Mtot = (cid:25)G ((cid:27) (cid:3))r(cid:3)cut (L=L(cid:3))
( )
It is worth emphasising that, by themselves, the individual galaxy halos do not contain
enough mass to reproduce all of the lensed features observed in the cluster. In other words we
{ {
must retain cluster-scale mass components associated with the brighter cluster galaxies (the
central cD (# ), #, # and #).
To constrain the composite mass model we (cid:12)rst de(cid:12)ne a (cid:31) estimator as the quadratic sum
of the di(cid:11)erences between the source parameters (position, orientation and ellipticity) for each set
of multiple images (see Table ), plus the observed shear as represented by the quadratic sum of
(cid:28)pot < (cid:28)I > sin(((cid:18)pot(cid:0) < (cid:18)I >)). We then minimise this estimator by varying the parameters of
the mass model. To stabilise and speed up the convergence we specify the location of the in(cid:12)nite
magni(cid:12)cation point in the fold or cusp images (i.e. the location of the symmetry break in the case
of # or the luminosity peak of # and the saddle between # and # in the case of the
arc # at z= . ).
The best (cid:12)ducial model resulting from the HST data is presented in Table . A contour plot
of the mass distribution is shown in Fig. where the shear-(cid:12)eld is also shown for a source plane at
zS = . This mass model is currently the most detailed derived for a cluster core, the detail only
being possible because of the combination of the high resolution shear map and the large number
of multiply-imaged sources.
Although the di(cid:11)erence between the HST and ground-based mass models is small when
considering global properties such as cluster mass/light ratio, we show in Section that there
can be variations in the lensing inversion for speci(cid:12)c arclets, depending upon their location. The
principle change is in the detailed granularity of the mass distribution leading to a more precise
inversion.
. Gravitational Lensing Formalism
We now turn to the primary purpose of the paper, namely to take our well-constrained mass
model for Abell and use it to derive statistical redshift distributions for the large sample of
faint arclets discussed in x.. In what follows we extend the original discussion of Kneib et al.
( a), developing a formalism for estimating the errors in the inversion redshifts of individual
{ {
galaxies. This will be particularly useful as we have a range of lensed features from relatively
bright arclets, many of which can be recognised as lensed on an individual basis, to fainter images
which can only be treated statistically... General Equations
The gravitational lensing formalism we use is based on the original treatise presented by
Kneib et al. ( a). The lens mapping is described by the transformation:
~uS = ~uI (cid:0) D ~r(cid:30)( ~uI ) ;
( )
where ~uS is the position of the source, ~uI is the position of the image, D is the dimensionless ratio
DLS =DOS and (cid:30) is the pro jected Newtonian potential normalized by =c.
A distant galaxy can be described to the (cid:12)rst order by (cid:12)ve geometrical parameters: its
centroid (xc ,yc ), complex deformation (cid:28) = (cid:28) ei(cid:18) and size s.
The (cid:12)rst moment of the weighed surface brightness (cid:22)(x; y ) distribution gives the position of
the centroid (xc ,yc ):
xc = (cid:22)W Z Z W (x; y ) (cid:22)(x; y )xdxdy
yc = (cid:22)W Z Z W (x; y ) (cid:22)(x; y )ydxdy ;
()
with
(cid:22)W = Z Z W (x; y ) (cid:22)(x; y )dxdy :
()
The weighting function W (x; y ) can be adjusted to minimise the error in the determination of the
centroid.The second order moment matrix M gives the shape of the galaxy ((cid:28) = (cid:28) ei(cid:18) ), i.e. its
equivalent ellipse of ma jor-axis a, minor-axis b and orientation (cid:18):
Mxy Myy ! / R(cid:18) a
b ! R(cid:0)(cid:18) ;
M = (cid:22)W Z Z W (x; y ) (cid:22)(x; y )xixj dxdy = Mxx Mxy
()
where R(cid:18) is the rotation matrix of angle (cid:18). Note that di(cid:11)erent weighting functions can be chosen
in computing the (cid:12)rst and second moment integrals, depending upon which is required with higher
{ {
accuracy. The weighting factor is more critical in dealing with ground-based data than with HST
images because of the e(cid:11)ects of seeing (Bonnet & Mellier , Kaiser et al. , Wilson, Cole &
Frenk ). In our analysis we used the simple weighting function:
W (x; y ) = ( if (cid:22) < (cid:22)I SO
()
if (cid:22) > (cid:22)I SO
The size parameter (s) is de(cid:12)ned as:
s = pdet M / ab:
()
and the deformation matrix D is:D = Mpdet M = (cid:14) + (cid:28)x
(cid:14) (cid:0) (cid:28)x ! ;
(cid:28)y
()
(cid:28)y
where (cid:28) = (cid:28)x+i(cid:28)y = (cid:28) ei(cid:18) is the complex deformation and (cid:14) = p + (cid:28) is the real distortion
parameter. In terms of the ma jor and minor axis these are:
(cid:14) = a + b
(cid:28) = a (cid:0) b
;
:
()
ab
ab
Further, the complex shear g and the complex ellipticity " are de(cid:12)ned as:
g = a (cid:0) b
(cid:14) = + g (cid:28) (cid:3) ;
()
a + b ;
and
" = a (cid:0) b
" = (cid:28)(cid:14) ;
( )
a + b ;
((cid:3) denotes the conjugate of a complex number).
The lensing equation for the moment matrix is given by (Kochanek, ):
MS = a(cid:0)MI ta(cid:0) ;
( )
where the subscript S refers to the source, I to the image and a(cid:0) is the inverse of the ampli(cid:12)cation
matrix, de(cid:12)ned as the Hessian of the lens mapping (eq. ):
(cid:0) (cid:20) (cid:0) (cid:13)x ! :
(cid:0) D@yy (cid:30) ! (cid:17) (cid:0) (cid:20) + (cid:13)x
a(cid:0) = (cid:0) D@xx (cid:30) (cid:0)D@xy (cid:30)
(cid:13)y
()
(cid:13)y
(cid:0)D@xy (cid:30)
{ {
(cid:20) and (cid:13) = (cid:13)x + i(cid:13)y = (cid:13) ei(cid:18)pot are the usual convergence and shear parameters. We denote
(cid:13) = D ~(cid:13) and (cid:20) = D ~(cid:20) to separate the distance and mass e(cid:11)ects. (cid:18)pot is the direction of the shear
(independent of the redshift of the source) and is de(cid:12)ned as:
@xy (cid:30)
tan((cid:18)pot ) =
@xx(cid:30) (cid:0) @yy (cid:30) :
()
Equivalently, for the potential we can de(cid:12)ne the parameters gpot , (cid:28) pot and (cid:14)pot :
gpot
gpot = (cid:13) (cid:0) (cid:20) ;
(cid:14) = + gpot(cid:28) (cid:3)pot :
()
(cid:0) gpotg(cid:3)pot ;
(cid:28)pot =
For a circular source using this notation, the lens transformation gives: gI = gpot, (cid:28)I = (cid:28)pot , etc.
The determinant of eq. gives the lensing transformation of the ob ject's size:
sS = j det a(cid:0) j sI :
()
Dividing eq. by eq. we have the lens equation for the deformation matrix:
j det a(cid:0) j a(cid:0) DI ta(cid:0) :
DS =
()
>From eq. , the lens equation for the complex deformation (cid:28) S is also derived:
sgn(det a(cid:0) )(cid:28) S = (cid:28) I (cid:0) (cid:28) pot (cid:16)(cid:14)I (cid:0) (cid:28)I <(gI g(cid:3)pot )(cid:17) :
()
The inverse equation is found by exchanging the subscripts I and S and the signs of (cid:28) pot and g (cid:3)pot.
This gives:
sgn(det a(cid:0) )(cid:28) I = (cid:28) S + (cid:28) pot (cid:16)(cid:14)S + (cid:28)S <(gS g (cid:3)pot)(cid:17) :
()
A vectorial representation of eq. is shown in Fig. . The complex deformation of the image is
just the vector sum of the intrinsic source shape and the induced deformation from the potential,
corrected in the strong lensing regime by a factor (cid:14)S + (cid:28)S <(gS g(cid:3)pot ). In the weak shear regime
(det a(cid:0) > ), the correction tends to unity and eq. becomes:
()
(cid:28) I = (cid:28) S + (cid:28) pot :
Using the local shear axes eq. reads:
sgn(det a(cid:0) )(cid:28)Sx = (cid:14)pot(cid:28)I x (cid:0) (cid:28)pot(cid:14)I = (cid:14)pot(cid:14)I ("I (cid:0) "pot ) ;
( )
{ {
sgn(det a(cid:0) )(cid:28)S y = (cid:28)I y :
( )
Note that j(cid:28)y j is a conserved quantity under the lens transformation.
.. Distribution in ellipticity and redshift
The source ellipticity distribution can be estimated from deep HST images of (cid:12)elds outside
rich clusters. A large sample of suitable (cid:12)elds are available in the Medium Deep Survey archive
(Gri(cid:14)ths et al. ). Analysis of these (Ebbels et al. ) reveals that the observed distribution
of image shapes for brighter galaxies is well (cid:12)tted by the functional form:
p((cid:28)Sx; (cid:28)S y ) = (cid:25)(cid:27) (cid:28) exp (cid:0) (cid:28) Sx + (cid:28) S y
! ;
()
(cid:27) (cid:28)
This distribution has a maximum at ((cid:28)x; (cid:28)y ) = ( ; ) and is also radially symmetric (because
of the random orientations of unlensed (cid:12)eld galaxies). We stress however that the form of this
distribution does depend strongly upon the size of the galaxies and their magnitudes (Ebbels et
al. ).
Since j(cid:28)y j is conserved by lensing, in the frame of the local shear, we have the conditional
probabilities:
p((cid:28)Sx ; (cid:28)S y ) = p((cid:28)Sx ; (cid:28)I y ) = p((cid:14)pot(zS )(cid:28)I x (cid:0) (cid:28)pot (zS )(cid:14)I ; (cid:28)I y ) = p(z jI ; mass)p(cid:28)Sy ((cid:28)S y ) :
()
In other words, the conditional redshift probability (given the image shape and the mass model)
is simply the source shape probability divided by that of (cid:28)S y . From eq. :
p(cid:25)(cid:27) (cid:28) exp (cid:0) ((cid:14)pot (zS )(cid:28)I x (cid:0) (cid:28)pot (zS )(cid:14)I )
! :
p(z jI ; mass) = p((cid:28)Sx ; (cid:28)S y )
p(cid:28)Sy ((cid:28)S y ) =
()
(cid:27) (cid:28)
which reproduces the intuitive prescription of Kneib et al. ( a) that the maximum of the
redshift probability function for a given image corresponds to the minimum deformation of the
source.When the image is outside the critical line, (cid:28)pot is an increasing function of redshift { with
positive (cid:28)I x if the orientation is within deg of the shear direction and negative otherwise. If
{ {
(cid:28)I x is positive but not too large, p(z jI ; mass) has a maximum for "I x = "pot (zS ) and the most
probable redshift is (cid:12)nite. However, when (cid:28)I x is too large, p(z jI ; mass) is an increasing function
of redshift leading to a most probable redshift, z = . If (cid:28)I x is negative then p(z jI ; mass) is a
decreasing function of redshift, yielding a most probable redshift, z = zlens (see Fig. ). In the
latter two cases no \sensible" estimate of the redshift of the galaxy can be derived.
.. Uncertainties in the redshift determination
We now discuss the uncertainties that arise when determining faint galaxy redshifts with a
gravitational telescope. There are three sources of error: those arising from image shape (errors in
the deformation (cid:28) I ), the lens mass model (errors in (cid:20), (cid:13) and (cid:18)pot ) and statistical errors introduced
by the contamination of the arclet sample by foreground or cluster galaxies. The (cid:12)rst two errors
are concerned with individual arclets, while the third a(cid:11)ects the properties of the sample as a
whole.
...
Individual errors
We begin by considering the relative error in z . Di(cid:11)erentiating gpot = D ~(cid:13)=( (cid:0) D ~(cid:20)), we have:
(z (cid:0) zL )D
z (cid:0) zL = dDD = ( (cid:0) D ~(cid:20)) dgpot
gpot (cid:0) d~(cid:13)~(cid:13) (cid:0) Dd~(cid:20) :
dz
()
D
The term D=(z (cid:0) zL )D is almost proportional to (z (cid:0) zL ), indicating that the accuracy of the
lensing-inferred redshifts is lower at large redshift (see Fig. ).
Moreover, for the maximum of redshift probability function:
"pot = (cid:14)pot d"I x"I x = (cid:14)pot (cid:18) d"I"I (cid:0) tan((cid:18)I )d(cid:18)I (cid:19) ;
gpot = (cid:14)pot d"pot
dgpot
()
and thus the total error in the estimate of the most probable redshift is:
dDD = ( (cid:0) D ~(cid:20))(cid:14)pot ( d"I"I (cid:0) tan((cid:18)I )d(cid:18)I ) (cid:0) d~(cid:13)~(cid:13) (cid:0) Dd~(cid:20) :
()
{ {
We now turn to the likely uncertainties in measured image ellipticities "I and position angles
(cid:18)I . The former has the following e(cid:11)ect:dDD = ( (cid:0) D ~(cid:20))(cid:14)pot d"I"I
:
()
An overestimate of "I leads (in the sub-critical part of the lens) to an over-estimated redshift.
If the image is close to a critical line then a small error in "I produces a large error in z since (cid:14)pot
diverges. However for the bulk of the arclets this is not a problem as we are not in the multiple
image regions and ( (cid:0) D ~(cid:20))(cid:14)pot is in general less than (as (cid:14)pot (cid:24) and ( (cid:0) D ~(cid:20)) < ). For
faint images there is a tendency to under-estimate the image ellipticity at large ellipticity and
over-estimate the ellipticity at small ellipticity, although these e(cid:11)ects can be statistically corrected.
However for small ellipticities and compact ob jects it is di(cid:14)cult to determine the true image
ellipticity.Errors in the measured orientation (cid:18)I :
dDD = (cid:0)( (cid:0) D ~(cid:20))(cid:14)pot tan((cid:18)I )d(cid:18)I :
()
have the same dependence on (cid:14)pot as the ellipticity. However, as the orientation is usually the best
measured characteristic of an image and because the error is symmetrically distributed, the bias is
less serious than for the ellipticities. Nevertheless, when (cid:18)I (cid:24) (cid:25)= ("I x (cid:24) ) the errors can become
very large.
Finally, the errors in the cluster mass model, in ~(cid:20):
dDD = (cid:0)Dd~(cid:20) = (cid:0) d(cid:6)(cid:6)crit ;
( )
and the error in ~(cid:13) :
dDD = (cid:0) d~(cid:13)~(cid:13) :
( )
demonstrate that an over-estimated local mass and shear lead to an under-estimated redshift,
with the dependence between the two being reasonably well-behaved. However, adding galaxy-size
{ {
components in the lens-model can dramatically change the intensity of the shear near a critical
area of the lens, therefore it is important to take these components into account.
An advantage of the Hubble Space Telescope is the stability of the high resolution imaging,
this minimises problems which plague ground-based studies of faint ob ject shapes (seeing,
tracking, (cid:12)eld astigmatism, and their time variability). While the HST capabilities are an order of
magnitude better than the ground-based facilities, the limiting factors for accurate measurement
of the shapes of faint and compact galaxies now become photon noise and pixel-sampling e(cid:11)ects.
A key point in determining reliable redshifts is the absolute calibration of the mass model.
This is best addressed using a number of spectroscopically-con(cid:12)rmed lensed features in the
cluster, while the morphology of the mass can be best determined using the geometry of any
multiply-imaged sources present (Mellier et al. , Kneib et al. , Smail et al. c). For
spectroscopic arcs, Abell is one of the best clusters for our purposes since Pello et al. ( )
have secured accurate redshifts for two of the giant arcs in the cluster core. Similarly, the presence
of at least multiply-imaged sources in Abell , identi(cid:12)ed using HST, means that we can
strongly constrain not only the absolute mass in the cluster core, but also the detailed form of
its distribution. We can thus expect that remaining uncertainties in the mass distribution will
predominantly arise from unresolved granularity on scales (cid:20) kpc not attached to any galaxy.
The fact that we can make such a statement attests to the detailed view of the cluster mass
provided by lensing.
... Sample selection contaminations
The (cid:12)nal uncertainty we must consider arises from contamination of the arclet catalog by
foreground galaxies and, in particular, cluster members. Indeed, considering only the number
of galaxies detected within the WFPC (cid:12)eld and comparing this to deep (cid:12)eld counts (Smail et
al. d), (cid:24) galaxies per magnitude are cluster members down to R(cid:24). (Fig. ). This
contamination is stronger in the center of the cluster than in the outer parts as the surface density
{ {
of galaxies within a cluster falls faster than =r.
If the contaminating galaxies are randomly orientated, the shear will be reduced below the
true value and hence we obtain an arti(cid:12)cially reduction of the mean redshift of the background
population. In the absence of a reliable distance separation on the basis of arclet colors, we have
developed a statistical method to estimate the unlensed contamination.
Lensing displaces the ellipticity distribution of faint sources from that observed for blank
(cid:12)elds (or unlensed sources) which should be centred on the null ellipticity; this is illustrated
schematically in Fig. . As discussed in x., redshifts can only be estimated for images with
orientations within deg of the predicted shear direction (i.e. arclets with (cid:28)I x > ). Images with
(cid:28)I x < must either be cluster members, foreground galaxies or, conceivably, lensed background
ob jects which are insu(cid:14)ciently deformed to move them into the (cid:28)I x > region. The number
of images with (cid:28)I x < therefore provides an upper limit on contamination by unlensed galaxies
and an improved estimate can be determined by considering those lying within the bulk of the
ellipticity distribution ((cid:28)I < (cid:27)(cid:28) and (cid:28)I x < ). By applying a (cid:25)= rotation prior to inversion, we
can also obtain an estimate of the contamination as a function of redshift and directly subtract
this spurious N (z ) from that derived for the total distribution. Although this only provides a
statistical correction for contamination, it gives a good indication of the stability of the derived
N (z ).
. Determining the Field Redshift Distribution to R '.
We now use the photometric catalog of faint arclet candidates discussed in x, together with
the lensing inversion method presented in x, to derive the likely redshift distribution N (z ) of
faint background galaxies viewed through the center of Abell .
In order to quantify the errors in our determination of N (z ) we must (cid:12)rst estimate
uncertainties in the shape measurements of individual arclets as a function of their size and
apparent magnitude. This is important in determining the useful limit of our HST image for
{ {
accurate inversion. To do this we simulated (cid:24) images of di(cid:11)erent known sizes and
ellipticities and then reproduced the detection characteristics applicable to our HST image. By
comparing the actual galaxy catalog to its simulated equivalent, the dispersion in the realised
shape and orientation can be examined as a function of apparent magnitude (Fig. ). The
formalism of x then gives, for each image, the likely redshift error arising from these observational
uncertainties.
The simulations are very helpful in revealing two important limitations that apply in deriving
redshift distributions from lensing data:
Firstly, we (cid:12)nd the redshift error does not track the measured apparent magnitude very well
for a realistic distribution of image properties, but depends more closely on the intrinsic shape
and signal/noise of each image. Clearly the most interesting region for consideration is that which
lies beyond the current spectroscopic limit, viz. R >. In the context of our relatively short HST
exposure of Abell , Fig. d shows that the uncertainty in the measurement of ellipticities
increases signi(cid:12)cant beyond R= (but strongly depend on the size and the ellipticity of the
ob jects) and thus inversion becomes highly uncertain. Although we can correct for this e(cid:11)ect
statistical ly (see x.., Fig. ), clearly the uncertainty in this correction could swamp the signal
from those sources for which reliable inversion is possible.
Secondly, even with adequate signal/noise for all images, (cid:1)z depends on z itself (Fig. ). A
single cluster lens can thus only usefully constrain the number of sources lying in a speci(cid:12)c redshift
range ( .< z <. for Abell ) although some information is available on the overall N (z ) as
well.The (cid:12)rst limitation is more serious as it emphasises that those samples for which
lensing-induced redshift distributions can be reliably determined are unlikely to be strictly
magnitude-limited as has been the case traditionally for ground-based spectroscopic surveys.
Notwithstanding the contamination from sources which are not amenable to inversion, a
magnitude-limited arclet sample would never produce a magnitude-limited source sample because
of the variable magni(cid:12)cations. However, the fact that little can be said about a subset of faint
{ {
sources is a more serious di(cid:14)culty when comparing with current model predictions which are
largely based on integrated magnitudes. Either it must be assumed that the compact sources are
a representative subset of those for which inversion is practical, or evolutionary models must take
into account the e(cid:11)ect of an areal threshold rather than an integrated magnitude. Conceivably
with much longer integrations, the signal/noise of each faint image will improve su(cid:14)ciently to
reduce the uncertainties.
As the source surface brightness and k-corrections depend strongly on redshift, the visibility
of a faint source is also a complex function of redshift and type. Although this is true of any
isophotally-selected faint galaxy sample and is not further distorted by the lensing process, as
the arclet population presumably probes to much higher redshift than the brighter spectroscopic
samples, the uncertainties in allowing for visibility losses are presumably much greater. A speci(cid:12)c
problem, raised originally by Smail et al. ( ), is the possibility that only dense star-forming
regions have su(cid:14)cient ultraviolet (cid:13)ux and high enough surface brightness to produce arclets visible
with HST.We now illustrate the above e(cid:11)ects in the context of the actual Abell catalogue. To
R ', we have candidate arclets and for each of these, our procedure delivers a likelihood
distribution for the true, unlensed apparent magnitude Rsource and the redshift z . We can apply
the contamination correction discussed in x.. to determine the mean redshift of those sources
with z > zcl and this can be compared with various predictions. This method is illustrated on the
Rsource {z plane in Figure together with the no evolution prediction for an R-limited sample
following the procedure described by Ellis ( ). The latter prediction is based on type-dependent
bJ luminosity functions and morphological proportions observed for the local (cid:12)eld population
(Loveday et al. ) transformed to the R-band using Hubble-sequence colours with k-corrections
taken from King & Ellis ( ). The results are also summarised in Table .
For R < there are too few arclets in our catalogue for meaningful results, but for
< R < the results indicate a gradual increase in mean redshift with apparent magnitude
(Fig. dashed-line). The mean redshift of the arclet population is reasonably close to the no
{ {
evolution expectations to R '. However, upon examination of the individual redshifts, there
appears to be an excess of low redshift arclets whose proportion is independent of magnitude and
whose origin could explain the trend towards low mean redshifts for faint arclets found earlier by
Smail et al. ( b) and Kneib et al. ( a). It is now clear, following the discussion above, that
this e(cid:11)ect arises because a fraction of the images have insu(cid:14)cient shear to be correctly inverted
and the residual uncertainties in the correction illustrated in Fig. can a(cid:11)ect the results at the
level where interesting scienti(cid:12)c conclusions are required. We can quantify this e(cid:11)ect by restricting
the technique to those images whose isophotal areas exceed pixels and ellipticity is larger than
.. As Fig. shows (dotted-dashed line), this leads to an increase in the mean redshift at all
magnitudes and the large ma jority of the `low-z' points disappear.
Out of a total sample of well-de(cid:12)ned arclets to R'., only two are beyond z ' and the
mean redshift at R ' is ' i.e. only slightly above the no evolution prediction. Thus, unless the
smaller arclets represent an entirely di(cid:11)erent population of sources or the intrinsic size of a source
is a strong function of redshift beyond ', the absence of a large number of very distant luminous
sources to R ' is a secure result.
It is important to recognise that, for a lensing cluster at z= ., Fig. b shows the mean
error in inverted redshift is high even for a well-de(cid:12)ned arclet. Typically, for the arclets amenable
to individual inversion, (cid:1)(cid:15)=(cid:15) ' . - . and thus, from Fig. b, a source at z ' could be placed
anywhere from < z <. A large sample size obviates the need for more precise inversion but we
also note that the redshift distribution would be veri(cid:12)ed via inversion through a well-constrained
cluster at higher redshift. An arclet sample viewed through a cluster at z ' . would reduce the
implied redshift error for a source at z ' by a factor of .
It may also be possible to verify the brighter arclets spectroscopically and this will clearly
lead to further improvements. Such con(cid:12)rmation can be made more e(cid:11)ective by selecting the
bluest cases with predicted redshifts z (cid:20) : where strong emission lines would lie in the range of
optical and near-infrared spectrographs. A subset of well-distributed arclets would represent a
valid test of our inversion since this depends on geometrical quantities and the cluster mass model,
{ {
both of which are independent of the photometric properties of the background sources. As an
encouragement for interested workers, we therefore list in Table the inverted redshifts for each of
the ma jor arcs and multiple images discussed in x. and those remaining arcs for which reliable
inversion is possible plotted on Fig. .
Finally, we turn to the e(cid:11)ect on the inversion of our improved HST-based mass model
compared to the earlier ground-based equivalent of Kneib et al. ( ). In Table we show the
mean inverted redshift after statistical correction for contamination for the ground-based (`old')
and HST-based (`new`) mass models. In both cases, the di(cid:11)erences are very minor and illustrate
that the principal uncertainty in inversion through Abell is no longer the global mass model
for the cluster.
Although inversion through well-constrained lenses is an extremely promising prospect, this
pilot study has shown that the ma jor observational limitation is the signal/noise of the required
shape parameters for the faint sources. This demonstrates the importance of securing deeper HST
exposures. A second revelation is the importance of developing a new approach in the construction
of model predictions. It seems unlikely that such faint sources can easily be constructed into
apparent magnitude-limited samples and thus much work is needed to produce surface brightness
limited predictions. Notwithstanding these di(cid:14)culties, Abell remains an exceptionally
promising cosmic lens and the opportunities for verifying or otherwise the predicted redshifts for
the brighter arclets are excellent. Such data will improve the mass model and lead to even tighter
constraints on the redshifts of sources beyond reach of ground-based spectrographs.
. Conclusions
. This paper is the (cid:12)rst attempt to constrain the redshift distribution of very faint galaxies by
using the HST to study images of sources which have been lensed by a massive foreground
cluster. Using multiple images and newly-discovered lensing features in the rich cluster
Abell , we have constructed a precise mass distribution which is more tightly constrained
{ {
that for any other cluster. With HST high-resolution data we have now su(cid:14)cient information
to constrain not only the mass pro(cid:12)le of the cluster but also to give limits on the masses of
individual cluster galaxies.
. Using this mass model and a new formalism we develop based on the observed image
parameters of faint sources, we demonstrate how it is possible to deduce the redshift
distribution of very faint galaxies viewed through the cluster as well as to account
statistically for contamination by unlensed sources and in understanding the various
uncertainties. First results are presented for a large sample of arclets to R '-., (cid:24)
magnitudes fainter than results from ground-based spectroscopy.
. Several limitations arise in the interpretation of our inverted redshifts when comparison is
made with evolutionary models. These limitations may help to explain earlier results which
have tended to yield arclet redshifts somewhat less than extrapolation of ground-based
spectroscopic data would imply. Even with the high resolution of the HST, it is di(cid:14)cult in
short integrations to measure faint galaxy shapes adequately to invert a magnitude-limited
It therefore appears more practical to invert area-limited samples and we
sample.
demonstrate that more reliable results are obtained using such a subset.
. We demonstrate that, notwithstanding the uncertainties and sample selection criteria we
have adopted, the absence of a large number of very distant sources (z > ) in our inverted
redshift distributions is a robust result. At R ', the mean redshift for samples corrected
for contamination or those based on individual arclets of high signal/noise is only ' .-..
. The brighter arclets, whose redshifts are estimated via our technique, are amenable to direct
spectroscopic examination. Such con(cid:12)rmation can be made more e(cid:11)ective by selecting the
bluest cases with predicted redshifts z (cid:20) : where strong emission lines would lie in the range
of optical and near-infrared spectrographs. Note that con(cid:12)rmation of a carefully-selected
subset of well-distributed arclets would still represent a valid test of our inversion technique.
The lensing inversion depends only on geometrical quantities and the cluster mass model,
both of which are independent of the photometric properties of the background sources.
{ {
We thank Bob Williams and Ray Lucas and other STScI sta(cid:11) for their assistance with the
rapid processing of this data. We acknowledge enthusiastic support from a number of colleagues
including Roger Blandford, Bernard Fort, Yannick Mellier, Jerry Ostriker, Roser Pell(cid:19)o, Peter
Schneider and Martin Rees. This paper is based on observations with the NASA/ESA Hubble
Space Telescope, obtained at the Space Telescope Science Institute, which is operated by the
Association of Universities for Research in Astronomy Inc. JPK gratefully acknowledges support
from an EC Fellowship and IRS from NATO and Carnegie Fellowships. WJC acknowledges
support from the Australian Department of Industry, Science and Technology, the Australian
Research Council and Sun Microsystems.
{ {
B-r(cid:3)
Multiple images
R (F W)
(cid:22)mean
z
.+ :(cid:0) :
.
/
. / .
. / .
./././. ././. /.
///
./. /. /.
.
.
.
.
.
.
. /./ . (. )
. /. /.
.(cid:6) :
. (cid:6) :
(cid:24).
H--
(cid:24).
.(cid:6) :
(cid:24).
H-
(cid:24).
.(cid:6) :
./.
. /
. / .
/ H
(cid:3) from Leborgne et al. ( ) when available, typical errors for red ob jects can be as high as .
mag (Kneib et al. ).
Table : List of con(cid:12)rmed and candidate multiple images from our HST study, along with
their ground-based colors and measured or predicted redshifts.
{ {
z+ Comments
zphoto
z-
zopt
(cid:22)R
Rcor
# id
(cid:3)
.- . disk galaxy
.
.
(cid:3)
.
.- .
.
.
compact + disk galaxies
.
.
. disk galaxy
. -.
.
.
.
.
.
.- .
.
.
.
.
.
.
.
(cid:3)
.-. compact galaxies
.
.
.-.
.
.
(cid:3)
.
.
.
.-.
.
.
(cid:3)
.-. extended galaxies
.
.
.
.
.
.
.-.
.
.
.
.
(cid:3)
.
.
.
.-.
.
.
.
.-.
.
.
.
.
.
.-.
.
.
.
.
.
.-.
.
.
.
.
.- .
.
.
.
.
.
.
.
.
.-.
.
.
.-.
.
.
.
counter-image of
.
.
.
.
.
. -.
.
.
.-.
.
.
.
.
.
.
.
.
.
.
.
.
.
cusp arc, disk of ?
H,, (cid:24). (cid:24).
.
.
.
fold arc
H,
(cid:24). (cid:24).
. counter-image of
.
.
. counter-image of
.
.
. counter-image of
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
Table : The catalog of candidate arclets from our HST/WFPC image. Ob jects with zphoto
represent candidate arclets from Pello et al. ( ) which lie within the HST/WFPC (cid:12)eld.
Ob jects denoted (cid:3) are either misidenti(cid:12)cations (edge-on galaxy or a close pair) or ob jects
suspected not to be lensed when comparison is made with the shear orientation.
{ {
(cid:18)
Cluster-Size
xc
yc
a/b
rc
(cid:27)
rcut
degree
(kpc)
Component
(arcsec)
(arcsec)
(km/s)
(kpc)
- .
#
.
.
.
.
.
.
.
#
-.
-.
.
.
.
.
.
#
-.
- .
.
.
.
.
.
.
.
.
#
.
- .
.
M/LV
rcut
rc
Galaxy-Size
Mtot
(cid:27)
M(cid:12) M(cid:12) /L(cid:12)
(kpc)
(km/s)
(kpc)
Component
for M(cid:3)V =-.
.
.
.
.
.
Table : Characteristics of the various components in the improved mass model of Abell .
Positions and orientations are de(cid:12)ned on the WFPC- image (Fig. a) with the position angle
(cid:18) increasing anti-clockwise from the X-axis.
arclet Ncorrarclet < z >old < z >new < z >nocorr
< z >N E N total
R
new
-
.
.
.
.
-
.
.
.
.
-
.
.
.
.
Table :
Inversion results using all arclets: < z >N E represents the mean expected for
arclet is the total number in each subsample, Ncorrarclet is that
no luminosity evolution. N total
after correction for contamination by unlensed images. < z >old and < z >new represent,
respectively, the mean arclet redshift using the ground-based (Kneib et al. ) and HST-
based mass models. For the HST model, < z >nocorr
indicates the mean redshift prior to
new
contamination correction.
{ {
REFERENCES
Bertin, E., Sextractor Manual .
Bertin, E., Arnouts, S., , preprint
Birkinshaw, M., Hughes, J.P., , ApJ, , .
Bonnet, H., Mellier, Y., , A&A, in press.
Brainerd, T., Blandford, R., Smail, I., , preprint SISSA/astro-ph .
Ebbels, T., et al. , in preparation.
Ellis, R.S., , in Unsolved Problems in Astrophysics, eds Bahcall, J. & Ostriker, J., Princeton,
in press.
Fort, B., Mellier, Y., , Astron. Astr. Rev., , .
Jones, M., Saunders, R., Alexander, P., Birkinshaw, M., Dilon, N., Grainge, K., Hancock, S.,
Lasenby, A., Lefevre, D., Pooley, G., , Nature, , .
Gri(cid:14)ths, R.E., Casertano, S., Ratnatunga, K.U., Neuchaefer,L.W., Ellis, R.S., Gilmore, G.,
Glazebrook, K., Santiago, B., Elson, R.A.W., Huchra, J.P., Windhorst, R.A., Pascarelle,
S.M., Green, R.F., Illingworth, G.D., Koo, D.C. & Tyson, A.J., , ApJ, , L
Holtzman, J.A., Burrows, C.J. , Casterno, S., Hester, J.J., Trauger, J.T., Watson, A.M., Worthey,
G., , preprint.
Kassiola, A., Kovner, I., , ApJ, , .
Kaiser, N., Squires, G., Broadhurst, T. , ApJ, in press. , preprint.
King, C.R. & Ellis, R.S., ApJ, , .
Kneib, J.-P., , PhD. Thesis, Universit(cid:19)e Paul Sabatier, Toulouse, France
Kneib, J.-P., Mellier, Y., Fort, B., Mathez, G., , A&A, , .
{ {
Kneib, J.-P., Mellier, Y., Fort, B., Soucail, G., Longaretti, P.Y., a, A&A, , .
Kneib, J.-P., Melnick, J., Gopal-Krishna, b, A&A, , L.
Kneib, J.-P., Mellier, Y., Pell(cid:19)o, R., Miralda-Escud(cid:19)e, J., Le Borgne, J.-F., B(cid:127)ohringer, H., Picat,
J.-P., , A&A,
in press.
Kochanek, C.S., , MNRAS, , .
Le Borgne, J.-F., Pell(cid:19)o, R. & Sanahuja, B., , A&AS, , .
Loveday, J., Peterson, B. A., Efstathiou, G., Maddox, S.J. , ApJ, , L.
Mellier, Y., Fort, B., Kneib, J.-P., , ApJ, , .
Pell(cid:19)o, R., Soucail, G., Sanahuja, B., Mathez, G., Ojero, E., , A&A, , L.
Pell(cid:19)o, R., Le Borgne, J.-F., Sanahuja, B., Mathez, G., Fort, B., , A&A, , .
Smail, I., Ellis, R.S., Fitchett, M. J., Norgaard-Nielsen, H.U., Hansen, L., Jorgensen, H.E. ,
MNRAS, , .
Smail, I., Couch, W.J., Ellis, R.S., Sharples, R.M., a, ApJ, , .
Smail, I., Dressler, A., Kneib, J.-P., Ellis, R.S., Couch, W.J., Sharples, R.M., Oemler, A., b,
ApJ,
in press.
Smail, I., Hogg, D.W., Blandford, R., Cohen, J.G., Edge, A.C., Djorgovski, S.G. c, MNRAS,
in press.
Smail, I., Hogg, D.W., Yan, L., Cohen, J., d, ApJ, , .
Wilson, G., Cole, S. & Frenk, C.S., , MNRAS, submitted
This manuscript was prepared with the AAS LATEX macros v. .
{ {
Fig. . (a) The full (cid:12)eld of our F W WFPC- exposure of Abell (z = :). (b)
Central portion showing that several multiply-imaged sources, numbered according to the
scheme of Le Borgne et al. ( ), are con(cid:12)rmed by virtue of their mirrored morphological
features (see text for details).
Fig. . Di(cid:11)erential galaxy counts within the Abell WFC image (. arcmin). The
dashed line de(cid:12)nes our estimated completeness limit at the % level. The dotted line
indicates (cid:12)eld counts in R from Smail et al. ( d). The dotted-dashed line is the cluster
galaxies counts estimated by subtracting the (cid:12)eld counts from the observed counts.
Fig. . The distribution of (cid:24) arclet candidates with < R < selected from the
HST image (thin lines) compared to those in the ground-based analysis of Pell(cid:19)o et al. ( )
(thick lines). The shear (cid:12)eld based on the HST sample illustrates the need for further mass
components associated with the brighter cluster galaxies.
Fig. . Shear map for the cluster center derived from the orientations and ellipticities of
the HST arclets. The most signi(cid:12)cant mass components are indicated. The new mass model
extends that of Kneib et al. ( ) by including ma jor mass components associated with
galaxies # and # (see Table ) as well as smaller halos around luminous cluster
members (see text for details). At the cluster redshift arcsec is equivalent to . kpc.
Fig. . Contour map for the adopted mass distribution and the shear map implied for a
source plane at zS = . Countours correspond from the lowest to the highest to a density
of ., ., ., ., ., ., ., ., . and . M(cid:12) kpc(cid:0) . At the cluster redshift arcsec is
equivalent to . kpc.
Fig. . Lens deformation diagram (see text for de(cid:12)nition of quantities).
{ {
Fig. . Redshift probability distribution for (cid:28)I x > (full and dashed line) and (cid:28)I x <
(dotted line) { see text for details.
Fig. . (a) D parameter versus source redshift for cluster lenses at z = : and z = :.
The solid line is for (cid:10) = and the dashed line for (cid:10) = :. (b), D=[(z (cid:0) zL )D ] versus
source redshift illustrating the higher the source redshift, the greater the uncertainty in the
inverted redshift. Note that the redshift of a distant source is more accurately derived using
a high redshift lens.
Fig. . Ellipticity distribution showing that observed is the sum of the lensed and unlensed
galaxies vectors.
Fig. . Simulated errors for image parameters relevant to lensing inversion as function
of apparent magnitude: (a) dispersion on the ellipticity (cid:27)" , (b) dispersion on the orientation
(cid:27)(cid:18) , (c) relative error (cid:27)"="mes and (d) relative error ("true (cid:0) "mes )="mes . Each data point
was determined from realisations of the same source., a dot denotes galaxies with small
ellipticities ((cid:15) < :), (*) denotes galaxies whose isophotal area < pixels, and (+) denotes
galaxies with intrinsic large ellipticities ((cid:15) > :) and large isophotal area > pixels.
Fig. . Mean redshift vs. intrinsic magnitude for various arclet samples. The solid line
represents the no-evolution prediction according to assumptions detailed in the text. The
dashed line represents the results for all arclet candidates after making a statistical correction
for foreground and cluster contamination. The dotted-dashed line is the same sample after
excluding images whose isophotal areas are smaller than pixels. Squares represent the
individual inverted redshifts of all arclets greater than pixels in area. Solid symbols denote
those with (cid:15) > : and open symbols those with (cid:15) < :.
{ {
Figure a
{ {
Figure b - Plate
{ {
Figure
{ {
Figure
{ {
Figure
{ {
Figure
{ {
Figure
pot
τ
I
τ
Sτ
2θ s
2θ
I
corrective term
{ {
Figure
{ {
Figure a
Figure b
{ {
Iyτ
Figure
"unlensed"
"lensed"
Ixτ
{ {
Figure
{ {
Figure
|
astro-ph/9507096 | 1 | 9507 | 1995-07-24T18:56:47 | The theoretical calculation of the Rossby number and the `non-local' convective overturn time for pre-main sequence and early post-main sequence stars | [
"astro-ph"
] | This paper provides estimates of convective turnover time scales for Sun-like stars in the pre-main sequence and early post-main sequence phases of evolution, based on up-to-date physical input for the stellar models. In this first study, all models have solar abundances, which is typical of the stars in the Galactic disk where most of the available data have been collected. A new feature of these models is the inclusion of rotation in the evolutionary sequences, thus making it possible to derive theoretically the Rossby number for each star along its evolutionary track, based on its calculated rotation rate and its local convective turnover time near the base of the convection zone. Global turnover times are also calculated for the complete convection zone. This information should make possible a new class of observational tests of stellar theory which were previously impossible with semi-empirical models, particularly in the study of stellar activity and in research related to angular momentum transfer in stellar interiors during the course of stellar evolution. | astro-ph | astro-ph | Ap. J. ms. number 32821, accepted
The theoretical calculation of the Rossby number
and the ‘non-local’ convective overturn time
for pre-main sequence and early post-main sequence stars
Yong -Cheol Kim and Pierre Demarque
Department of Astronomy, and Center for Solar and Space Research
Yale University, Box 208101, New Haven, CT 06520-8101
E-Mail: [email protected], [email protected]
ABSTRACT
This paper provides estimates of convective turnover time scales for Sun-like stars
in the pre-main sequence and early post-main sequence phases of evolution, based on
up-to-date physical input for the stellar models. In this first study, all models have
solar abundances, which is typical of the stars in the Galactic disk where most of the
available data have been collected. A new feature of these models is the inclusion of
rotation in the evolutionary sequences, thus making it possible to derive theoretically
the Rossby number for each star along its evolutionary track, based on its calculated
rotation rate and its local convective turnover time near the base of the convection
zone. Global turnover times are also calculated for the complete convection zone. This
information should make possible a new class of observational tests of stellar theory
which were previously impossible with semi-empirical models, particularly in the study
of stellar activity and in research related to angular momentum transfer in stellar
interiors during the course of stellar evolution.
Subject headings: general – stars: interiors – stars: evolution
5
9
9
1
l
u
J
4
2
1
v
6
9
0
7
0
5
9
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
– 2 –
1.
Introduction
There is a rapidly growing body of observations relating to the study of solar-stellar
phenomena which will lead to a better understanding of the internal dynamics of the pre-main
sequence and early post-main sequence evolutionary phases of stars like the Sun. But before this
wealth of data can be fully understood, the role of the convection zone structure, its depth and
overturn time scale, and its interaction with rotation need to be clarified. This is of importance
for understanding not only the mechanism of angular momentum transfer in stars, but also the
evolution of rotating stars, both in the pre main-sequence and post main-sequence phases. The
interaction of rotation with convection is now widely believed to be responsible for the generation
of stellar dynamos and the observed stellar magnetic activity and activity cycles.
Durney and Latour (1978) (see also Durney, Mihalas and Robinson 1981, and Durney and
Robinson 1982) made an important step forward in relating the principles of mean-field dynamo
theory to the observations. They showed that if a stellar dynamo is responsible for the observed
stellar activity, there should be a relation between magnetic activity and the characteristics of
rotating stellar convection zones. With the help of dimensional arguments, they pointed out
the significance of the Rossby number (proportional to the ratio of the rotation period to the
convective turnover time) in the dynamo mechanism. Since in the dynamo model, the dynamo
action is believed to take place at the base of the convection zone, anchored in the radiative
layers just below the convective interface, the convective turnover time of the deepest part of the
convection zone is the most relevant in the evaluation of the Rossby number. Soon afterwards, the
availability of precise rotation periods for magnetically active stars (Baliunas et al. 1983) made it
possible to test this hypothesis in a semi-empirical way by combining convective turnover times
derived from model convection zones, such as the models of Gilman (1980), with the observed
rotation periods (Noyes 1983). This was done by a number of researchers (Mangeney and Praderie
1984; Hartmann et al. 1984; Noyes et al. 1984) for stars near the main sequence and for pre-main
sequence stars (Simon et al. 1985). Since then a large number of studies have been performed
correlating different types of magnetic activity indices to such semi-empirically derived Rossby
numbers and to other parameters such as the stellar rotation rate (Basri 1987; Shrijver and Rutten
1987; Simon and Fekel 1987; Dobson and Radick 1989).
The pattern of convective velocities as a function of effective temperature and age derived
from stellar models, which serve as input in Rossby number calculations, depend sensitively on the
particulars of the stellar interior models, either input parameters such as chemical composition, or
mass, or physics input such as opacities or the equation of state used. In addition, because of the
well-known non-linearity of the equation of stellar structure, it is unadvisable to construct stellar
envelope models by simple inward integration without applying the interior boundary conditions,
as the early calculations frequently did; fully consistent interior models are needed. An important
step in relating activity observations to self consistent stellar evolutionary tracks and the predicted
evolutionary changes in convective overturn times was made by Gilliland (1985, 1986). Other
calculations of Rossby numbers, based on complete main sequence stellar models, have also been
– 3 –
published by Rucinski and VandenBerg (1986, 1990). Since then, rapid progress has been made in
our knowledge of stellar opacities and equation of state (see e.g. Rogers and Iglesias 1994), and
much improved models of the Sun and Sun-like stars can be constructed (Guenther et al 1992;
Chaboyer et al. 1995; Guenther, Kim and Demarque 1995).
The theory of rotating stellar evolution has also advanced. The work of Endal and Sofia (1978,
1981), which included the spin-down due to a stellar wind, and introduced into stellar evolution
the effects of various rotationally induced mixing processes acting on different time scales, thus
relating angular moment transfer to internal mixing, opened up new ways of confronting theory
and observation (Pinsonneault et al. 1989, 1990; Chaboyer et al. 1995).
The purpose of this paper is to provide estimates of turnover time scales for Sun-like stars
in the pre-main-sequence and early post main-sequence phases of evolution, based on up-to-date
physical input for the stellar models. In this first study, all models have solar abundances, which
are typical stars in the Galactic disk. Another new feature of these models is the inclusion of
rotation. Because the evolution of internal rotation has been included in the models, it is possible
to derive theoretically the Rossby number for each star along its evolutionary track based on the
theoretical estimates for both the convective turnover time and the rotation rate of the convection
zone. These internally self consistent models should make possible a new class of observational
tests of stellar theory which were impossible with semi-empirical models.
We describe convection by the mixing-length formalism in the usual way. While it is known
that the convective velocities near the stellar surface are not well described by the mixing-length
approximation (Kim et al. 1995a,b), the convective turnover time scales calculated here are
dominated by the conditions near the base of the convection zone, where the temperature gradient
is for all practical purposes adiabatic, and the mixing-length approximation is known to provide
an adequate description of convection, at least in an average sense (Chan and Sofia 1989; Lydon
et al. 1992). For this reason, the Rossby number estimates should be little affected (sub ject to
a constant scale factor) by improvements in our understanding of convection. We emphasize,
however, that for many other purposes, such as describing the outer layers where radiation plays
a dominant role, the mixing length approximation is inadequate, and more refined convection
models that take into consideration the interplay between convection and radiation, are needed
(Kim et al. 1995a,b). This conclusion applies in particular for understanding the structure of the
transition superadiabatic layer at the top of the convection zone. It is also likely to apply for
understanding the behavior of magnetic fields, the details of the generation of acoustic noise in
stellar chromospheres, and the driving of p-modes in Sun-like stars.
Section 2 describes the series of stellar models with masses ranging from 0.5M⊙ to 1.2M⊙
which were evolved from the fully convective pre-main-sequence Hayashi phase to the sub-giant
phase. The calculation of the convective turnover time and of the Rossby number and their
evolution as a function of time are considered in Sections 3 and 4, respectively. Finally, we briefly
discuss the results in section 5.
– 4 –
2. Calculations
2.1. Stellar models
A series of stellar models with masses ranging from 0.5 to 1.2M⊙ (in 0.1M⊙ increments),
have been evolved from a fully convective pre-main sequence model to the sub-giant phase. The
OPAL opacities tables (Iglesias and Rogers 1991), constructed for the solar mixture of Anders and
Grevesse(1989) were used, together with the Kurucz (1991) low temperature opacities. The Kurucz
(1992) model atmospheres served as surface boundary conditions. The numerical tolerances and
input physics were identical for all evolutionary runs, and similar to those adopted by Chaboyer et
al. (1995). All models used the parameters derived for the standard solar model, where the initial
X , Z , and the mixing length ratio α are varied until a solar model at the solar age of 4.55Gyr
(Guenther 1989) has the observed solar values of luminosity, radius, and Z/X . In addition, the
solar surface rotational velocity and 7Li depletion were used to calibrate the rotation and diffusion
parameters of all evolutionary sequences, as described in sections 2.2 and 2.3 below. The solar
model in this calibration matches the solar radius and luminosity to within 0.01%, while the
surface Z/X matches the observed value to within 1.0%. The model also reproduced the observed
solar rotation rate and Li depletion to within 1.5%. Table 1 summarizes the characteristics of the
models and their input parameters.
Figure 1 shows the evolutionary tracks in the H-R diagram. For the internal rotation rates
considered here, rotation has a negligible effect on both the rate of evolution and the path of the
evolutionary track in the H-R diagram (Pinsonneault et al. 1989; Deliyannis et al. 1989).
2.2. Rotation
All models used in this paper have been constructed using a version of the Yale Rotating
Stellar Evolution Code (Prather 1976, Pinsonneault 1988). Recently, the YREC has been improved
in the microscopic diffusion and its interaction with rotational mixing (Chaboyer et al. 1995). The
calculation has been carried out using this improved version.
The evolutionary sequences were started from fully convective pre-main sequence models in
the Hayashi phase. At first, the whole star rotates as a rigid body, and spins up as it contracts.
The torque due to the stellar wind then takes over and spins the star down (see below), and in the
process of transferring out internal angular momentum, progressively depletes 7Li in the star. The
evolution of the internal angular momentum distribution follows the approach of Pinsonneault et
al (1989), which includes the effects of rotationally induced instabilities in the radiative layers.
This results in a state of differential rotation. The Sun seems to be overdepleted in 7Li compared
with other stars of its age and spectral class. Since the amount of mixing in a star increases as its
initial rotational velocity increases, it is likely that the Sun was initially a rapid rotator. Therefore,
even though observations show that the ma jority of T-Tauri stars have rotation velocities around
– 5 –
10km/s, we have chosen an initial rotation velocity of 30km/s for the solar model. Observations
of T-Tauri stars also indicate that there is no large difference in surface rotation rates between
high and low mass stars over the mass range we study, i.e. 0.5 ∼ 1.2M⊙ (Bouvier 1991; Bouvier
et al. 1993), and for this reason, we have the applied the same rotation parameters to all masses.
Figure 2 shows the evolution of the rotation period of our theoretical models. It is important to
note that changing the initial rotation velocity affects the time-scale for early spin-down, but does
not change appreciably the final configuration. This is due to the fact that our calibration requires
the solar model to rotate at the present rotation rate of the Sun (the adopted initial rotational
velocity affects primarily only the present 7Li abundance, which is not particularly relevant to the
convective turnover calculations of this paper). Note in Figure 2 that the rotation period depends
sensitively on mass for a given age.
When calculating the evolution of the Rossby number, the adopted wind law is the most
important input of our rotating models, as internal structural effects of rotation are minimal in
these models which rotate relatively slowly. A modified version of Kawaler’s parameterization
for the loss of angular momentum due to magnetic stellar wind (Kawaler 1988) has been used
(Chaboyer et al. 1995). It is given by:
dJ
dt
= fkKwR2−N M −N/3 M 1−2N/3 ω1+4N/3
(ω < ωcrit),
= fkKwR2−N M −N/3 M 1−2N/3 ωω4N/3
crit
dJ
dt
where R is the radius in units of the solar radius (R⊙ ), M is the mass in units of the solar
mass (M⊙ ), ωcrit introduces a saturation level into the angular momentum loss law (set to
1.5 × 10−5 s−1 ), Kw = 2.036 × 1033 (cid:0)1.452 × 109 (cid:1)N in cgs units, and M is the mass-loss rate in unit
of 10−14M⊙yr−1 (set to 2.0). Here, it is primarily the exponent N in the wind model (a measure
of the magnetic field geometry), which determines the rate of angular momentum loss with time.
We have adopted the value of 1.5 for N , which reduces to the empirical Skumanich (1972) law
near the main sequence i.e.,
(ω ≥ ωcrit),
vrot ∝ τ −.51 ,
where vrot is equatorial rotation velocity, and τ is the age in Gyr. The constant factor in the wind
model, fk , determines the total amount of the angular momentum loss. We adjust fk for a given
N to give the solar surface rotation velocity at the solar age. We use the observed value at about
30 degrees from the equator, 1.86km/s, since this value is close to the mean value which from the
seismology data, the interior of the Sun appears to approach (Libbrecht and Morrow 1991)
For the sake of completeness, the transport of angular momentum and chemical elements have
also been taken into account in the models. Two types of rotation-induced mixing – the dynamical
shear instabilities and the Solberg-Hoiland instability – and three type of secular instabilities –
meridional circulation, the Goldreich-Schubert-Fricke instability, and the secular shear instability
– were included in the calculations (Chaboyer et al. 1995, Pinsonneault et al. 1989).
– 6 –
In this study, the uncertain effects of other secular instabilities are treated as free parameter,
and the diffusion coefficients are set to be the same as that of meridional circulation, fc . The
value of fc is fixed in the solar model by requiring the model lithium depletion to match the solar
value at the solar age. The depletion of Li is inferred by comparing the cosmic abundance with
the photospheric abundance. A comparison of the meteoric abundance with the photospheric
abundance shows the depletion to be a factor of 140+40
−30 (Anders and Grevesse 1989). We have
therefore set the solar 7Li depletion factor equal to 140.
2.3. Diffusion
The microscopic diffusion coefficients of Michaud and Proffitt (1993) have been used. They
have the advantage of being valid not only for 4H e and 1H , but also for 3H e, 6Li, 7Li, and
9B e. Comparison with the diffusion coefficients of Thoul et al. (1994) indicates that the
Michaud-Proffitt coefficients are good to within 15%. Finally we note that when diffusion is taken
into account, the surface Z/X is not a constant during a stellar evolution calculation. As 4H e
diffuse with respect to hydrogen with the relatively short time scale, the model structures are
affected. Measurements of the solar photosphere do not actually determine Z – they measure Z/X
(i.e.
[F e/H ]). The Anders and Grevesse (1989) photospheric mixture with meteoritic Fe gives
Z/X = 0.0267 ± 0.001. Thus, our model of the present Sun was constrained to match this number.
3. Global turnover time and Rossby number
The close connection between stellar rotation and its chromospheric emission can be described
in terms of general stellar dynamo models. In the mean-field dynamo theory, a dimensionless
parameter, the dynamo number, characterizes the model behavior. The dynamo number is
essentially proportional to the inverse square of the Rossby number, NR , which is the ratio of the
stellar rotation period to the local convective turnover time (Durney and Latour 1978; Noyes et
al 1984). Thus, in principle, one could draw a theoretical Rossby number vs. magnetic activity
diagram.
In practice, however, our knowledge on stellar convection is too limited to calculate ‘correct’
convective turnover times. The characteristic length scales as well as the velocities are not well
known. Even when one decides to resort to the mixing length approximation, there are still
uncertainties: the mixing length ratio α is assumed to be the same for all stars with different
masses and/or at different evolutionary stage, which is probably not be quite correct. In addition,
some assumption must be made as to where in the convection zone the dynamo process is
operating, since the convective overturn time is strongly depth dependent. For example, Gilman
(1980) set the characteristic convective overturn time equal to the convective overturn time
one scale height above the bottom of the convective zone. On the other hand, Gilliland (1986)
– 7 –
determined the turnover time at a distance of half of the mixing length above the base of the
convection zone.
3.1. Global convective turnover time
To depict the characteristics of convection at each stellar evolution stage, two parameters
have been calculated; the ‘global’ convective turnover time, and the Rossby number. For the
characteristic time scale of convective overturn, the ‘non-local’ (or ‘global’) convective turnover
time has been calculated at each time step. It is:
τc = Z R∗
Rb
where Rb is the location of of the bottom of the surface convection zone, which is defined where
the ∇ − ∇ad = 0, R∗ is the total radius of the stellar model, and v is the local convective velocity.
Figure 3 shows the evolution with time of the convective turnover time. Note that in the pre-main
sequence phase, τc varies rapidly with time. Near the main sequence, τc remains nearly constant
and is primarily a function of mass.
dr
v
3.2. Rossby number
For the Rossby number calculation, the characteristic convective overturn time was set equal
to the ‘local’ convective overturn time at a distance of half of the mixing length α HP
above the
2
base of the convection zone. The ratio of the rotation period to this characteristic convective
overturn time is used to characterize the Rossby number in the deep convection zone, where
dynamo generation of magnetic fields is thought to occur.
NR = 2πv/αΩHP
where v is the characteristic convective velocity, α is the mixing length ratio, Ω is rotational
velocity, and HP is the local pressure scale height. It turns out that the evolution of the
‘local’ turnover time is the same as the ‘non-local’ one, except for a scaling factor.
The evolution of the inverse squared Rossby number is illustrated in Figure 4. This quantity,
sometimes called the ‘dynamo number’, is believed to be proportional to the strength of magnetic
activity. Note that the dynamo number depends on both the age and the mass of the star.
4.
Isochrones
Theoretical isochrones offer the opportunity to test stellar evolution theory in star clusters
where stars are coeval and formed from a gas cloud of uniform composition. Conversely, when
properly calibrated, isochrones can become a powerful tool to study the properties of field stars.
– 8 –
Isochrones were constructed using the evolutionary tracks for the ages of 0.2, 0.5, 0.7, 1.0, 4.55
(the solar age), 10, and 15 Gyr. Their characteristics are listed in Table 2. Figure 5 shows a plot
of isochrones of the non-local turnover time vs. log Tef f (the solid lines). For comparison, a few
isochrones of the local turnover time are shown (the dotted lines), in the same figure. Isochrones
of the non-local turnover time vs. rotation period are given in Figure 6. In Figure 7, rotation
period vs. log Tef f is shown, where the increase of log Tef f can be understood as the increase of
the stellar mass, because of the proximity of the main sequence. The right most point of each line
is for 0.5 M⊙ . Assuming our treatment of stellar rotation is correct, then one can use Figure 7 to
uniquely determine stellar mass and age from the effective temperature and the rotation period.
Figure 8 shows the the inverse square of the Rossby number, N −2
R vs. log Tef f . Once empirical
relations between N −2
R and magnetic activity indices are determined, one can use Figure 8 for
determination of the age and the mass of a star by observing its effective temperature and an
activity index. Figure 9 is the plot of N −2
R vs. rotation period, where for an isochrone, the right
most point of the line represent the lowest mass. We see that, given the assumptions implicit in
our discussion, our grid of theoretical evolutionary tracks provide the means to determine the age
and the mass of a star from a measurement of its rotation period and an activity index.
5. Discussion
Estimates of turnover time scales and the Rossby number are provided, for Sun-like stars
in the pre-main sequence and early post-main sequence phases of evolution, based on up-to-date
physical input for the stellar models, and including rotation.
We expect the results in this paper to be robust, since the convective turnover timescale
is weighted toward the deepest part of the convection zone, where the shortcomings of the
mixing length approximation are least important. This is the reason why our ‘global’ and ‘local’
convective time scale give the same result except for a scaling factor (e.g. Figure 5). This is
consistent with recent numerical simulations of convection (e.g. Chan and Sofia 1989; Kim et al.
1995b) which confirm the validity of the mixing length approximation in the limit of deep and
efficient convection.
In this study, all models have solar abundances; they will therefore find applications in the
interpretation of the rotational history and magnetic activity indices for stars in young star
clusters and Sun-like field stars, which are the most common stars in our part of the Galactic disk.
Caution must be exercised, however, with stars with chemical composition that differ appreciably
from solar. Both the depth of the convection zone and the convective velocities are known to
depend on opacities and equation of state. As more detailed observations about the rotational
properties and magnetic activity of very metal-poor and very metal rich stars become available,
the sensitivity of convective turnover timescales to chemical composition parameters will need to
be explored in some detail.
– 9 –
YCK would like to thank B. Chaboyer and M. Pinsonneault. This work was supported in
part by NASA grants NAG5-1486 and NAGW-2469 to Yale University.
REFERENCES
Anders, E. and Grevesse, N. 1989, Geochim. Cosmochim. Acta, 53, 197
Baliunas, S. L. Vaughan, A. H., Hartmann, L., Middelkoop, F., Mihalas, D., et al. 1983, ApJ, 275,
752
Basri, G. 1987, ApJ, 316, 377
Bouvier, J. 1991, in Angular Momentum Evolution of Young Stars, ed. S. Catalano and J. R.
Stauffer (Dordrecht: Kluwer), 41
Bouvier, J., Cabrit, S., Fernandez, M. E. L. and Matthews, J. M. 1993, A&A, 272, 176
Chaboyer, B., Demarque, P. and Pinsonneault, M. H. 1995, ApJ, 441, 865
Chaboyer, B., Demarque, P., Guenther, D. B. and Pinsonneault, M. H. 1995, ApJ, 446, 445
Chan, K. -L. and Sofia, S. 1989, ApJ, 336, 1022
Deliyannis, C. P., Demarque, P. and Pinsonneault, M. H. 1989, ApJ, 347, L73
Dobson, A. K. and Radick, R. R. 1989, ApJ, 344, 907
Durney, B. R. and Latour, J. 1978, Geophys Ap Fluid Dyn, 9, 241
Durney, B. R., Mihalas, D. and Robinson, R. D. 1981, PASP, 93, 587
Durney, B. R. and Robinson, R. D. 1982, ApJ, 253, 290
Endal, A. S. and Sofia, S. 1978, ApJ, 220, 279
Endal, A. S. and Sofia, S. 1981, ApJ, 243, 625
Gilliland, R. L. 1985, ApJ, 299, 286
Gilliland, R. L. 1986, ApJ300,339
Gilman, P. 1980, in Stellar Turbulence, IAU Colloq. No. 51, ed. D. Gray, J. Linsky (New York:
Springer-Verlag), 19
Green, E. M., Demarque, P. and King, C. R. 1987, Revised Yale Isochrones and Luminosity
Functions, Yale University Observatory, New Haven
Guenther, P. 1989, ApJ, 39, 1156
Guenther, D. B., Demarque, P., Kim, Y. -C. and Pinsonneault, M. H. 1992, ApJ, 387, 372
Guenther, D. B., Kim, Y. -C. and Demarque, P. 1995, in preparation
Hartmann, L. W., Baliunas, S. L., Duncan, D. K. and Noyes, R. W. 1984 ApJ, 279, 778
Iglesias, C. A. and Rogers, F. J. 1991, ApJ, 371, 408
– 10 –
Kawaler, S. D. 1988, ApJ, 333, 236
Kim, Y. -C., Fox, P. A., Sofia, S. and Demarque, P. 1995a ApJ, 442, 422
Kim, Y. -C., Fox, P. A., Demarque, P. and Sofia, S. 1995b ApJ, submitted
Kurucz, R. L. 1991, in Stellar Atmospheres: Beyond Classical Models, ed. L. Crivellari, I. Hubeny
and D. G. Hummer (Dordrecht:Kluwer), 440
Kurucz, R. L. 1992, in The Stellar Population of Galaxies, IAU Symposium No.149, ed. B. Barbuy
and A. Renzini (Dordrecht:Kluwer), 225
Libbrecht, K. G. and Morrow, C. A. 1991, in Solar Interior and Atmosphere, ed. A.N. Cox, W.C.
Livingston and M.S. Matthews (Tucson: Univ. of Arizona Press), 479
Lydon, T. J., Fox, P. A. and Sofia, S. 1992, ApJ, 397, 701
Mangeney, A. and Praderie, F. 1984, A&A, 130, 143
Michaud, G. and Proffitt, C. R. 1993, in Inside the Stars, IAU Colloq. 137, ed. A. Baglin and W.
W. Weiss (San Francisco: ASP), 246
Noyes, R. W. 1983, in Solar and Stellar Magnetic Fields:Origin and Coronal Effects, ed. J. O.
Stenflo (Dordrecht:Reidel), 133
Noyes, R. W., Hartmann, L. W., Baliunas, S. L., Duncan, D. K. and Vaughan, A. H. 1984, ApJ,
279, 763
Pinsonneault, M. H. 1988, Ph.D. dissertation, Yale University
Pinsonneault, M. H., Kawaler, S. D. and Demarque, P. 1990, ApJS, 74, 501
Pinsonneault, M. H., Kawaler, S. D., Sofia, S. and Demarque, P. 1989, ApJ, 338, 424
Prather, M. J. 1976, Ph.D. dissertation, Yale University
Rogers, F. J. and Iglesias, C. A. 1994, Science, 263, 50
Rucinski, S. M. and VandenBerg, D. A. 1986, PASP, 98, 669
Rucinski, S. M. and VandenBerg, D. A. 1990, AJ, 90, 1279
Shrijver, C. J. and Rutten, R. G. M. 1987, A&A, 177, 155
Simon, T. and Fekel, F. C. 1987, ApJ, 316, 434
Simon, T, Herbig, G. and Boesgaard, A. M. 1985, ApJ, 293, 551
Skumanich, A. 1972, ApJ, 171, 565
Thoul, A. A., Bahcall, J. N. and Loeb, A. 1994, ApJ, 421, 828
This preprint was prepared with the AAS LATEX macros v4.0.
– 11 –
Figure Captions
Figure 1: The evolutionary tracks in the theoretical HR-diagram. Several isochrones have also
been drawn for 1 Gyr to 15 Gyr. The lower age isochrones are indisguishable from the 1 Gyr
isochrone on the scale of this diagram.
Figure 2: The rotation period as a function of age and mass. The slope near the main sequence
of each mass curve corresponds to the Skumanich law with a proportionality factor which
depends on mass.
Figure 3: The non-local convective turnover time as a function of age and mass. Near the main
sequence, τc remains nearly constant with time for a given mass.
Figure 4: The dynamo number (∝ N −2
R ), often used as a measure of magnetic activity strength,
as a function of age and mass. This plot (together with Figures 8 and 9) provides the means
of calibrating magnetic activity indices along the main sequence of a star cluster.
Figure 5: The global and local convective turnover times as a function of effective temperature
and age.
Figure 6: Global convective turnover time as a function of rotation period and age.
Figure 7: Rotation period as a function of effective temperature and age.
Figure 8: The dynamo number as a function of effective temperature and age.
Figure 9: The dynamo number as a function of rotation period and age.
– 12 –
Table 1.
Input physics
Parameter
Input
Mass
Mixing length ratio
Weight fraction of hydrogen, X
Weight fraction of all heavy elements, Z
Mixture of heavy elements
The exponent in the wind model, N
The constant factor in the wind model, fk
The diffusion coefficient, fc
Initial rotation velocity
Opacity tables
Atmosphere
Equation of state
The microscopic diffusion
0.5 ∼ 1.2M⊙
1.86315
0.70952
0.01926
Anders-Grevesse(1989)
1.5
17.4837
0.05575
30 km s−1
OPAL
with Kurucz opacity tables for low temperature
Kurucz model atmosphere
The standard implementation
with Debye-Huckel correction
Michaud and Proffitt (1993)
– 13 –
Table 2.
Isochrones
M/M⊙
log Tef f
(B − V ) a
log L/L⊙
b
τc
(day)
N −2
R
c
Rotation Period
(day)
0.5
0.6
0.7
0.8
0.9
1.0
1.1
1.2
0.5
0.6
0.7
0.8
0.9
1.0
1.1
1.2
0.5
0.6
0.7
0.8
0.9
1.0
1.1
1.2
3.54
3.59
3.64
3.68
3.72
3.75
3.78
3.80
3.53
3.59
3.64
3.68
3.72
3.75
3.78
3.80
3.53
3.59
3.64
3.68
3.72
3.75
3.78
3.80
1.55
1.37
1.16
0.99
0.83
0.69
0.57
0.49
1.60
1.37
1.16
0.99
0.83
0.69
0.57
0.49
1.60
1.37
1.16
0.99
0.83
0.69
0.57
0.49
0.20 Gyr
-1.50
-1.19
-0.88
-0.61
-0.36
-0.14
0.06
0.24
0.50 Gyr
-1.53
-1.18
-0.87
-0.60
-0.35
-0.13
0.07
0.26
0.70 Gyr
-1.53
-1.18
-0.87
-0.60
-0.35
-0.13
0.08
0.27
125.10
104.06
83.19
66.50
53.15
40.27
27.49
13.95
137.40
106.56
84.84
67.18
53.07
39.97
26.97
13.14
140.32
107.55
84.88
67.40
52.96
40.00
26.76
12.87
6.18
27.58
28.40
24.13
18.90
13.30
7.73
2.48
11.13
13.57
11.77
9.31
7.01
4.82
2.56
0.69
9.26
10.14
8.42
6.57
4.92
3.37
1.77
0.46
23.41
9.18
7.23
6.31
5.69
5.19
4.67
4.13
19.19
13.44
11.46
10.25
9.33
8.57
7.94
7.42
21.32
15.69
13.59
12.22
11.12
10.21
9.44
8.81
– 14 –
Table 2—Continued
M/M⊙
log Tef f
(B − V ) a
log L/L⊙
b
τc
(day)
N −2
R
c
Rotation Period
(day)
0.5
0.6
0.7
0.8
0.9
1.0
1.1
1.2
0.5
0.6
0.7
0.8
0.9
1.0
1.1
1.2
0.5
0.6
0.7
0.8
0.9
1.0
1.1
1.2
3.53
3.59
3.64
3.69
3.72
3.75
3.78
3.80
3.53
3.59
3.64
3.69
3.73
3.76
3.78
3.80
3.53
3.59
3.64
3.69
3.73
3.76
3.78
3.78
1.60
1.37
1.16
0.95
0.83
0.69
0.57
0.49
1.60
1.37
1.16
0.95
0.78
0.65
0.57
0.49
1.60
1.37
1.16
0.95
0.78
0.65
0.57
0.57
1.00 Gyr
-1.53
-1.18
-0.87
-0.59
-0.34
-0.12
0.09
0.29
2.00 Gyr
-1.53
-1.18
-0.86
-0.58
-0.32
-0.09
0.14
0.34
4.55 Gyr
-1.52
-1.16
-0.84
-0.54
-0.26
0.00
0.25
0.44
141.14
108.04
85.43
67.49
52.74
39.86
26.32
12.40
142.19
108.88
85.26
67.28
52.62
39.25
25.00
11.44
144.87
110.05
85.10
66.70
51.21
37.49
24.82
26.42
7.15
7.30
5.90
4.55
3.40
2.31
1.19
0.30
4.11
3.73
2.91
2.22
1.63
1.07
0.53
0.13
1.80
1.54
1.19
0.89
0.65
0.41
0.22
0.28
24.45
18.61
16.29
14.68
13.37
12.28
11.34
10.56
32.61
26.23
23.24
20.98
19.19
17.72
16.23
15.07
49.92
40.97
36.26
32.77
29.83
27.66
25.14
23.40
– 15 –
Table 2—Continued
M/M⊙
log Tef f
(B − V ) a
log L/L⊙
b
τc
(day)
N −2
R
c
Rotation Period
(day)
0.5
0.6
0.7
0.8
0.9
1.0
0.5
0.6
0.7
0.8
0.9
3.53
3.59
3.65
3.70
3.74
3.75
3.54
3.60
3.66
3.71
3.74
1.60
1.37
1.11
0.91
0.74
0.69
1.55
1.33
1.07
0.87
0.74
10.00 Gyr
-1.50
-1.13
-0.78
-0.45
-0.11
0.28
15.00 Gyr
-1.48
-1.10
-0.73
-0.34
0.13
145.61
109.32
83.95
65.26
50.58
50.94
144.95
108.32
83.08
64.97
68.27
0.73
0.62
0.47
0.33
0.24
0.32
0.39
0.33
0.24
0.18
0.34
78.38
64.26
56.79
52.86
48.69
42.01
107.04
87.34
78.18
70.55
53.99
a Revised Yale Isochrones and Luminosity Functions (Green et al. 1987)
b τc = R R∗
dr
v(r) , where Rb is the location of the bottom of the surface convection zone, R∗
Rb
is the total radius of the stellar model, and v(r) is the convective velocity as a function of
radius.
c NR = 2πv/αΩHP where v is the local convective velocity at a distance of the half of the
mixing length α HP
above the base of the convection zone, α is the mixing length ratio, Ω
2
is rotational velocity, and HP is the local pressure scale height.
|
0707.1316 | 2 | 0707 | 2007-08-30T18:22:07 | Fast Collisionless Reconnection Condition and Self-Organization of Solar Coronal Heating | [
"astro-ph",
"physics.plasm-ph",
"physics.space-ph"
] | I propose that solar coronal heating is a self-regulating process that keeps the coronal plasma roughly marginally collisionless. The self-regulating mechanism is based on the interplay of two effects. First, plasma density controls coronal energy release via the transition between the slow collisional Sweet-Parker regime and the fast collisionless reconnection regime. This transition takes place when the Sweet--Parker layer becomes thinner than the characteristic collisionless reconnection scale. I present a simple criterion for this transition in terms of the upstream plasma density (n_e), the reconnecting (B_0) and guide (B_z) magnetic field components, and the global length (L) of the reconnection layer: L < 6.10^9 cm [n_e/(10^{10}/cm^3)]^(-3) (B_0/30G)^4 (B_0/B_z)^2. Next, coronal energy release by reconnection raises the ambient plasma density via chromospheric evaporation and this, in turn, temporarily inhibits subsequent reconnection involving the newly-reconnected loops. Over time, however, radiative cooling gradually lowers the density again below the critical value and fast reconnection again becomes possible. As a result, the density is highly inhomogeneous and intermittent but, statistically, does not deviate strongly from the critical value which is comparable with the observed coronal density. Thus, in the long run, the coronal heating process can be represented by repeating cycles that consist of fast reconnection events (i.e., nanoflares), followed by rapid evaporation episodes, followed by relatively long periods (1-hour) during which magnetic stresses build up and simultaneously the plasma cools down and precipitates. | astro-ph | astro-ph | Draft version October 29, 2018
Preprint typeset using LATEX style emulateapj v. 08/22/09
7
0
0
2
g
u
A
0
3
]
h
p
-
o
r
t
s
a
[
2
v
6
1
3
1
.
7
0
7
0
:
v
i
X
r
a
FAST COLLISIONLESS RECONNECTION CONDITION AND SELF-ORGANIZATION OF SOLAR CORONAL
HEATING
Dmitri A. Uzdensky1
(Dated: August 27, 2007)
Draft version October 29, 2018
ABSTRACT
I propose that solar coronal heating is a self-regulating process that keeps the coronal plasma
roughly marginally collisionless. The self-regulating mechanism is based on the interplay of two
effects. First, plasma density controls coronal energy release via the transition between the slow
collisional Sweet -- Parker regime and the fast collisionless reconnection regime. This transition takes
place when the Sweet -- Parker layer becomes thinner than the characteristic collisionless reconnection
scale. I present a simple criterion for this transition in terms of the upstream plasma density (ne),
the reconnecting (B0) and guide (Bz) magnetic field components, and the global length (L) of the
reconnection layer: L . 6 · 109 cm (ne/1010cm−3)−3 (B0/30 G)4 (B0/Bz)2. Next, coronal energy
release by reconnection raises the ambient plasma density via chromospheric evaporation and this,
in turn, temporarily inhibits subsequent reconnection involving the newly-reconnected loops. Over
time, however, radiative cooling gradually lowers the density again below the critical value and fast
reconnection again becomes possible. As a result, the density is highly inhomogeneous and intermittent
but, statistically, does not deviate strongly from the critical value which is comparable with the
observed coronal density. Thus, in the long run, the coronal heating process can be represented by
repeating cycles that consist of fast reconnection events (i.e., nanoflares), followed by rapid evaporation
episodes, followed by relatively long periods ( . 1 hr) during which magnetic stresses build up and
simultaneously the plasma cools down and precipitates.
Subject headings: magnetic fields -- MHD -- Sun: corona -- Sun: magnetic fields -- Sun: flares
1. INTRODUCTION
In this paper I address some aspects of solar coronal
heating (see Aschwanden et al. 2001b and Klimchuk 2006
for recent reviews) in the context of Parker's nanoflare
model (Parker 1972, 1983, 1988). A more concise version
of this work is presented in Uzdensky (2006, 2007).
Since the main heating process in the nanoflare model
is magnetic reconnection, I first discuss what we have
learned about reconnection in the past 20 years or so
(see § 2). I argue that, even though we still do not have
a complete picture of reconnection, there now appears to
be some consensus in the magnetic reconnection commu-
nity about some of its fundamental aspects. One of the
main goals of this paper is to use this emerging knowl-
edge to shed some new light on the old coronal heating
problem.
In this paper I purposefully adopt a rather
conservative approach: I invoke only those very few re-
sults that seem to be relatively firmly established and try
not to rely on those details of reconnection physics that
are still under vigorous debate. Specifically, there is now
strong evidence coming from numerous numerical simu-
lations and some laborious laboratory experiments that
there are two main modes of reconnection: slow Sweet --
Parker reconnection taking place in collisional plasmas
where classical resistive MHD applies; and fast Petschek-
like reconnection in collisionless plasmas (in § 2.1). The
transition between these two regimes seems to be rather
sharp. An approximate condition for this transition
1 Department of Astrophysical Sciences, Princeton University,
Peyton Hall, Princeton, NJ 08544 -- Center for Magnetic Self-
Organization (CMSO); [email protected].
can be formulated as a relationship between the global
length L of the reconnecting system and the electron col-
lisional mean-free path inside the layer, λe,mfp (see § 2.2).
Furthermore, this condition can be cast in terms of the
plasma density n, the reconnecting magnetic field B0,
and L; namely, one can define a critical density nc(L, B0)
below which reconnection switches to the fast regime.
In the strong-guide field case, Bz ≫ B0, the condition
is modified; namely, the critical density for transition
to fast reconnection is suppressed by a factor of order
(Bz ≫ B0)2/3 (see § 2.4).
In § 2.5 I address various
alternative ideas and caveats that may complicate the
above simple picture.
I then apply these results to the active solar corona
(in § 3), viewed withing the nanoflare model. My main
point here is that the corona should be regarded as a self-
regulating machine keeping itself (in a statistical sense)
around marginal collisionality. This conclusion comes
from the interplay between the way the plasma density
controls reconnection via the above collisionless recon-
nection transition, and the way the coronal magnetic en-
ergy release due to reconnection in turn controls the am-
bient gas density via chromospheric evaporation. The
coronal heating process is then highly intermittent and
inhomogeneous; it can be thought of a sequence of char-
acteristic energy-circulation cycles that occur simulta-
neously in a broad range of spatial, temporal, and en-
ergy scales. Each such elementary cycle consists of sev-
eral phases: (1) a fast reconnection event (a nanoflare)
causing (2) an evaporation episode filling the loop with
hot dense plasma, followed by (3) a longer period during
which the magnetic stresses build up and the plasma den-
2
D. A. Uzdensky
sity goes down due to slow radiative cooling (and thermal
conduction). It is the collisionless reconnection condition
that makes this scenario (in particular, the last phase)
possible. Indeed, an important point here is that, even if
a current sheet is formed, it will stay in the slow Sweet --
Parker state if the density is large enough. This will
continue for some time, until the reconnecting magnetic
field becomes strong enough and/or the ambient plasma
density becomes low enough for the system to transi-
tion to the fast collisionless reconnection regime. This
enables a non-trivial amount of free magnetic energy to
be accumulated before a sudden release. This energy
amount, along with the characteristic time-scale between
nanoflares and the characteristic coronal density, is de-
termined, statistically, by the balance between the rate
at which current sheets are created and amplified by the
photospheric footpoint motions and the efficiency of ra-
diative cooling. The collisionless reconnection condition
plays a key role in this process as a mediator and ulti-
mately governs the statistical distribution of the coronal
reconnection events (flares).
In § 4, I discuss some of the open questions that, in
my view, need to (and can) be addressed in the near
future, in order to see whether the physical picture put
forward in this paper is correct and what modifications
should be made to improve it. This section is mostly tar-
geted towards researchers doing numerical simulations of
reconnection and also to experimentalists and observers.
Finally, I present my conclusions in § 5.
I also would like to make a clarifying remark about
the terms "collisional" and "collisionless" that will be
used many times throughout this paper. Often, by
"collisionless" one means a regime in which Ωe > νe,i,
i.e., when the electron gyro-frequency is higher than the
electron-ion collision rate. Equivalently, this means that
λe,mfp > ρe.
In this paper, on the other hand, the
terms "collisional" and "collisionless" will involve com-
paring λe,mfp not with the gyro-radius, but instead with
the system size L, or, more precisely, with L divided by
the square root of the ion/electron mass ratio. As will
be explained in §§ 2.2 -- 2.3 [specifically, see eq. (4) and
the discussion at the end of § 2.3], this choice of termi-
nology turns out to be more relevant to the reconnection
problem.
2. FAST COLLISIONLESS RECONNECTION CONDITION
2.1. The Existence of Two Reconnection Regimes
Magnetic reconnection research started 50 years
ago with the Sweet -- Parker (SP) theory (Sweet 1958;
Parker 1957) for solar flares. As was realized immediately
at that time, this relatively simple and elegant theory
could not reproduce the very short (∼ 103 sec) reconnec-
tion timescales required by flare observations; instead,
it predicted reconnection times of order a few weeks or
months. After several years, however, Petschek (1964)
proposed a modification to the classic Sweet -- Parker the-
ory that apparently resulted in a much higher reconnec-
tion rate, thus eliminating the main contradiction with
observations. He realized that the main bottleneck in the
Sweet -- Parker reconnection model is the need to have a
reconnection layer that is both thin enough for the resis-
tivity to be important and at the same time thick enough
for the plasma to be able to flow out. The result of this
compromise is the famous Sweet -- Parker scaling for the
thickness δSP of the reconnection layer and for the recon-
nection velocity vrec:
δSP
L
=
vrec
VA
=r η
LVA
≡ S−1/2 ,
(1)
where L is the global length of the reconnection layer,
of order the system size, η is the magnetic diffusivity,
and VA is the Alfv´en speed corresponding to the recon-
necting (in-plane) magnetic field component. Since the
Lundquist number in the solar corona is usually very
large (1012 or greater), the layer is very thin and the
resulting reconnection rate is very small. Furthermore,
Petschek (1964) proposed that this difficulty can be cir-
cumvented if the reconnection region has a certain special
structure: the famous Petschek configuration, with four
slow-mode shocks attached to a small Sweet -- Parker cen-
tral diffusion region. As he showed, this structure gave
an additional geometric factor that could lead to a much
faster reconnection rate.
I would like to emphasize the special importance of
Petschek's idea for reconnection in astrophysical sys-
tems, including the solar corona. As I will discuss below,
there are several local physical effects (e.g., Hall effect
or anomalous resistivity) that can indeed broaden the
reconnection layer and prevent it from collapsing down
to the very small Sweet -- Parker thickness δSP. However,
each of these processes comes with its own microscopic
physical scale, e.g., the ion gyro-radius ρi or the ion
collisionless skin-depth di ≡ c/ωpi. These scales are
determined by the local values of the basic plasma pa-
rameters inside the reconnection layer, such as magnetic
field, density, or temperature. The important point is
that they know nothing about the overall global system
size L. Astronomical systems are often astronomically
large, however. That is, L is typically much larger than
any microscopic physical scale δ and than the Sweet --
Parker layer thickness (which is a hybrid length-scale).
Therefore, any simple Sweet -- Parker-like analysis would
give a reconnection rate vrec/VA scaling as δ/L ≪ 1; this
would not be rapid enough to be of any practical inter-
est. Thus, we come to Conclusion I: irrespective of the
actual microphysics inside the layer, Petschek's mecha-
nism, or a variation thereof, is absolutely necessary for
sufficiently fast astrophysical reconnection, including so-
lar flares (and even micro- and nanoflares).
Because Petschek's model was able to reproduce the
very short observed flare timescales, it quickly became
very popular and people have believed in it for the next
20 years. With the advent of computer simulations, how-
ever, the original Petschek model came under fire.
In
particular, several numerical studies (e.g., Biskamp 1986;
Scholer 1989; Ugai 1992, 1999; Ma & Bhattacharjee
1996; Uzdensky & Kulsrud 1998, 2000; Erkaev et al.
2000, 2001; Biskamp & Schwarz 2001; Birn et al. 2001;
Malyshkin et al. 2005; Cassak 2005, 2006) showed that
in resistive MHD with uniform resistivity (and, by infer-
ence, with resistivity that is a smooth function of plasma
parameters, e.g., Spitzer, see Biskamp & Schwarz 2001)
Petschek's mechanism fails and Sweet -- Parker scaling ap-
plies instead. The basic physical reason for this, as has
been elucidated analytically by Kulsrud (2001), (see also
Uzdensky & Kulsrud 2000 and Malyshkin et al. 2005)
Solar Coronal Heating
3
is that the transverse (i.e., perpendicular to the current
sheet) magnetic field component, which is necessary to
sustain Petschek's shocks, is swept away by the Alfv´enic
flow out of the layer so rapidly that it cannot be regen-
erated by the nonuniform resistive merging of the recon-
necting magnetic field component.
In addition, a strong evidence for the existence
of a slow Sweet-Parker reconnection mode in col-
lisional
resistive-MHD plasmas has been obtained
from recent laboratory studies in the Magnetic Re-
connection Experiment (MRX; Yamada et al. 1997)
in the high-collisionality regime
et al. 1998;
Trintchouk et al. 2003; Kuritsin et al. 2006).
(Ji
This enables us to draw Conclusion II: In the col-
lisional regime, when classical resistive MHD applies,
Petschek's fast reconnection mechanism does not work;
slow Sweet -- Parker reconnection process takes place in-
stead.
A very natural question to ask next is whether fast
Petschek-like reconnection possible in a collisionless
plasma, where resistive MHD does not apply. There is
now a growing consensus that the answer to this question
is YES. In Solar and Space physics, of course, there has
long been plentiful observational evidence for fast colli-
sionless reconnection, in solar flares (e.g., Tsuneta et al.
1984, 1992, 1996; Masuda et al. 1994, 1995; Shibata 1995,
1996; Yokoyama et al. 2001), and in the Earth magne-
tosphere, both in the magnetopause (e.g., Mozer et al.
2002, 2003, 2004) and in the magnetotail (Oieroset et al.
2001; Nagai et al. 2001). The evidence for fast collision-
less reconnection has been further significantly strength-
ened by recent laboratory measurements in the MRX
(Ji et al. 1998, 2004; Yamada et al. 2006) and other de-
vices (e.g., Brown 1999; Egedal et al. 2007; Brown et al.
2006). These laboratory measurements, however, have
not been able to elucidate the special role of the Petschek
mechanism in accelerating reconnection. On the other
hand, over the past decade or so, several theoretical
and numerical studies have indicated that fast reconnec-
tion enhanced by the Petschek mechanism (or a vari-
ation thereof) does indeed take place in the collision-
less regime. Moreover, it appears that there may be
two mechanisms of collisionless reconnection. Physically,
these two mechanisms are very different from each other;
nevertheless, they both appear to lead to the establish-
ment of a Petschek-like configuration, which enhances
the reconnection rate. The two regimes in question are:
• Hall-MHD reconnection, involving two-fluid effects
in a laminar flow configuration (e.g., Mandt et al.
1994; Kleva et al. 1995; Biskamp et al. 1995,
1997; Ma & Bhattacharjee 1996; Lottermoser &
Scholer 1997; Shay et al. 1998, 1999, 2001; Hesse
et al.
1999, 2001;
Birn et al. 2001; Rogers et al. 2001; Pritchett 2001;
Breslau & Jardin 2003; Cassak et al. 2005, 2006;
Daughton et al. 2006).
1999; Bhattacharjee et al.
• Spatially-localized anomalous resistivity due to
plasma micro-turbulence that is excited when the
current density exceeds a certain threshold (e.g.,
Coppi & Friedland 1971; Smith & Priest 1972;
Coroniti & Eviatar 1977; Kulsrud 2001, 2005; Kul-
srud et al. 2005). This seems to lead to a Petschek
configuration with the length of the inner diffusion
region being on the order of the resistivity localiza-
tion scale (e.g., Ugai & Tsuda 1977; Sato & Hayashi
1979; Ugai 1986, 1992, 1999; Scholer 1989; Biskamp
& Schwarz 2001; Erkaev et al. 2001; Kulsrud 2001,
2005; Uzdensky 2003; Malyshkin et al. 2005). The
corresponding reconnection rate is then fast enough
to explain the relatively-short timescales observed
in solar flares (e.g., Kulsrud 2001, 2005; Uzden-
sky 2003).
Recent experimental evidence from the MRX exper-
iment indicates that both anomalous resistivity (e.g,
Ji et al. 2004) and Hall effect (Ren et al. 2005; Ya-
mada et al. 2006; Brown et al. 2006; Egedal et al. 2007)
are present and may be important in collisionless recon-
nection. However, at present, it is still not clear which
one of these two mechanisms works in a given physical
situation (if at all). Also not known is whether these
two mechanisms can operate simultaneously in a given
system and how they interact with each other.
In any case, from the point of view of our present
discussion, the important thing is that one does get a
Petschek-enhanced fast reconnection if the plasma is col-
lisionless (in the sense that will be specified below, see
§ 2.2). Thus, we can draw Conclusion III: a Petschek-
enhanced fast reconnection does happen in the collision-
less regime.
To sum up, there are two regimes of astrophysical mag-
netic reconnection: slow Sweet -- Parker reconnection in
resistive-MHD with classical collisional resistivity, and
fast Petschek-like reconnection in collisionless plasmas.
In other words, in order for reconnection to be fast, it
needs Petschek's mechanism to operate and that in turn
requires the reconnection layer to be collisionless.
In fact, whenever we observe violent and rapid ener-
getic phenomena that we interpret as reconnection, it is
always in relatively tenuous plasmas. I am not aware of
any counter-examples and would be would be very inter-
ested in learning about them. Is there any evidence for
fast large-scale reconnection events in collisional astro-
physical environments?
2.2. The Fast Reconnection Condition: Zero Guide
Field Case
How can one quantify the transition between the two
regimes? For clarity, let us first consider the case with
zero guide field, Bz = 0.
[In the Bz 6= 0 case some of
the arguments and results presented in this section will
be somewhat modified; however, the main concepts and
conclusions will remain similar, as we show in the next
section (see § 2.4).]
When there is no guide field (strictly anti-parallel field
merging), the condition for fast reconnection can be for-
mulated roughly as (Kulsrud 2001, 2005; Uzdensky 2003;
Cassak et al. 2005; Yamada et al. 2006)
δSP < di ≡
c
ωpi
.
(2)
What this condition means is the following. As a recon-
nection layer forms, its thickness δ becomes smaller and
smaller. If condition (2) is not satisfied, then this thin-
ning saturates at δ = δSP; then, reconnection proceeds
in the slow Sweet -- Parker regime. On the other hand, if
4
D. A. Uzdensky
condition (2) is satisfied, then various two-fluid and/or
kinetic effects kick in as soon as δ drops down to about di
or so, i.e., well before the resistive term becomes impor-
tant. Then, the reconnection processes necessarily in-
volves collisionless, non-classical-resistive-MHD physics.
This results in a rapid increase in the reconnection rate,
as has been recently documented experimentally in the
MRX (Yamada et al. 2006).
Note that, by construction, δSP that enters in equa-
tion (2) is to be calculated using the classical Spitzer
resistivity. Therefore, and since we are talking about col-
lisional and collisionless reconnection, it is instructive to
express the above condition in terms of the classical elec-
tron mean free path due to Coulomb collisions, λe,mfp. It
is pretty straight-forward to show (Yamada et al. 2006)
that
,
(3)
δSP
di
∼(cid:18) L
λe,mfp(cid:19)1/2(cid:18)βe
me
mi(cid:19)1/4
Here, βe is the ratio of the electron pressure inside the
layer to the pressure of the reconnecting magnetic field
component (B2
0/8π) outside the layer. Thus, using equa-
tion (2), we see that a reconnection layer is collisionless
when
e
λe,mfp .
(4)
L < Lc ≡ λe,mfppmi/βeme ≃ 40 β−1/2
The collisionless reconnection condition in a similar form
has been first published by Yamada et al. (2006). Their
condition actually differs from equation (4) by a factor
of 2. Since the discussion here is very qualitative, I re-
gard this difference as unessential. Moreover, to make
the present discussion more clear, I systematically ignore
numerical factors of order 1 everywhere in this paper.
We can go a little bit further. The mean free path
λe,mfp can be written as
k2
BT 2
e
λe,mfp ∼
nee4 log Λ
≃ 7 · 107cm n−1
10 T 2
7 ,
(5)
where we set the Coulomb logarithm equal to 20 and
where n10 and T7 are the central layer electron den-
sity ne and temperature Te given in units of 1010 cm−3
and 107 K, respectively.
Combining equations (4)
and (5), the criterion for fast collisionless reconnection
can now be formulated as a condition on the layer's
length L in terms of the values of ne and Te at the center
of the Sweet -- Parker reconnection layer:
e
e
n−1
10 T 2
λe,mfp ≃ 3·109cm β−1/2
L < Lc ≡ 40 β−1/2
7 . (6)
Notice that the critical layer length Lc strongly de-
pends on the central electron temperature Te. The
main cause for this high sensitivity is the strong tem-
perature dependence of the electron mean free path and
hence Spitzer resistivity. It is therefore very important
to figure out what the electron temperature inside the
Sweet -- Parker reconnection layer should be. In a situa-
tion where the ambient (upstream) pressure is already
high, βupstream ≥ 1, the conversion of magnetic en-
ergy to plasma via reconnection is not going to affect
the plasma temperature and density considerably.
In
this case, both the plasma temperature and the density
should be roughly uniform across the layer; then one can
just substitute their upstream values into equation (6).
The story then ends here.
However,
in this paper we will be mostly inter-
ested in environments that are (globally) magnetically-
dominated (and hence almost force-free), such as the so-
lar corona and coronae of other stars and accretion disks.
In this case, the ambient thermal pressure is smaller than
the reconnecting magnetic field pressure. It is then im-
portant to distinguish the temperature inside the layer
from the ambient coronal temperature (which is much
lower). The simplest first step towards determining Te is
to note that, in the case with zero guide field, the condi-
tion of pressure balance across the layer dictates that
β ≡
8πnekB(Te + Ti)
B2
0
= 1 .
(7)
Next, assuming that Te inside a Sweet -- Parker reconnec-
tion layer is approximately equal to (and in any case not
much smaller than) the ion temperature Ti -- the self-
consistency of this assumption will be demonstrated at
the end of this section -- we see that βe should generally
be close to 1/2 (and in any case of order 1).
Then, using the pressure-balance condition (7), we can
express the central electron temperature in terms of B0
and ne as
Te =
B2
0 /8π
2kBne
≃ 1.4 · 107 K B2
1.5 n−1
10 ,
(8)
where B1.5 is the outside magnetic field B0 expressed
in units of 30 G. Upon substituting this expression into
equation (5), we get
λe,mfp ≃ 1.5 · 108cm n−3
10 B4
1.5 ,
(9)
and so the fast collisionless reconnection condition can
be written in the final form as
L < Lc(n, B0) ≃ 6 · 109cm n−3
10 B4
1.5 .
(10)
We see that the condition L < Lc is often satisfied in the
solar corona (e.g., in solar flares) and is definitely always
satisfied in the Earth magnetosphere.
2.3. Plasma Temperature and Density inside the
Sweet -- Parker Reconnection Layer
Note that the density that enters in the above expres-
sions is, by definition, the density at the center of the
reconnection layer. In general, it is not known a priori
and we would like, instead, to get expressions involving
the background (upstream) density. Equation (7) gives
us one relationship between the two desired quantities
(central ne and Te). However, we need to know both of
them and by itself, equation (7) does not tell us whether
the central pressure is increased (from a very low ambi-
ent value) to the required equilibrium level by raising the
density or the temperature. I will now argue, however,
that, in the regime that is relevant here, the latter is
more likely to be the case, i.e., that the pressure increase
is mostly due to the increased temperature, whereas the
density should change relatively little.
Indeed, as is evident from the above discussion, the
main reason for the strong density dependence in equa-
tion (10) is the Te-dependence of the classical Spitzer
resistivity, combined with the pressure balance condi-
tion (7). Therefore, what we are mostly interested in here
are the central electron density and temperature in the
Solar Coronal Heating
5
context of the resistive Sweet -- Parker theory. To deter-
mine them, we need more detailed information. Namely,
we need to consider one more equation that we have not
discussed up until now -- the equation of energy conser-
vation. In general this equation should take into account
Ohmic heating and all the relevant loss mechanisms, such
as advection, radiation, and electron thermal conduction.
Let us consider a fluid element as it travels through the
reconnection layer. During its transit, the fluid element is
subject to Ohmic heating and to all the above loss mech-
anisms. The characteristic time for the advection term
is of course just the Alfv´en transit time τA ∼ L/VA --
the time a fluid element spends inside the layer. During
this time the element is continuously heated by Ohmic
heating, whose rate per unit volume can be estimated
within the Sweet -- Parker theory as
4πδ(cid:19)2
η′j2 ∼ η′(cid:18) cB0
∼
B2
0
4π
η
δ2 =
B2
0
4π
ηS
L2 =
B2
0
4π
L
VA
, (11)
where the magnetic diffusivity η is related to the resistiv-
ity η′ via η = η′c2/4π.2 Thus, during its transit, the fluid
element acquires just the right amount of heat to raise
its temperature to about the equipartition level given by
equation (8) (Uzdensky 2003). We now need to see under
what conditions one can ignore various loss mechanisms,
namely, radiation and thermal conduction.
First, it is easy to see that, in the case of the solar
corona, the radiative losses are indeed negligible on the
time scales of interest (τA). Indeed, the radiative cooling
time is determined by the cooling function Q(T ) via
τrad ∼
2cvnkBT
n2Q(T )
=
3kBT
nQ(T )
.
(12)
In the temperature range of interest to the solar corona,
106 K ≤ Te ≤ 108 K, the cooling function varies between
10−23 and 10−22 erg sec−1 cm3 (e.g., Rosner et al. 1978;
Priest 1984, p. 88; Cook et al. 1989). Then, taking
n = 1010 cm−3, the typical radiative cooling times are of
order 10 min (for T = 2 · 106 K) and longer (up to many
days for T = 108 K). In any case, this is much longer
than the characteristic Alfv´en crossing time τA which is
usually a few seconds.
On the other hand, the smallness of the radiative losses
in other astrophysical situations cannot be taken for
granted, especially in High-Energy Astrophysics appli-
cations, such as coronae of accretion disks around black
holes and neutron stars and also gamma-ray bursts. If
radiative losses are important, then the above estimates
for the central temperature would no longer apply and
correspondingly the condition for transition to the fast
reconnection regime would have to be modified.
Next, going back to the solar corona, we need to con-
sider the effect of electron thermal conduction on the
central temperature. Note that electrons are still well
magnetized throughout most of the reconnection layer
(this will be even more so in the presence of a guide
field). Therefore, the electron thermal conduction is
2 I would like to add one extra word of caution here. The Sweet --
Parker theory that we are relying on here does not take into account
the large variation of the Spitzer resistivity across the layer as the
electron temperature increases by a large factor with respect to its
relatively low ambient value. This brings in an extra uncertainty
into our estimates.
highly anisotropic, and so we shall estimate the effects
of parallel and perpendicular thermal conduction sepa-
rately. We shall start with the heat conduction along the
magnetic field. Since we interested in the collisional case,
λe,mfp ≪ L, the parallel heat transport is in the diffu-
sive regime and can be described by a one-dimensional
random walk of electrons along the magnetic field. The
characteristic parallel thermal conduction time-scale can
then be estimated as
τe,cond,k ∼
λe,mfp
ve,th (cid:18) L
λe,mfp(cid:19)2
=
L
L
ve,th
λe,mfp
∼ τAs(cid:18) me
βmi(cid:19) L
λe,mfp
,
(13)
where τA and β correspond to the reconnecting field B0.
Using expression (4), we can write this expression as
τe,cond,k ∼ τA
L
Lc
β−1 .
(14)
In the magnetically-dominated environment of the so-
lar corona, where the thermal pressure outside of the
reconnection region is small, we expect β ∼ 1; it can
conceivably be smaller, but we never expect it to be sig-
nificantly higher than 1. Therefore, we estimate that
τe,cond,k ≥ τA
L
Lc
.
(15)
Thus, we see that if we are in the regime of inter-
est -- that is, in the expected regime of applicability
of the collisional Sweet -- Parker theory, L > Lc, -- then
τe,cond,k > τA and hence the energy losses due to parallel
electron thermal conduction are not important.
A similar line of arguments can be made to show that
the characteristic time for perpendicular (across the re-
connection layer) electron thermal conduction is in fact
automatically always comparable to the Alfv´en cross-
ing time.
Indeed, the cross-field electron conduction
timescale in the Sweet -- Parker regime can be estimated
as
τe,cond,⊥ =
,
(16)
δ2
SP
De,⊥
where De,⊥ ∼ ρ2
e ve,th/λe,mfp is the classical electron
collisional diffusivity across the magnetic field. Using
βe ∼ 1, we see that τe,cond,⊥ should scale as
τe,cond,⊥ ∼ τA
δ2
SP
d2
i r mi
me
λe,mfp
L
.
(17)
Upon plugging equation (3) into this expression, we im-
mediately find that all of the factors on the right-hand-
side cancel and we simply see that
τe,cond,⊥ ∼ τA ,
(18)
i.e., that the two timescales are automatically compara-
ble to each other.
Combined with our previous result regarding the ef-
fect of the parallel thermal conduction, we conclude
that energy losses due to both parallel and perpendic-
ular electron thermal conduction out of the layer are, at
best, only marginally important in the collisional Sweet --
Parker regime. Heat losses are still present but are not
likely to cause a decrease in the central electron temper-
ature by more than a factor of order unity. This means
6
D. A. Uzdensky
that the jump in the plasma pressure, required to main-
tain the pressure balance with the outside reconnecting
magnetic field, should be attributed mostly to the in-
crease in the plasma temperature (due to Ohmic heat-
ing), whereas the density should not vary across the layer
by more than a factor of a few. Thus, for our rough es-
timates, the density that enters in our estimate (8) for
the central temperature can be taken to be the ambi-
ent plasma density. As long as we are in the collisional
regime, L > Lc, this result is consistent with the energy
balance inside the layer that includes Ohmic heating,
heat advection by the bulk plasma outflow, and electron
thermal transport. As we shall see in the next section,
this will be true even in the presence of a guide field.
It is also interesting to check that the assumption Ti ≃
In principle, this could
Te is not strongly violated.
be a worry, since the collisional electron-ion energy-
equilibration rate is suppressed by the large mass ratio;
hence, in general, the electron and ion temperatures in
the layer need not be equal. For example, ions might have
been much hotter than the electrons and hence might
have provided the bulk of the pressure support against
the outside magnetic field. The electron temperature
in this case would have been far below the equiparti-
tion value (about 107 K). Correspondingly, the electron
mean-free path would be much lower than that given by
equation (9).
The electron-ion temperature equilibration time, τEQ,
is by a factor of mi/me longer than the electron colli-
sion time, τe. Making use of the the pressure balance
cs ∼ (Te/mi)1/2 ∼ VA, we can express τe in terms
of λe,mfp and the Alfv´en speed as τe ≃ λe,mfp/ve,th ∼
(λe,mfp/cs)pme/mi ∼ (λe,mfp/VA)pme/mi. Then, we
can compare τEQ with the Alfv´en crossing time, which
is the characteristic time a a fluid element spends inside
the layer:
τEQ ∼ τA
λe,mfp
L r mi
me
∼ τA
Lc
L
.
(19)
Thus we see that, if we are in the collisional regime as
defined by equation (4), the electrons and ions experi-
ence enough collisions with each other to equalize their
temperatures while they transit through the layer. Thus,
Ti ≃ Te should be a decent approximation in the Sweet --
Parker regime.
Note that the plasma density also enters the collision-
less reconnection condition via di. The straight-forward
logic of the above picture dictates that this quantity is
to be estimated in the collisional regime. However, it
is plausible that a corresponding estimate based on the
collisionless regime also has some relevance; in this case
it is interesting to compare the two. The task of estimat-
ing the central ne in the collisionless regime is more diffi-
cult than estimating it in the Sweet -- Parker regime, since
there are many uncertainties here. However, it is also less
important, since most of the strong density dependence
in equation (10) comes from the the steep temperature
(and hence, indirectly, density) dependence of the Spitzer
resistivity. Nevertheless, we still shall try to discuss the
question of whether the density can vary significantly
across a collisionless reconnection layer. If the fast colli-
sionless reconnection is due to anomalous resistivity, then
the effective λe,mfp is determined by wave-particle colli-
sions and hence will be relatively short. This will result,
again, in suppression of parallel electron thermal conduc-
tion and hence in high temperature and correspondingly,
according to the pressure balance condition (7), in a rela-
tively mild density change with respect to the upstream
value (see also Uzdensky 2003). Then the above esti-
mates are likely to be still valid (although in this case
one might have to deal with a possibility that Ti ≫ Te).
On the other hand, if the fast reconnection is due to
the Hall effect, then we also may expect the density to
be roughly uniform across the layer. This is because
the ions become demagnetized on scales less than di and
hence cannot develop structures on smaller scales. The
dynamics of ions inside the reconnection layer can be
summarized by noting that they are pulled directly into
the layer by the bipolar electric field, which accelerates
them up to the Alfv´en speed, and then are ejected out
along the layer by the Lorentz force associated with the
quadrupole out-of-plane magnetic field (e.g., Uzdensky &
Kulsrud 2006). At any rate, one does not expect to see
a significant variation of ion, and hence electron, den-
sity across the Hall reconnection layer (apart from the
di-thick density depletion layers that run along the sep-
aratrices, as reported by Shay et al. 2001).
I also would like to make a little side remark about my
use of terms here. In this paper I use the terms "colli-
sional" and "collisionless" a lot and I would like to clarify
what I mean by them. Notice that the collisionless re-
connection condition (4) is different from the customary
usage of the term "collisionless". Often, by "collision-
less" one means a regime in which Ωe > νe,i, i.e., when
the electron gyro-frequency is faster than the electron-ion
collision rate. Equivalently, this means that ρe < λe,mfp.
In this paper, on the other hand, the terms "collisional"
and "collisionless" involve comparing λe,mfp not with the
electron gyro-radius, but instead with the system size L
divided by the square root of the ion/electron mass ratio,
as seen from (4).
Finally, I would like to make a comment about the use
of the term "collisionless". As I mentioned in § 1, in
the present paper this term is understood in the sense of
equation (4). That is, a system is called "collisionless"
when λe,mfp & L/40, roughly speaking. This definition
is different from the condition λe,mfp > ρe that is of-
ten used. Note that, when λe,mfp > ρe, or, equivalently,
Ωe > νe,i, then the Hall term in the generalized Ohm law
formally cannot be neglected. In plasmas of interest to
this paper, e.g., in the solar corona, this condition is al-
ways very well satisfied. In this sense, the coronal plasma
is definitely always collisionless. Our collisionless recon-
nection condition (2), on the other hand, is not always
satisfied.
In fact, there is a large and astrophysically-
important region in parameter space where δSP > di but
where the plasma is nevertheless collisionless in the sense
that Ωe > νe,i, and where therefore the Hall effect may be
significant (see also Cassak et al. 2005). In the above dis-
cussion I have assumed that, in this intermediate regime,
reconnection still proceeds slowly, in the purely resistive
Sweet -- Parker mode, as suggested by numerous resistive-
MHD simulations. One should be aware, however, that
most of these simulations completely ignore the Hall ef-
fect; thus, it is not clear to what extent their results can
be relied upon in this intermediate regime. In fact, the
recent numerical work by Cassak et al. (2005), who in-
Solar Coronal Heating
7
clude both resistivity and Hall effect, indicates that both
slow Sweet -- Parker and fast Hall modes may operate in
the intermediate regime, depending on the previous his-
tory of the system. In particular, they observe that, when
one considers a gradual initial thinning of the reconnec-
tion layer, the system always stays in the slow regime,
even if Ωe > νe,i, as long as δSP > di. Since it is ex-
actly this scenario that is of interest in the context of
our model, we feel we can use condition (2) with confi-
dence.
2.4. Reconnection with a Guide Field
In the real world the zero-guide-field case considered
in the previous section is rare. It may be encountered in
some special circumstances; for example it may be rele-
vant to reconnection in the Earth magnetotail. Gener-
ally, however, there is some guide field, Bz 6= 0, present.
This is especially so in the context of the nanoflare model
for heating of solar coronal loops by braiding of individ-
ual elemental flux strands that comprise a larger loop.
In this case the guide field (i.e., the axial field along the
general loop direction) is always present and is generally
much stronger than the reconnecting field B0. Another
situation where guide field is dominant over the recon-
necting magnetic field is laboratory plasma devices, such
as tokamaks. Note that reconnection releases the energy
of the weaker B0-field; the energy of the guide field is
not available. However, the presence of a guide field,
especially if it is strong, may, in principle, change the
structure of the reconnection layer. Therefore, in this
section we consider the effect of the guide field on our
collisionless reconnection condition.
The presence of the guide field modifies the collision-
less reconnection condition. This is because, whereas the
characteristic layer thickness in resistive MHD is not af-
fected by the guide field and is still given by the same
Sweet -- Parker scaling (1), the characteristic scale for col-
lisionless reconnection layer changes. In particular, ac-
cording to the published literature (e.g., Kleva et al.
1995; Rogers et al. 2001; Cassak et al. 2007), the layer
thickness for Hall collisionless reconnection becomes the
ion acoustic Larmor radius, ρs ∼ cs/Ωi, where Ωi corre-
sponds to the total upstream magnetic field and where
cs ∼ (Te/mi)1/2 is the ion-acoustic speed. This result
has been recently confirmed experimentally (Egedal et al.
2007). Then, the criterion for the onset of fast reconnec-
tion in the Hall-MHD regime becomes (e.g., Cassak et al.
2007)
δSP < ρs .
(20)
According to the general scheme presented in the pre-
ceding section, we first need to consider the Sweet-Parker
layer. Again, we want to estimate the value of the Spitzer
resistivity at the center of the reconnection layer. There-
fore, we need to determine the central electron temper-
ature. One of the difficulties here is that, if there is a
guide field, then the cross-layer force-balance in the form
of equation (7) is no longer applicable. This is because
the pressure of the reconnecting component of the out-
side magnetic field, B2
0 /8π, no longer has to be balanced
at the center of the layer by the plasma pressure alone. It
can be partially balanced by an increase in the guide field
pressure due to a modest compression. In particular, a
very strong guide field, Bz ≫ B0, dominates the pressure
both inside and outside of the current layer and can be
changed only slightly. Therefore, it essentially ensures
incompressibility: nin ≃ nout. The temperature then
decouples from the pressure balance and has to be deter-
mined from some other considerations, namely, from the
energy equation. Fortunately, most of the analysis pre-
sented in the previous section can be carried over to the
strong-guide-field case without much change. In partic-
ular, the plasma experiences Joule heating as it travels
through the layer and, as long as one is in the collisional
regime, the parallel electron thermal conduction is not
able to cool the layer by a large factor. Therefore, we ex-
pect the temperature inside the layer to increase roughly
to the equipartition level given by equation (8). However,
this equation can no longer be regarded as equality, but
just as a rough estimate. For the purposes of this paper,
though, this is good enough.
Now,
let us discuss the right-hand side of condi-
tion (20). If the guide field is not very large compared
with the reconnecting field, then one expects ρs ∼ di,
and hence the criterion (2) is not changed substantially.
However, in the strong-guide-field case, we expect to have
ρs ≪ di (assuming, as we always do throughout this pa-
per, that upstream β ≪ 1). Then, the criterion for tran-
sition to fast collisionless Hall reconnection is modified
structurally. It now becomes
δSP < ρs ∼ di (cid:18) 8πneTe
z (cid:19)1/2
B2
∼ di β1/2
e
B0
Bz
,
(21)
where βe is based on the central electron temperature Te
and density and on the upstream reconnecting field com-
ponent B0. Note that, according to logic of our paper,
this βe is to be estimated in the collisional regime. Nev-
ertheless, it may also be useful to get some feeling for
what βe (and hence ρs) should be immediately after the
transition to fast reconnection. As we noted above, a
strong guide field makes the plasma nearly incompress-
ible, so that ne is roughly uniform. The question then
is what one should take for the electron temperature in-
side the layer? There is a considerable uncertainty here.
In particular, there is no good reason for Ti to be close
to Te in the collisionless regime. For example, Hsu et al.
(2001) present experimental evidence for strong ion heat-
ing Ti ≫ Te inside a collisionless reconnection layer in the
no-guide-field case in the MRX, although the rise in Ti
is considerably smaller in the presence of a guide field.
In any case, with all these reservations, the simplest ap-
proach one can adopt is to take βe ∼ 1 in the above
expression. Then, the condition for fast collisionless re-
connection in the strong guide field case becomes
δSP
ρs
∼(cid:18) L
λe,mfp(cid:19)1/2 (cid:18) me
mi(cid:19)1/4 Bz
B0
< 1 ,
(22)
which, after our usual manipulations, can be rewritten
as
L < Lc(Bz) =r mi
me
Bz(cid:19)2
λe,mfp (cid:18) B0
Bz(cid:19)2
1.5(cid:18) B0
10 B4
≃ 6 · 109 cm n−3
(23)
Thus, the critical global length Lc in this case becomes
8
D. A. Uzdensky
smaller by a factor (Bz/B0)2 and so the collisionless re-
connection condition is harder to satisfy. Also note that,
for fixed ne and Bz, Lc becomes very sensitive to the
reconnecting field component: Lc ∼ B6
0.
2.5. Caveats and Alternatives
With all the above said, there are still many caveats
and alternative ideas that may potentially alter the above
picture. It is important to be aware of them and try to
gage their effect on the existence and the form of the fast
collisionless reconnection condition. I will discuss some
of them in this section.
First, there are a number of ideas that may, if cor-
rect, represent a potential challenge to our Conclusion II
(that collisional resistive-MHD reconnection is slow). For
example, Lazarian & Vishniac (1999) suggested that re-
connection may be relatively rapid in pure resistive MHD
in the presence of externally imposed three-dimensional
MHD turbulence (see, however, Kim & Diamond 2001
for a dissenting view). As far as I know, this idea has
not yet been tested in three-dimensional (3D) numeri-
cal simulations, although it is definitely an interesting
possibility that deserves further study. Such studies are
indeed already underway (Vishniac 2007, private com-
munication). Furthermore, even in the two-dimensional
case, resistive-MHD Sweet -- Parker-like current layer may
become unstable to tearing instability when its aspect
ratio exceeds a certain large number (probably a hun-
dred) (Bulanov et al. 1978; Forbes & Priest 1983;
Biskamp 1986; Lee & Fu 1986; Malyshkin 2005, private
communication; Loureiro et al. 2007). Tearing may effec-
tively result in an enhanced hyper-resistivity that would
dominate over the normal resistivity (e.g., Strauss 1988).
What will be the effect of the nonlinear development of
this instability on the overall reconnection rate is still
not clear. In addition, Dahlburg et al. (1992) identified a
mechanism, based on a secondary ideal-MHD instability,
that leads to the excitation of fully three-dimensional tur-
bulence within a reconnecting current layer. Finally, in a
recent numerical simulation with both uniform resistivity
and Hall effect, Cassak et al. (2005) observed an interest-
ing hysteresis phenomenon: as the resistivity was turned
down, there was a transition from slow Sweet -- Parker to
fast Hall reconnection, in accordance with condition (2).
However, as the resistivity was turned back up to a level
higher than that corresponding to (2), the system still
stayed in the fast mode. This puzzling behavior needs to
be investigated further.
Second, there are recent studies relevant to Conclu-
sion III (that collisionless reconnection is fast). In par-
ticular, even though most of the authors agree that Hall
reconnection is much faster than Sweet -- Parker, there is
still a significant disagreement about exactly how fast it
actually is. For example, Drake and his collaborators
report a universal reconnection rate of about 0.1 VA,
independent of the electron microphysics that is ul-
timately responsible for breaking the field lines (e.g.,
Biskamp et al. 1997; Shay et al. 1998, 1999, 2001; Cas-
sak et al. 2005; see also Birn et al. 2001). On the other
hand, others argue that it should not be universal but
should instead be a function of the system parameters,
such as the electron/ion mass ratio (e.g., Bhattacharjee
et al. 2001).
In addition, Daughton et al. (2006) re-
cently raised a concern about the validity of periodic
boundary conditions used in the majority of previous
numerical studies.
In particular, they argued that the
reported very high reconnection rate may be an arti-
fact of such boundary conditions. Furthermore, they
performed 2D particle simulations with the appropriate
open boundary conditions and observed that, after an
initial transient peak consistent with the high value re-
ported by Shay et al.
(1999, 2001), the reconnection
rate went down and settled at a significantly lower level
(Daughton et al. 2006). Similar conclusions were reached
independently by Fujimoto (2006) using periodic 2D par-
ticle simulations but in a very large. After a year-long
debate, however, this controversy finally seems to have
been resolved. It is now agreed by both sides that the
electron dissipation region develops a two-scale structure
with an extended high-velocity electron jet in the outflow
direction; however, the dimensionless reconnection rate
still remains high (vrec/VA ≫ di/L), although perhaps
not quite as high as previously believed (e.g., Shay et al.
2007; Karimabadi et al. 2007).
Finally, I would like to mention that the picture pre-
sented in the above two sections, implies certain physi-
cal conditions and thus is not expected to be universally
valid in all astrophysical environments. For example, in
the case of weakly-ionized plasma in molecular clouds
(e.g., in the context of star formation), reconnection dy-
namics is strongly modified by recombination processes
and may become significantly faster than Sweet -- Parker
even in the collisional regime (Zweibel 1989; Heitsch
& Zweibel 2003). Also,
in the case of reconnection
of super-strong magnetic fields in magnetar magneto-
spheres (B0 & 1014 G), the density of the released mag-
netic energy is so large that the pressure inside the layer
is dominated by radiation. The resulting temperature
is so high that a large number of electron-positron pairs
is produced; this means that the number of particles is
not conserved, and so the above estimates would have
to be modified (see, e.g., Thompson 1994; Lyutikov &
Uzdensky 2003; Uzdensky & MacFadyen 2006).
Other issues that are worth-mentioning and that may
substantially complicate the simple physical picture pre-
sented in this paper are: impulsive, bursty reconnection
(e.g., Bhattacharjee et al. 1999; Bhattacharjee 2001;
Daughton et al. 2006, 2007, in preparation); and recon-
nection in three dimensions (e.g., Longcope 1996).
3. SOLAR CORONAL HEATING AS A SELF-REGULATED
PROCESS
3.1. General Picture
The physical picture presented in this paper has a num-
ber of interesting implications for the solar corona that
I will discuss in this section.
At the present, the theory of magnetic reconnection
has not yet reached the level of maturity that would al-
low us to make accurate predictions regarding when the
transition between the slow and fast reconnection modes
will occur. Moreover, there is still no consensus on the
exact nature of the fast collisionless reconnection mecha-
nism and on the actual scalings of the reconnection rate
in either fast or slow reconnection (see section § 2.5 for
more discussion). However, what is important for us in
this paper is just the sole fact of existence of such a tran-
sition and that it has to do with the plasma collisionality.
The transition doesn't actually have to be razor-sharp,
Solar Coronal Heating
9
and in reality there may be not two but more different
regimes with different reconnection rates and hence more
than one such transition. Also, the transition condition
may depend on other physical parameters, such as the
presence of a guide field (see § 2.4) or, perhaps, the ef-
fectiveness of radiative cooling, which we have omitted
in this paper. Whatever the case, the existence of two
reconnection regimes with a collisionality-related transi-
tion between them is a very important fact and we would
like to exploit it in various astrophysical applications. In
this section, I will apply this condition [namely, eqs. (4)
and (10)] to the problem of solar coronal heating.
I propose that coronal heating is a self-regulating pro-
cess that works to keep the corona marginally collision-
less, in the sense of equations (2) and (4) (see also Uz-
densky 2006, 2007).
According to Parker's nanoflare theory (Parker 1972,
1983, 1988), as long as the twisting and braiding of coro-
nal loops by photospheric footpoint motions and flux
emergence episodes keep producing thin current sheets
in the corona (e.g., van Ballegooijen 1986), magnetic dis-
sipation in these current sheets leads to continuous coro-
nal heating.
In reality, this heating, of course, is not
uniformly distributed; it is instead strongly intermittent,
localized in both space and time (Rosner et al. 1978;
hereafter RTV78). This basic picture of coronal heating
has been shown to work in several 3D MHD numerical
simulations (e.g., Galsgaard & Nordlund 1996; Hendrix
et al. 1996; Gudiksen & Nordlund 2002, 2005).
Now, the overall, integrated heating density, i.e., the
rate of magnetic dissipation per unit volume, depends on
the reconnection rate in these sheets, e.g., on whether
reconnection is fast (e.g., Petschek) or slow (Sweet --
Parker). This is actually a rather subtle point. Indeed,
in a steady state, the energy dissipated in the corona per
unit time should be equal to the power pumped into the
corona magnetically from the solar surface. And it is
not obvious why and how the energy-pumping rate de-
pends on what happens in the corona. For example, if
the corona were enclosed in a fixed volume, then the en-
ergy dissipation per unit volume would be fixed (as in
driven MHD turbulence in a box). However, it is impor-
tant to recognize that the coronal volume is not fixed! If,
for example, reconnection were to suddenly slow down,
then more energy would be pumped in than the corona
could dissipate; then, in order to accommodate this ad-
ditional free magnetic energy, the corona would respond
by increasing its scale-height. That is, coronal magnetic
structures would grow in height until, finally, the total
dissipation in the corona became equal to the total input
from the photosphere. Because of the increased volume,
the magnetic dissipation per unit volume is decreased.
In addition, as coronal magnetic structures grow in
height, especially when H & R⊙, the amount of energy
pumped into the corona by the footpoint motions may go
down. This is because the work done by footpoint mo-
tions is proportional to vfp·BhorBvert/4π. As the coronal
structures grow in height without increasing their lateral
size, the horizontal field, Bhor, decreases, whereas the
vertical field component, Bvert, does not change. Corre-
spondingly, the overall power pumped from the photo-
sphere into the corona goes down.
As follows from equation (10), the reconnection mode
is determined by the global scale L of the reconnection
layer and by the basic physical parameters characteriz-
ing the plasma in the layer (i.e, ne, Te, and B0). The
typical values of L and B0 are determined by the scale
and strength of the magnetic structures emerging from
the Sun and by the scale of footpoint motions (e.g., the
meso-granular scale). Therefore, for the purposes of the
present discussion, I will regard L and B0 as fixed and
then ask what determines the coronal density and tem-
perature.
Following this line of reasoning, let us invert equa-
tion (10) and view it as the condition for the plasma
density. That is, let us introduce a scale-dependent crit-
ical density, nc, below which the reconnection process
transitions from the slow collisional Sweet -- Parker regime
to a fast collisionless regime:
nc(Bz . B0) ∼ 2 · 1010 cm−3 B4/3
1.5 L−1/3
9
,
(24)
where L9 is the global reconnection layer length ex-
pressed in units of 104 km. In the strong-guide-field case,
Bz ≫ B0, the corresponding expression becomes:
nc(Bz ≫ B0) ∼ 2 · 1010 cm−3 B4/3
1.5 L−1/3
9
Bz(cid:19)2/3
(cid:18) B0
.
(25)
It is interesting to note that these values are close to
those observed in active solar corona. I suggest that this
is not a pure coincidence.
The reason for this is that there is an important feed-
back between coronal energy release and the coronal den-
sity. This feedback is due to the fact that the gas high
in the corona actually comes from evaporation from the
surface along the field lines that just underwent recon-
nection.
In fact, it is essential to consider the coronal
heating not only as the process of increasing the coro-
nal plasma's temperature, but also as that of increas-
ing its density (e.g., Klimchuk 2006; Aschwanden et al.
2007).
Indeed, a part of the energy released in a re-
connection event is rapidly conducted along the field by
fast (perhaps, nonthermal) electrons to the surface and
is deposited in a denser photospheric and chromospheric
plasma. This leads to a localized heating at the foot-
points of the post-reconnected magnetic loops and to
a subsequent chromospheric evaporation along them, a
well-documented phenomenon in solar corona (e.g., As-
chwanden et al. 2001a,b and Klimchuk 2006; see, how-
ever, Brown et al. 2000 and Aschwanden et al. 2007
for an evidence that chromospheric evaporation occurs
in response to chromospheric, rather than coronal, heat-
ing events). As a result, these loops become filled with a
dense and hot plasma. The density rises and may now ex-
ceed nc. This will shut off any further reconnection (and
hence heating) involving these post-reconnective loops
until they again cool down, which occurs on a longer,
radiative timescale.
Let us consider a couple of examples of how this might
work. First, let us consider the case when the plasma
density is relatively high and so the radiative cooling
time, τrad, is relatively short compared with the timescale
on which the coronal magnetic structures change due to
footpoint motions, τfp; however, it is of course still long
compared with the fast reconnection timescale. The op-
posite limiting case will be discussed later in this section.
Let us suppose that due to field-line twisting, a recon-
10
D. A. Uzdensky
necting structure is set up in the corona at t = 0. Let this
structure be characterized by the current sheet length L
and the reconnecting field component B0. Furthermore,
let us further suppose that, initially, the ambient density
of the background plasma is higher than nc(L, B0). Then
the reconnection layer is collisional and reconnection pro-
ceeds very slowly, in the Sweet -- Parker mode. That is,
there is almost no reconnection at all and hence coronal
heating is weak. The surrounding plasma gradually cools
by radiation (and also, in general, by thermal conduc-
tion) and the pressure scale height gradually goes down.
The gas gradually precipitates to the surface. Then the
density of the plasma entering the layer decreases and at
some point becomes lower than the critical density. The
reconnection process then suddenly switches to the col-
lisionless regime. Petschek-like fast reconnection ensues,
and the rate of magnetic energy dissipation greatly in-
creases. A flare commences. Some fraction of the energy
released by reconnection is transported by the electron
conduction along the reconnected field lines down to the
base, where it is deposited in the much denser photo-
spheric or chromospheric plasma. This in turn leads to
a massive evaporation along the same field lines. As a
result, the newly-reconnected loops are now populated
with relatively dense and hot plasma. They cool down
only slowly via conduction and radiation losses, keeping
their relatively high density for an appreciable length of
time. If, during this time, these loops become twisted or
somehow get in contact with other loops, they are now
not likely to reconnect rapidly, since their plasma density
is above critical. This inhibits further coronal heating in
the given region. In fact, we can speculate that for any
further outbursts of coronal activity in the given region
to occur, one has to wait for the gas in post-reconnective
loops to cool down significantly, which occurs on a longer,
radiative timescale.
Thus we see that, although highly intermittent and in-
homogeneous, the corona is working to keep itself roughly
at about the height-dependent critical density given by
equation (24) or (25). Correspondingly, the background
coronal temperature should be such that results in a
density scale-height that is just large enough to popu-
late the corona up to the critical density level at a given
height. In this sense, coronal heating regulates itself (Uz-
densky 2006, 2007).
An important remark is that loop brightness in soft
x-rays and UV does not necessarily mean that this loop
is actively undergoing fast reconnection at the given mo-
ment of time. Instead, it just reflects the fact that the
loop has a high plasma density (since radiation intensity
is proportional to n2). This implies an episode of strong
evaporation from the surface in the recent past, presum-
ably caused by a preceding fast-reconnection coronal en-
ergy release in this loop. The nearby coronal regions
that are darker have evidently lower density, and they
would be ripe for reconnection from the point of view of
the collisionless reconnection condition. However, they
probably are filled with a magnetic field that does not
have complex topology with current sheets, since the un-
derlying reason for coronal activity is always the genera-
tion of small coronal current sheets as a result of field-line
braiding driven by a complicated pattern of the footpoint
motions on the solar surface.
The above radiative-cooling-dominated scenario was
presented here first only because it brings out more
clearly the role that plasma density plays in controlling
the reconnection regime.
It also highlights the role of
chromospheric evaporation in controlling the plasma den-
sity. However, there is an alternative version of this pic-
ture corresponding to the opposite situation, τfp < τrad.
In this case one can repeat the same arguments as those
given above but instead of the slow initial evolution of
the plasma density due to gradual radiative cooling, one
can invoke the slow evolution of the reconnecting mag-
netic field strength caused by the motion of the loop foot-
points (as was described by Cassak et al. 2006). In other
words, one can regard the ambient plasma density n as
fixed on the timescale of interest, whereas the reconnect-
ing field component B0 gradually increases (for simplicity
we also keep L fixed). Then, the current sheet gradually
builds up, while staying in the Sweet -- Parker mode, and
then rapidly switches over to the fast collisionless mode
once B0 has exceeded a certain threshold. This scenario
is in fact closer in spirit to the original Parker nanoflare
model (Parker 1983, 1988), and also to the discussion by
Cassak et al. (2005, 2006). Instead of defining nc(L, B0),
one can just as well rewrite the fast reconnection con-
dition in terms of a critical current-sheet strength, i.e.,
the critical reconnecting magnetic field component, Bc,
expressed as a function of L and n:
Bc(n, L) ≃ 20 G L1/4
9 n3/4
10 .
(26)
In the case of a strong guide field, Bz ≫ B0, the critical
reconnecting field is
Bc(n, L, Bz) ≃ 30 G L1/6
9 n1/2
10 B1/3
z,2 .
(27)
where Bz,2 ≡ Bz/(100 G).
As I will show below,
in the long run one can ex-
pect some sort of statistical steady state with τfp ∼ τrad.
Then, this second scenario is just as likely to take place
as the first one. In fact, in reality they probably both
occur alongside each other.
I would like to emphasize that the phenomenon de-
scribed in the first part of this paper (i.e., the transition
from slow to fast collisionless reconnection) is not the
whole coronal heating story, although, in my view, it is
an important part of it. Conceptually, it is probably
about one quarter of the whole story. The other impor-
tant pieces of the puzzle are: (i) complex dynamic inter-
action of a large number of coronal magnetic loops, ulti-
mately responsible for the generation of current sheets
in the corona by footpoint motions; this requires the
development of an appropriate statistical description of
loops (e.g., Uzdensky & Goodman 2007, in preparation);
(ii) chromospheric evaporation following a reconnection
event and motions of plasma along the loop; and (iii) ra-
diative (and thermal-conduction) energy losses and the
resulting precipitation of the material back to the so-
lar surface. All these essential physics ingredients (or
at least some effective representation of them) will even-
tually have to be incorporated into a single comprehen-
sive theoretical (probably numerical) model of the solar
corona.
It is important to note that each of these processes
comes with its own characteristic timescale. The recon-
nection time in the fast regime, τrec, is very short. For the
Solar Coronal Heating
11
purposes of our discussion, we shall regard it as instanta-
neous. The characteristic time for material evaporation
following a nanoflare is also relatively short. The two
longer time-scales in the problem are the characteristic
timescale for generation and growth of new current sheets
by photospheric footpoint motions, τfp, and the radiative
cooling timescale τrad. As we shall show below, the sys-
tem tends to adjust itself to a statistical equilibrium in
which these two timescales are comparable (of order one
hour in the solar corona). Finally, the longest time scale
of all is the Sweet -- Parker reconnection time. It is much
longer than one hour, and is therefore irrelevant in our
problem. That is, the energy release during the Sweet --
Parker phase is not important. However, this does not
mean that the existence of the slow Sweet -- Parker regime
is not important; it is in fact extremely important, since
it allows for a strong current sheet to build up, accumu-
lating a large amount of free magnetic energy.
This discussion brings us to one of the most interesting
issues in solar coronal heating -- intermittency. The key
questions here are: what is the distribution of (spatial,
temporal, and energy) scales on which magnetic dissipa-
tion occurs? what scales are responsible for most of the
energy release? In my opinion, the collisionless reconnec-
tion condition is, at the end of the day, very important
for answering these questions. In particular, I think this
condition plays a key role in establishing the spectrum
of the energy release events, e.g., the observed power-law
distribution of flare energies.
3.2. Long-term Evolution of a Coronal Loop
To illustrate the above points, let us consider the fol-
lowing model problem. One can raise the question of
whether the self-regulating behavior of the nonlinear dy-
namical system described above can exhibit an oscilla-
tory behavior (Parker 2006, private communication), es-
pecially in light of a hysteresis-like phenomenon reported
by Cassak et al. (2005). Of course, in general this issue
is difficult to assess within the present model, since it
first needs to be incorporated in the overall multi-loop
dynamics (e.g., Uzdensky & Goodman 2007, in prepara-
tion); that is, the energy release in a given fast recon-
nection event does not, strictly speaking, directly affect
the reconnection process that produced it. However, it
does lead both to relaxation of the magnetic stress and
also, importantly, to the increase in density in the post-
reconnected flux tube (or more precisely, in the elemen-
tary magnetic thread within the larger flux tube). This
will in turn temporarily inhibit any reconnection involv-
ing this given field thread even if a new current sheet
is formed. This inhibition will last for some time un-
til either the plasma density goes down because of the
radiative cooling or the reconnecting field component in
the new current sheet grows to a larger strength.
Let us consider a simple illustrative example that
demonstrates this oscillatory behavior and the secular
evolution of the system on a longer timescale. Consider
a composite coronal flux tube, initially filled with a rela-
tively low plasma density n. For simplicity, let us imagine
that that this tube consists of two separate elementary
flux threads that are being wrapped around each other
steadily by the sheared photospheric motions. Conse-
quently, a current sheet gradually builds up inside the
tube, in the spirit of the nanoflare model. Again for
simplicity, let us assume that the length L of this cur-
rent sheet and the axial magnetic field strength Bz in
the tube stay constant at all times. Then, as the ele-
mental flux threads are wrapped around each other, the
reconnecting field component, B0, gradually increases.
Because the plasma density is low, the radiative cooling
time is relatively long and therefore we can regard the
plasma density constant during this phase. On the other
hand, also because the density is low, the critical field
strength is also low, according to equation (27). That
is, the current sheet will become collisionless relatively
early, at some small value of B0. Correspondingly, the
amount of free magnetic energy attributed to the current
sheet at that moment will also be small and the resulting
reconnection event will be relatively weak. After the re-
connection event, the magnetic field becomes unwrapped
again (almost completely, since the hysteresis behavior
reported by Cassak et al. 2005, 2007 implies that fast
reconnection proceeds to the end); then the footpoint-
driven process of current-sheet build-up starts all over
again. Now, if the plasma density in the tube under
consideration were fixed at a constant small value ne,
then (since L and Bz are also kept fixed) all the sub-
sequent fast reconnection events would be triggered at
the same critical value of B0 = Bc(n, L, Bz). Corre-
spondingly, they would all be characterized by a rel-
atively small amount of energy released, on the order
of V B2
c (ne, L, Bz)/8π, where V is the volume of the re-
gion involved in the given reconnection event. At the
same time, since weaker current sheets are easier to pro-
duce, there would be following each other at a relatively
rapid rate.
However, because of the mass exchange between the
surface of the Sun and the corona, the ambient density
is not fixed a priori.
In particular, if one starts with
a small initial density, the first few reconnection events
will be energetically small, but, nevertheless, they will
eventually lead to an increase in the plasma density via
chromospheric evaporation. This will make the critical
magnetic field stronger and hence the amount of free
energy stored between reconnection events and released
through them larger. At the same time, such events will
be more rare, since it takes longer to build up a stronger
current sheet. Therefore, an increase in plasma density
shifts the dominant energy-release scale towards larger
energies. Thus, in this picture, the plasma density in
the tube increased over time in a step-like manner:
it
increases by a certain amount in the aftermath of each
reconnection event and stays roughly flat between them.
The amount of plasma pumped into the corona as a re-
sult of each event is roughly speaking proportional to the
magnetic energy released in this event. The latter is in
turn a monotonically-increasing function of density, ac-
cording to equation (27). That is, we have, keeping L
and Bz fixed: δn ∼ B2
c (n) ∼ n. Thus, the relative in-
crease is then independent of density; it can be roughly
estimated as follows. Consider a loop of constant cross-
sectional area L2 and length Lk > L; the loop's volume
is then Vloop ∼ Lk L2. Assuming that the reconnection
layer extends all the way along the loop, the volume V
of the region involved in the reconnection event is sim-
ilar: V ∼ Vloop ∼ Lk L2. Then, the magnetic energy
released in the event is of order Erec ∼ (B2
0 /8π) Lk L2.
12
D. A. Uzdensky
Next, let us say that a certain fraction ǫ < 1 of this
energy goes towards filling the entire loop with plasma
evaporated from the surface. The coefficient ǫ depends
on the portion of the released energy that goes to the
energetic electrons and is transported by them to the
photosphere, times the fraction of this deposited energy
that goes to material evaporation (as opposed to that im-
mediately radiated away from the footpoints). Overall, ǫ
describes the fraction of the flare energy that eventually
will be radiated by post-coronal loops as soft X-rays, UV,
etc., but not hard X-rays. Immediately after the evapo-
ration episode, the energy ǫErec is divided between the
increase in the gravitational energy, δEgrav, and in the
thermal energy, δEth ≃ 3 δn kBTloopVloop, of the coronal-
loop plasma (for simplicity of discussion, I dropped the
contribution to δEth associated with the increase in the
gas temperature; in reality, of course, both n and Tloop
should increase together).
If the height of the loop is
larger than the density scale-height, then these two con-
tributions are likely to be comparable, according to the
virial theorem. However, in the case of a compact loop,
whose height is less than the thermal scale-height, the
thermal energy increase will be larger than δEgrav.
In
either case, we expect δEth ∼ ǫErec, from which we im-
mediately find
δn
n
∼ ǫ
B2
c
8π
1
3nkBTloop
=
2ǫ
3 βloop
,
(28)
where βloop ≡ 16π nkBTloop/B2
is the characteristic
c
plasma-beta based on the critical reconnecting field.
Substituting the expression (27) for Bc, we get
c,1.5 ∼ 0.07 Tloop,6 B−2/3
βloop ≃ 0.07 n10 Tloop,6 B−2
z,2 L−1/3
,
(29)
9
and so
δn
n
∼ 10 ǫ T −1
loop,6 B2/3
z,2 L1/3
9
.
(30)
There are also numerical factors of order unity which we
have ignored. It is reasonable to expect that this ratio
should be roughly of order unity. For now let us just
assume that it is not large, i.e., δn/n . 1.
As mentioned above, in a realistic scenario the loop
temperature should also increase at each step, in concert
with the density. This means that, to get into the coronal
loop, chromospheric gas would need to be heated to a
temperature that is higher than at the previous step.
Then, each erg of the released energy would be able to lift
a smaller amount of gas, i.e., the evaporation efficiency of
coronal heating would decrease over time. To illustrate
this point, let us, for example, adopt the famous RTV
scaling for the loop temperature (RTV78), even though it
was derived assuming steady heating and thus is not valid
for the impulsive situation considered here (see below
for more discussion). Recasting the RTV78 scaling in
terms of the electron density instead of the pressure, we
write T RTV
k,9 . Substituting this into
equation (30), we get
loop ≃ 3 · 106 K n1/2
10 L1/2
δn
n
∼ 3 ǫ n−1/2
10 B2/3
z,2 L1/3
9 L−1/2
k,9 ∼ n−1/2 .
(31)
Next, the time separation between the steps also in-
creases with the increased threshold magnetic field. In
particular, let us assume that the footpoint driving twists
up the field lines and thus creates and then strength-
ens current sheets (or perhaps, quasi-separatrix lay-
ers, see, e.g., Galsgaard & Nordlund 1996; Titov &
Hornig 2002) at some constant rate. This is, of course,
an over-simplification.
In reality, current sheets may
be created as a consequence of the nonlinear develop-
ment of kink-like instabilities of twisted magnetic loops
(Dahlburg et al. 2005). Then, the growth of B0 with
time is not likely to be a smooth function, but instead
may involve sudden jumps. Here, however, we would like
to avoid such complications. Thus, let us consider, as an
illustration, a simple model in which the reconnecting
field component increases linearly between reconnection
events,
dB0
dt
= γBz ,
(32)
where γ = const. Then, using equation (27), the time δt
it takes the reconnecting field to grow a given critical
value Bc(n) (after a complete relaxation during the pre-
ceding reconnection event) is δt = γ−1Bc(n)/Bz ∼ n1/2.
Therefore, as long as the relative increase in density at
each step is not large, the long-term (t ≫ δt) evolution
of the system can approximately be described by a dif-
ferential equation:
dn
dt
∼
δn(n)
δt(n)
∼ const ,
(33)
corresponding to a linear rise n(t) ∼ t. Accordingly, the
emission measure of the loop should increase as t2. [Al-
ternatively, if we neglect the evolution of the loop tem-
perature in the above analysis and instead keep Tloop =
constant, we then get δn ∼ n and hence n(t) ∼ t2, with
the emission measure rising as t4.]
This growth will continue until one of the following two
processes will intervene. First, it may happen, as the
density builds up, the critical value of B0 will become so
large (e.g., a sizable fraction of Bz), that the equilibrium
shape of the entire loop will be affected. For example,
the loop may become external-kink-unstable and become
helical (sigmoidal loop), which with further twisting may
result in a large-scale disruption with a catastrophic en-
ergy release, a large flare. This scenario, by the way, sug-
gests that coronal loops should gradually (on the time
scale of hours) brighten up before a major disruption,
because a higher plasma density (manifested as a higher
emission measure) leads to a higher critical field strength
and hence allows a larger amount of free magnetic energy
to be accumulated without being prematurely dissipated
via smaller reconnection events.
On the other hand, there is another, less violent out-
come that is also possible. Indeed, so far we have been ne-
glecting radiative cooling. However, as the plasma den-
sity gradually builds up due to a series of chromospheric
evaporations caused by reconnection events, the radia-
tive cooling time decreases (as n−1). At the same time,
as we have seen above, the time interval between subse-
quent reconnections becomes longer and longer. Eventu-
ally, a point may be reached when the amount of plasma
precipitated during δt will become equal to the amount
of material injected into the loop during an evaporation
episode. After that point, there will be no further net
Solar Coronal Heating
13
secular change in the coronal density; the system will
attain a stable limit cycle behavior. To evaluate the ex-
act conditions characterizing this state one will need to
calculate the amount of material evaporated following a
given small flare and the rate at which the plasma returns
to the photosphere as a result of gradual cooling. This
would require a detailed model of the thermal structure
along loop (e.g., along the lines of RTV78), but this can
(and should!) definitely be done. For now, however, we
just want to get some qualitative feeling. Therefore, we
shall just say, tentatively, that the system enters this sta-
ble cyclic regime (with no secular density gain) when the
radiative cooling time becomes equal to the time between
reconnection events. This yields the following condition:
δt(n) = γ−1 Bc(n)
Bz
= τrad(n) .
(34)
This equation can now be viewed as a condition that
determines the long-term equilibrium plasma density in-
side the loop, n∗. Substituting equation (12) for the
radiative cooling time and equation (27) for Bc, we ob-
tain n∗ as a function of the loop temperature Tloop, γ, L,
and Bz:
n∗ ≃ 2 · 1010 cm−3 B4/9
z,2 L−1/9
9
(γτrad,0)2/3(cid:18) Tloop,6
Q−22 (cid:19)2/3
,
(35)
where Q−22 ≡ Q(Tloop)/(10−22 erg cm3 s−1) and where
for convenience we defined τrad,0 as the radiative cool-
ing time corresponding to n10 = 1, Tloop = 1 MK,
and Q−22 = 1: τrad,0 ≃ 400 sec.
Correspondingly, the characteristic threshold recon-
necting magnetic field is
B0,∗ = Bc(n∗, L, Bz)
≃ 45 G L1/9
9 B5/9
z,2 (γτrad,0)1/3(cid:18) Tloop,6
Q−22 (cid:19)1/3
,(36)
and the characteristic time interval between subsequent
reconnection events is
∆t∗ = τrad(n∗)
≃ 200 sec B−4/9
z,2 L1/9
9
(γτrad,0)−2/3(cid:18) Tloop,6
Q−22 (cid:19)1/3
,(37)
The characteristic amount of energy released in each
reconnection event comprising the limit cycle is
E∗ ≃ Lk L2 B2
0,∗
8π
≃ 8 · 1028 erg ×
× Lk,9 L2+2/9
9
B10/9
z,2
(γτrad,0)2/3(cid:18) Tloop,6
Q−22 (cid:19)2/3
,(38)
and hence the time-averaged rate of magnetic dissipation
in the loop is
H∗ =
E∗
∆t∗
≃ 4 · 1026 erg sec−1 ×
× Lk,9 L2+1/9
9
B14/9
z,2
(γτrad,0)4/3(cid:18) Tloop,6
Q−22 (cid:19)1/3
,(39)
Note that most of the above parameters have weak
dependence on L, except for E∗ and H∗, whose strong L-
dependence is almost entirely due to the volume involved.
To make any further progress, we need two more rela-
tionships: (1) the cooling function Q(T ) and (2) a scal-
ing for the loop temperature Tloop. Note that both are
subject to significant uncertainties. In particular, the ra-
diative cooling function exhibits a complicated behavior
in the relevant temperature range (1-10 MK); for exam-
ple, according to Cook et al. (1989), it decreases sharply
between T = 1 MK and T = 3 MK, and then stays
nearly flat for T > 3 MK. Therefore, in principle, one
should not expect simple power-law scalings to result at
all. Furthermore, there are still significant disagreements
in the literature regarding Q(T ) (e.g., Raymond et al.
1976; Raymond & Smith 1977; RTV78; Cook et al. 1989;
Landi & Landini 1999; Aschwanden et al. 2000a). For
these reasons, any realistic quantitative analysis of ra-
diative cooling in the context of the present model is a
highly non-trivial task, adding even more ambiguity to
this already complicated picture. Getting into such a
high level of sophistication and detail would exceed the
overall level of accuracy of our model. We shall therefore
leave it for a future study.
In addition to the cooling function, we need an ex-
pression for the loop temperature, Tloop, which strongly
affects plasma cooling, in terms of the other loop parame-
ters. In principle, the temperature, or, more generally, its
distribution along the loop, is determined by the hydro-
static balance in conjunction with the energy transport
along the loop, which includes thermal conduction, ra-
diative losses, and distributed heating. Such an analysis
has been first performed by RTV78 for the simplest case
of uniform and steady heating; as a result, they derived
the famous RTV scaling for the loop-top temperature:
T RTV
loop = 1.4 · 103(pLk)1/3, where p is the loop pressure
and Lk is its length (which is different and usually larger
than our current-sheet length L; in our model, we shall
regard it as constant). Now, it is true that there is a lot
of disagreement in the modern literature regadring the
applicability of the RTV scaling to the real solar corona
and the observational support is doubtful (e.g., Porter
& Klimchuk 1995; Cargill et al. 1995; Kano & Tsuneta
1995; Aschwanden et al. 2000b,2001a). This is not sur-
prising taking into account that the RTV scaling was de-
rived assuming stationary and uniform heating, and thus
ignored the impulsive and intermittent nature of coronal
energy dissipation (which is an intrinsic property of my
model). In addition, to develop their theory, RTV78 re-
lied on an old cooling function due to Raymond (see be-
low), whereas somewhat different cooling functions have
been used in recent years (Cook et al. 1989; Landi &
Landini 1999; Aschwanden et al. 2000a).
Nevertheless, despite its significant limitations, the
RTV scaling has been highly influential in solar physics.
It is still a highly-respected common standard, against
which solar physicists often measure their theories and
observations. Therefore, just to present an illustrative
example, let us now combine the RTV scaling with our
model. First, to be consistent, we need to adopt the
cooling function due to Raymond (Raymond et al. 1976;
14
D. A. Uzdensky
Raymond & Smith 1977) that was used by RTV78:3
QRTV(T ) ≃ 2 · 10−22 T −2/3
6
erg cm3 s−1 ,
(40)
applicable for 2 -- 10 MK (see Appendix A of RTV78).
Substituting this function into scalings (35) -- (39), we get
9
T 10/9
loop,6 (γτrad,0)2/3 ,
z,2 L−1/9
9 T 5/9
z,2 L1/9
n∗ ≃ 1010 cm−3 B4/9
B0,∗ ≃ 40 G B5/9
z,2 L1/9
loop,6 (γτrad,0)1/3 ,
9 T 5/9
∆t∗ ≃ 160 sec B−4/9
B10/9
E∗ ≃ 5 · 1028 erg Lk,9 L2+2/9
H∗ ≃ 3 · 1026 erg sec−1 Lk,9 L2+1/9
loop,6 (γτrad,0)−2/3 ,
(43)
z,2 T 10/9
loop,6 (γτrad,0)2/3 ,(44)
B14/9
z,2 ×
(41)
(42)
9
9
× T 5/9
loop,6 (γτrad,0)4/3 .
(45)
Next, recasting the RTV loop-temperature scaling in
terms of the electron density instead of the pressure,
loop ≃ 3 · 106 K n1/2
T RTV
10 L1/2
k,9 ,
and substituting it into the above scalings, we get:
(46)
(47)
(48)
(49)
(50)
in order to confirm, modify, or refute various elements of
my model.
First, the whole picture hinges, in part, on the premise
that reconnection is slow in the collisional regime. Al-
though in the above discussion I have for definiteness as-
sumed that it is as slow as in the classical Sweet -- Parker
model, I don't actually think that this needs to be the
case. If the true rate of classical resistive-MHD reconnec-
tion turns out to be much faster than Sweet -- Parker, the
overall picture would still stand, qualitatively, as long
this rate is still much smaller than the collisionless re-
connection rate. The only thing that matters is that
there are two regimes, slow and fast, and the transition
between them is governed by the collisionality of the re-
connection layer.
Nevertheless, from the point of view of my picture,
studies of slow collisional (resistive MHD) reconnection
are just as important as those of fast collisionless recon-
nection. One would like to tighten up the case for slow
reconnection. Thus, I would like to encourage more stud-
ies of resistive reconnection, to really confirm that it is
slow. Thus, it is important to address the following spe-
cific issues, best using numerical simulations:
1) Does the presence of small-scale 3D MHD turbu-
lence enhance reconnection rate, as suggested by Lazar-
ian & Vishniac (1999)? If it does, how large is the en-
hancement and what determines it? Could it be as fast
as collisionless reconnection, in which case it would con-
stitute a challenge to the picture presented in this paper?
To address this issue, one would have to perform a 3D
MHD simulation involving a regular large-scale recon-
nection layer with a super-imposed 3D MHD turbulence.
Just as importantly, one needs to investigate the possibil-
ity that 3D MHD turbulence is spontaneously generated
inside the layer by the secondary-instability mechanism
of Dahlburg et al., (1992, 2005), and evaluate its effect
on the time-averaged reconnection rate.
2) What is the effect of the 2D tearing instability inside
the resistive Sweet -- Parker layer, on the overall, time-
averaged reconnection rate? It is expected that tearing
instability may become important in reconnection layers
that are very long, with aspect ratios L/δ > 100.
It
may then make the reconnection process inherently time-
dependent, bursty. Therefore, in order to investigate this
issue, one would have to perform only a 2D simulation,
but with a very high resolution (corresponding to S >
104) and with a very long duration (to be able to average
over many bursts).
3) The majority of existing numerical studies of
resistive-MHD reconnection use a constant uniform re-
sistivity. It is important to check whether the main re-
sults will be the same if one uses the actual Spitzer re-
sistivity. This is especially important in the context of a
magnetically-dominated environment, such as the solar
corona, where the temperature inside the reconnection
layer may be much higher than the ambient tempera-
ture. As a result, the Spitzer resistivity may vary by a
factor of a hundred across the layer, being much smaller
at its center.
4) One should further pursue numerical studies of mag-
netic reconnection that include both resistive MHD and
Hall effect, along the lines of the recent work by Cas-
sak et al. (2005, 2006, 2007). Of particular interest is the
k,9 γ3/2
∗ ≃ 6 · 1010 cm−3 Bz,2 L−1/4
L5/4
nRTV
−3 ,
9
k,9 γ3/4
z,2 L−1/8
L9/8
loop,∗ ≃ 7 · 106 K B1/2
T RTV
−3 ,
9
k,9 γ3/4
0,∗ ≃ 80 G B5/6
z,2 L1/24
L5/8
BRTV
−3 ,
k,9 γ−1/4
L5/8
∗ ≃ 900 sec B−1/6
z,2 L1/24
−3
B5/3
k,9 L2+1/12
∗ ≃ 2.5 · 1029 erg L9/4
ERTV
z,2 γ3/2
∆tRTV
9
9
9
,
−3 , (51)
H RTV
∗
≃ 3 · 1026 erg sec−1 ×
B11/6
× L13/8
k,9 L2+1/24
9
z,2 γ7/4
−3 ,
(52)
where, in addition, we introduced γ−3 ≡ γ · 103 sec and
used τrad,0 = 400 sec.
A similar exercise could be conducted, for example,
for an entirely different cooling function (Cook et al.
1989; Landi & Landini 1999; Aschwanden et al. 2000a)
and/or for a different scaling law [e.g., that derived by
Cargill et al.
(1995) or, observationally, by Kano &
Tsuneta (1995)].
These estimates give us the characteristic scales of
micro- and nanoflare-like energy-release events in a
bright coronal loop of a fixed length Lk, fixed cross-loop
spatial scale L, and a fixed axial magnetic field Bz, sub-
ject to continuous braiding at a fixed rate γ.
It is of
course understood that this is just an idealized model
set-up and that, in practice, using such typical values
straight-forwardly may not be meaningful because of the
highly inhomogeneous nature of the solar corona.
4. DISCUSSION: WHAT NEEDS TO BE DONE
In this section I will discuss several open questions rel-
evant to the physical picture presented in this paper --
questions that I would like to see answered in the near
future. Correspondingly, I will describe possible studies
(mostly numerical) that I think ought to be performed
3 Note that, to derive their scaling, RTV78 actually used an
approximation of the Raymond cooling function: Q(T ) ∼ T −1/2.
Solar Coronal Heating
15
intermediate regime, when δSP > di, ρi but at the same
time Ωe > νe. The objective of such simulations would
be to observe the transition from slow to fast reconnec-
tion within one simulation run as the upstream plasma
parameters, such as B0 or n, are gradually changed, pass-
ing the critical point. Whereas the results published by
Cassak et al. (2005, 2006, 2007) are already very impor-
tant, the relevant parameter space is large and similar
studies need to be extended to other regimes, most no-
tably to a situation where βupstream ≪ 1. Also, these
results need to be confirmed by other groups, preferably
with a more realistic electron-to-ion mass ratio, a larger
box size, and higher numerical resolution.
5) Future numerical studies of collisional reconnection
should include realistic energy balance taking into ac-
count both Ohmic heating and the electron thermal con-
duction. The goal here would be to calculate the values
of Te and ne at the center of the reconnection region.
Also, one needs to study temperature equilibration be-
tween ions and electrons.
6) Effect of viscosity on resistive-MHD reconnection
needs to be assessed.
7) The effect of a strong guide field on resistive-MHD
reconnection and on the transition to the fast regime
(e.g., Cassak et al. 2007) needs to be investigated in more
detail.
8) Finally, I would like to strongly encourage further
experimental (laboratory) studies of collisional reconnec-
tion, especially in the large-S limit.
In addition to the above questions related to collisional
reconnection, there are several important issues related
to collisionless reconnection: (1) What is its physical na-
ture of anomalous resistivity due to wave-particle interac-
tions and under what conditions is it excited? (2) What
is the structure of the Petschek-like reconnection layer
for a given functional form of anomalous resistivity and
what is the resulting reconnection rate in terms of the
basic physical parameters? Does the anomalous resistiv-
ity spread along the separatrices (Petschek's shocks) or is
it present only in the small central region? If the latter is
the case, then how does the plasma crossing the shocks
gets heated? (3) When does laminar a Hall-effect (or,
in general, two-fluid) reconnection take place? How do
Hall effect and anomalous resistivity co-exist and interact
with each other? (4) How rapid is two-fluid reconnec-
tion? What parameters affect the reconnection rate in
this regime? (5) What is the effect of guide field on trig-
gering and saturation of Hall-regime reconnection and
on anomalous resistivity? (6) Is collisionless reconnec-
tion process bursty and, if so, what is the time-averaged
reconnection rate? (7) What is the overall partitioning of
the released energy between the bulk kinetic energy, elec-
tron and ion thermal energies, and nonthermal particle
acceleration?
Finally, in order to be realistic, future numerical simu-
lations of the solar corona should include all of the follow-
ing (c.f., Miyagoshi & Yokoyama 2003; Klimchuk 2006):
(1) flux emergence processes and random motions of the
field-line footpoints; (2) physically-motivated prescrip-
tion for transition to fast reconnection, such as the one
suggested in this paper; such a prescription would thus
play a role of a subgrid model used in a large-scale MHD
simulation of the corona; (3) mass exchange between the
corona and the solar surface (e.g., chromospheric evap-
oration and plasma precipitation); (4) optically-thin ra-
diative energy losses and thermal conduction (including
the contribution due to nonthermal electrons) along the
magnetic field lines.
5. CONCLUSIONS
Magnetic reconnection research started 50 years ago
in the field of Solar Physics, with the Sweet -- Parker
(Sweet 1958; Parker 1957) model for solar flares, followed
by the Petschek (1964) theory a few years later. These
studies tackled the most fundamental micro-physical as-
pects of the reconnection layer. Over time, however, the
focus of solar reconnection research has shifted away from
local basic physics of reconnection, despite the fact that
reconnection has been confirmed observationally to be
the key process responsible for magnetic energy release
in flares (e.g., Tsuneta 1996, Yokoyama et al. 2001).
Instead, the forefront of research on the fundamental
physics of reconnection has moved in recent years to
other fields, most notably, to Space Physics, where de-
tailed in-situ measurements using Earth-orbiting space-
craft are now available (e.g., Oieroset et al. 2001; Na-
gai et al. 2001; Mozer et al. 2002), and to Laboratory
Plasma Physics, where several dedicated experiments
(Yamada et al. 1997; Egedal et al. 2007; Brown 1999)
have already made fundamental contributions to our un-
derstanding of reconnection. Most importantly, a lot of
progress has been recently made using numerical simula-
tions, again, mostly in the context of the Earth Magne-
tosphere.
The basic paradigm that emerges as a result of all
these studies can be summarized as follows (see § 2.1).
The starting point is the realization that there are in-
deed two reconnection regimes. The first one is a slow
(Sweet-Parker) resistive-MHD regime that is realized in
relatively dense, collisional plasmas. The second one
is a fast (Petschek-like) regime that takes place in col-
lisionless plasmas. The mechanism for the fast colli-
sionless reconnection can be either a locally-enhanced
anomalous resistivity due to micro-instabilities triggered
when the current density exceeds a certain threshold;
or the Hall effect.
In either case, one can formulate
an approximate condition for the transition between the
slow collisional and the fast collisionless regimes.
If
the guide field is not much larger than the reconnect-
ing magnetic field, this condition is δSP < di, where
δSP is the thickness of the Sweet -- Parker reconnection
layer and di
is the collisionless ion skin depth. One
can further rewrite this condition in terms of the clas-
sical electron mean free path λe,mfp inside the layer as
L < Lc ∼ (mi/me)1/2 λe,mfp, where L is the global sys-
tem size (Yamada et al. 2006). Due to the strong tem-
perature dependence of λe,mfp, this form of the condi-
tion highlights the need to estimate the electron temper-
ature Te at the center of the layer. Using considerations
of pressure balance and energy conservation (see § 2.3),
one can express Te in terms of the reconnecting magnetic
field B0 and the background plasma density n. Then, the
collisionless reconnection condition can be recast in terms
of L, B0 and n, see equation (10) in § 2.2. In the case
with a strong guide field Bz ≫ B0, the corresponding
condition is δSP < ρi and this leads to an additional fac-
tor (B0/Bz)2 in the expression for the critical length Lc
[see eq. (23)].
16
D. A. Uzdensky
One of the main driving forces behind this paper is the
author's desire to bring the recently-obtained knowledge
about reconnection back to Solar Physics and to use it
productively to build a better understanding of the solar
corona. Although most of the present discussion is also
relevant to solar flares, in this paper I focus predomi-
nantly on the problem of solar coronal heating (see § 3).
In the context of Parker's (1983, 1988) nanoflare the-
ory of coronal heating, magnetic energy release in the
solar corona takes place in the form of many unresolved
reconnection events (nanoflares). One of the most im-
portant features of this picture is the intermittent char-
acter of energy release. Random footpoint motions lead
to continuous twisting of elementary magnetic strands
around each other, which, in turn, leads to the forma-
tion of many small current sheets in the corona. Current
sheets may form either in finite time, as suggested by
Parker (1983, 1986), or exponentially in time, as was
demonstrated by van Ballegooijen (1986) and later nu-
merically by Galsgaard & Nordlund (1996); for our pur-
poses, it does not matter which one is correct. What
matters is that thin current layers do eventually form.
Free magnetic energy accumulates for a while and is then
suddenly released in distinct fast reconnection events. In
the present paper, I suggest that the transition between
the slow and fast reconnection regimes plays a key role
in determining when a given nanoflare will take off and
how much energy will be released (see also Cassak et al.
2005). My model can thus be regarded as an alterna-
tive to the secondary-instability mechanism proposed by
Dahlburg et al. (2003). Furthermore, I argue that the
fact that the fast reconnection condition involves the am-
bient plasma density is an important part of the story.
The reason for this is that the density in the corona is not
a fixed constant; in a given loop, it constantly changes in
response to radiative cooling and chromospheric evapo-
ration caused by coronal energy-release events. The ba-
sic picture here is the following. A coronal energy re-
lease leads to an increase in density, thus making the
plasma more collisional. This temporarily inhibits fast
reconnection in the given region until the density de-
creases again (on the radiative cooling timescale) to be-
low a certain critical value. At this point fast reconnec-
tion again becomes possible. On the longer time-scale,
a quasi-periodic behavior is established, characterized by
repeated cycles that include fast reconnection events, fol-
lowed by chromospheric evaporation episodes, followed
by relatively long (∼ 1 hour) and steady periods during
which free magnetic energy in the loop builds up and the
plasma gradually cools down. Thus, coronal heating can
be viewed as a self-regulating process that statistically
keeps the density roughly near the critical value for the
fast reconnection transition. In other words, the system
constantly fluctuates around the state of marginal colli-
sionality as defined by the collisionless reconnection con-
dition. In the long-term statistical equilibrium, a balance
is maintained in which the amount of plasma pumped
into a coronal loop as a result of an evaporation episode
is equal to the amount of plasma drained down to the sur-
face during the gradual radiative cooling stage that takes
place between two subsequent fast-reconnection events.
The characteristic equilibrium density, the time interval
between reconnection episodes, and the energy released
in each such episode can be estimated in terms of the
loop's longitudinal magnetic field, its characteristic size,
and the footpoint driving rate (see § 3).
Finally, I believe that the physical framework de-
veloped in this paper should also be applicable to
magnetically-dominated coronae of other astrophysical
objects, such as other stars and accretion disks (Good-
man & Uzdensky 2007).
I am grateful to S. Antiochos, P. Cassak, J. Goodman,
H. Ji, J. Klimchuk, R. Kulsrud, E. Parker, M. Shay, and
M. Yamada for stimulating discussions and encouraging
remarks and to the anonymous referee for useful sugges-
tions.
This work is supported by National Science Founda-
tion Grant No. PHY-0215581 (PFC: Center for Mag-
netic Self-Organization in Laboratory and Astrophysical
Plasmas).
REFERENCES
Aschwanden, M. J, et al. 2000a, ApJ, 535, 1047
Aschwanden, M. J., Nightingale, R. W., & Alexander, D.
2000b, ApJ, 541, 1059
Aschwanden, M. J., Schrijver, C. J., & Alexander, D.
2001a, ApJ, 550, 1036
Aschwanden, M. J., Poland, A. I., & Rabin, D. M. 2001b,
ARA&A, 39, 175
Aschwanden, M. J., Winebarger, A., Tsiklauri, D., Peter,
H. 2007, ApJ, 659, 1673
Bhattacharjee, A., Ma, Z. W., & Wang, X. 1999, JGR,
104, 14543
Bhattacharjee, A., Ma, Z. W., & Wang, X. 2001, Phys.
Plasmas, 8, 1829
Birn, J. et al. 2001, J. Geophys. Res., 106, 3715
Biskamp, D. 1986, Phys. Fluids, 29, 1520.
Biskamp, D., Schwarz, E., & Drake, J. F. 1995, PRL 73,
3850
Biskamp, D., Schwarz, E., & Drake, J. F. 1997, Phys.
Plasmas, 4, 1002
Biskamp, D., & Schwarz, E. 2001, Phys. Plasmas, 8, 4729
Breslau, J. A. & Jardin, S. C. 2003, Phys. Plasmas, 10,
1291
Brown, J. C., Krucker, S., Gudel, M., & Benz. A. O.
2000, A&A, 359, 1185
Brown, M. R. 1999, Phys. Plasmas, 6, 1717
Brown, M. R., Cothran, C. D., & Fung, J. 2006, Phys.
Plasmas, 13, 056503
Bulanov, S. V., Syrovatsky, S. I., & Sakai, J. 1978, JETP
Lett., 28, 177
Cargill, P. J., Mariska, J. T., & Antiochos, S. K. 1995,
ApJ, 439, 1034
Cassak, P., Shay, M., & Drake, J. 2005, Phys. Rev. Lett.,
95, 235002
Cassak, P., Drake, J., & Shay, M. 2006, ApJ, 644, L145
Cassak, P., Drake, J., & Shay, M. 2007, Phys. Plasmas,
14, 054502
Cook, J. W., Cheng, C.-C., Jacobs, V. L., & Antiochos,
S. K. 1989, ApJ, 338, 1176
Coppi, B. & Frieldland, A. B. 1971, ApJ, 169, 379
Coroniti, F. V. & Eviatar, A. 1977, ApJS, 33, 189
Dahlburg, R. B., Antiochos, S. K., & Zang, T. A. 1992,
Phys. Fluids B, 4, 3902
Dahlburg, R. B., Klimchuk, J. A., & Antiochos, S. K.
2005, ApJ, 622, 1191
Solar Coronal Heating
17
Daughton, W., Scudder, J., & Karimabadi, H. 2006,
Phys. Plasmas, 13, 072101
Egedal, J., Fox, W., Katz, N., Porkolab, M., Reim, K.,
& Zhang, E. 2007, Phys. Rev. Lett., 98, 015003
Erkaev, N. V., Semenov, V. S., & Jamitzky, F. 2000,
Phys. Rev. Lett., 84, 1455
Erkaev, N. V., Semenov, V. S., Alexeev, I. V., & Biernat,
H. K. 2001, Phys. Plasmas, 8, 4800
Forbes, T. G. & Priest, E. R. 1983, Solar Phys., 84, 169
Fujimoto, K. 2006, Phys. Plasmas, 13, 072904
Galsgaard, K. & Nordlund, A. 1996, JGR, 101, 13445
Goodman, J. & Uzdensky, D. A. 2007, in preparation
Gudiksen, B. V. & Nordlund, A. 2002, ApJ, 572, L113
Gudiksen, B. V. & Nordlund, A. 2005, ApJ, 618, 1020
Heitsch, F. & Zweibel, E. G. 2003, ApJ, 583, 229
Hendrix, D. L., van Hoven, G., Mikic, Z., & Schnack,
D. D. 1996, ApJ, 470, 1192
Hsu S. C., Carter, T. A., Fiksel, G., Ji, H., Kulsrud,
R. M., & Yamada, M. 2001, Phys. Plasmas, 8, 1916
Huba, J. D. & Rudakov, L. I. 2004, Phys. Rev. Lett.,
93, 175003
Hesse, M., Schindler, K., Birn, J., & Kuznetsova, M.
1999, Phys. Plasmas, 6, 1781
Ji, H., Yamada, M., Hsu S., & Kulsrud, R. 1998, Phys.
Rev. Lett., 80, 3256
Ji, H., Terry, S., Yamada, M., Kulsrud, R. Kuritsyn, A.,
& Ren, Y. 2004, Phys. Rev. Lett., 92, 115001
Kano, R. & Tsuneta, S. 1995, ApJ, 454, 934
Karimabadi, H., Daughton, W., & Scudder, J. 2007, Geo-
phys. Res. Lett., 34, L13104
Kim, E.-J., & Diamond, P. H. 2001, Phys. Lett. A, 291,
407
Kleva, R. G., Drake, J. F., & Waelbroeck F. L., 1995,
Phys. Plasmas, 2, 23
Klimchuk, J. A. 2006, Solar Phys., 234, 41
Kulsrud, R. M. 2001, Earth, Planets and Space, 53, 417
Kulsrud, R. M. 2005, "Plasma Physics for Astrophysics",
Princeton Univ. Press, Princeton
Kulsrud, R., Ji, H., Fox, W., & Yamada, M. 2005, Phys.
Plasmas, 12, 082301
Kuritsyn, A., Yamada, M., Gerhardt, S., Ji, H., Kulsrud,
R., & Ren, Y., 2006, Phys. Plasmas, 13, 055703
Landi, E. & Landini, M. 1999, A&A, 347, 401
Lee, L. C. & Fu, Z. F. 1986, J. Geophys. Res., 91, 6807
Lottermoser, R.-F. & Scholer, M. 1997, JGR, 102, 4875
Loureiro, N. F., Schekochihin, A. A., & Cowley, S. C.
2007, submitted to PRL; preprint (astro-ph/0703631)
Lyutikov, M. & Uzdensky, D. 2003, ApJ, 589, 893
Ma, Z. W. & Bhattacharjee, A. 1996, Geophys. Res.
Lett., 23, 1673
Malyshkin, L. M., Linde, T., & Kulsrud, R. M. 2005,
Phys. Plasmas, 12, 102902
Masuda, S., Kosugi, T., Hara, H., Tsuneta, S., &
Ogawara, Y. 1994, Nature, 371, 495
Masuda, S., Kosugi, T., Hara, H., Sakao, T., Shibata,
K., & Tsuneta, S. 1995, PASJ, 47, 677
Miyagoshi, T. & Yokoyama, T. 2003, ApJ, 593, L133
Mozer F. S., Bale, S. D., & Phan, T. D. 2002, Phys. Rev.
Lett., 89, 015002
Nagai, T., Shinohara, I., Fujimoto, M., Hoshino, M.,
Saito, Y., Machida, S., & Mukai, T. 2001, JGR, 106,
25929
Oieroset, M., Phan, T.D., Fujimoto, M., Lin, R.P., &
Lepping, R.P. 2001, Nature, 412, 414
Parker, E. N. 1957, J. Geophys. Res., 62, 509
Parker, E. N. 1972, ApJ, 174, 499
Parker, E. N. 1983, ApJ, 264, 642
Parker, E. N. 1988, ApJ, 330, 474
Petschek, H. E. 1964, AAS-NASA Symposium on Solar
Flares, (National Aeronautics and Space Administration,
Washington, DC, 1964), NASA SP50, 425.
Porter, L. & Klimchuk, J. 1995, ApJ, 454, 499
Priest, E. 1984, Solar Magnetohydrodynamics, (Reidel,
Dordrecht, 1984)
Raymond, J. C., Cox, D. P., & Smith, B. W. 1976, ApJ,
204, 290
Raymond, J. C. & Smith, B. W. 1977, ApJS, 35, 419
Ren, Y., Yamada, M., Gerhardt, S., Ji, H., Kulsrud, R.,
& Kuritsyn, A., 2005, Phys. Rev. Lett., 95, 055003
Rogers, B. N., Denton, R. E., Drake J. F., & Shay, M. A.
2001, Phys. Rev. Lett. 87, 195004
Rosner, R., Tucker, W. H., & Vaiana, G. S. 1978, ApJ,
220, 643 (RTV78)
Sato, T. & Hayashi, T. 1979, Phys. Fluids, 22, 1189
Scholer, M. 1989, J. Geophys. Res., 94, 8805.
Shay, M. A. & Drake, J. F. 1998, JGR, 103, 9165
Shay, M. A., Drake, J. F., Rogers, B. N., & Denton, R. E.
1999, Geophys. Res. Lett., 26, 2163
Shay, M. A., Drake, J. F., Rogers, B. N., & Denton, R. E.
2001, JGR, 106, 3759
Shay, M. A., Drake, J. F., & Swisdak, M. 2007, submitted
to PRL; preprint (arXiv:0704.0818)
Shibata, K., Masuda, S., Shimojo, M., Hara, H.,
Yokoyama, T., Tsuneta, S., Kosugi, T., Ogawara, Y.
1995, ApJ, 451, L83
Shibata, K. 1996, Adv. Space Res., 17(4/5), 9
Smith, P. F. & Priest, E. R. 1972, ApJ, 176, 487
Strauss, H. R., 1988, ApJ, 326, 412
Sweet, P. A. 1958, in IAU Symp. 6, Electromagnetic
Phenomena in Cosmical Physics, ed. B. Lehnert, (Cam-
bridge: Cambridge Univ. Press), 123.
Thompson, C. 1994, MNRAS, 270, 480
Titov, V. S. & Hornig, G. 2002, Adv. Space Res., 29,
1087
Trintchouk, F., Yamada, M., Ji, H., Kulsrud, R. M. &
Carter, T. A. 2003, Phys. Plasmas 10, 319
Tsuneta, S., Takakura, T., Nitta, N., Makishima, K.,
Murakami, T., Oda, M., Ogawara, Y., Kondo, I., Ohki,
K., & Tanaka, K. 1984, ApJ, 280, 887
Tsuneta, S., Hara, H., Shimizu, T., Acton, L. W., Strong,
K. T., Hudson, H. S., & Ogawara, Y. 1992, PASJ, 44, L63
Tsuneta, S. 1996, ApJ, 456, 840
Ugai, M., & Tsuda, T. 1977, J. Plasma Phys., 17, 337.
Ugai, M. 1986, Phys. Fluids, 29, 3659
Ugai, M. 1992, Phys. Fluids B, 4, 2953
Ugai, M. 1999, Phys. Plasmas, 6, 1522
Uzdensky, D. A., & Kulsrud, R. M. 1998, Phys. Plasmas,
5, 3249
Uzdensky, D. A., & Kulsrud, R. M. 2000, Phys. Plasmas,
7, 4018.
Uzdensky, D. A. 2003, ApJ, 587, 450
Uzdensky, D. A. & Kulsrud, R. M. 2006, Phys. Plasmas,
13, 062305
Uzdensky, D. A. 2006, ArXiv Astrophysics e-prints,
astro-ph/0607656
Uzdensky, D. A. 2007, Mem. Soc. Ast. It., 78, 317; ArXiv
Astrophysics e-prints: astro-ph/0702699
Uzdensky, D. A. & MacFadyen, A. I. 2006, ApJ, 647,
18
D. A. Uzdensky
1192
van Ballegooijen, A. A. 1986, ApJ, 311, 1001
Yamada, M., Ji, H., Hsu, S., Carter, T., Kulsrud, R.,
Bretz, N., Jobes, F., Ono, Y., & Perkins, F. 1997, Phys.
Plasmas, 4, 1936
Yamada, M., et al. 2006, Phys. Plasmas, 13, 052119
Yokoyama, T., Akita, K., Morimoto, T., Inoue, K., &
Newmark, J. ApJ, 546, L69
|
0807.0460 | 1 | 0807 | 2008-07-02T21:58:22 | Gravitational fragmentation and the formation of brown dwarfs in stellar clusters | [
"astro-ph"
] | We investigate the formation of brown dwarfs and very low-mass stars through the gravitational fragmentation of infalling gas into stellar clusters. The gravitational potential of a forming stellar cluster provides the focus that attracts gas from the surrounding molecular cloud. Structures present in the gas grow, forming filaments flowing into the cluster centre. These filaments attain high gas densities due to the combination of the cluster potential and local self-gravity. The resultant Jeans masses are low, allowing the formation of very low-mass fragments. The tidal shear and high velocity dispersion present in the cluster preclude any subsequent accretion thus resulting in the formation of brown dwarfs or very low-mass stars. Ejections are not required as the brown dwarfs enter the cluster with high relative velocities, suggesting that their disc and binary properties should be similar to that of low-mass stars. This mechanism requires the presence of a strong gravitational potential due to the stellar cluster implying that brown dwarf formation should be more frequent in stellar clusters than in distributed populations of young stars. Brown dwarfs formed in isolation would require another formation mechanism such as due to turbulent fragmentation. | astro-ph | astro-ph |
Mon. Not. R. Astron. Soc. 000, 000–000 (0000)
Printed 27 December 2017
(MN LaTEX style file v2.2)
Gravitational fragmentation and the formation of brown
dwarfs in stellar clusters
Ian A. Bonnell1⋆, Paul Clark2 and Matthew R. Bate3
1 SUPA, School of Physics and Astronomy, University of St Andrews, North Haugh, St Andrews, Fife, KY16 9SS.
2 Institut fuer Theoretische Astrophysik, Albert-Ueberle-Str. 2, 69120, Heidelberg, Germany
3 School of Physics, University of Exeter, Stocker Road, Exeter, EX4 4QL
27 December 2017
ABSTRACT
We investigate the formation of brown dwarfs and very low-mass stars through the
gravitational fragmentation of infalling gas into stellar clusters. The gravitational po-
tential of a forming stellar cluster provides the focus that attracts gas from the sur-
rounding molecular cloud. Structures present in the gas grow, forming filaments flowing
into the cluster centre. These filaments attain high gas densities due to the combina-
tion of the cluster potential and local self-gravity. The resultant Jeans masses are low,
allowing the formation of very low-mass fragments. The tidal shear and high velocity
dispersion present in the cluster preclude any subsequent accretion thus resulting in
the formation of brown dwarfs or very low-mass stars. Ejections are not required as
the brown dwarfs enter the cluster with high relative velocities, suggesting that their
disc and binary properties should be similar to that of low-mass stars. This mecha-
nism requires the presence of a strong gravitational potential due to the stellar cluster
implying that brown dwarf formation should be more frequent in stellar clusters than
in distributed populations of young stars. Brown dwarfs formed in isolation would
require another formation mechanism such as due to turbulent fragmentation.
Key words: stars: formation – stars: luminosity function, mass function – globular
clusters and associations: general.
1
INTRODUCTION
Brown dwarfs, having masses less than 0.08 M⊙, are seen to
be nearly as frequent as stars and to have many of the same
properties during their youth such as circumstellar discs,
binary companions, and chromospheric activity (Burgasser
et al. 2007; Luhman et al. 2007; Scholz, Jawyawardhana &
Wood 2006). There have been several proposed mechanisms
to explain the origin of brown dwarfs (Whitworth et al.
2007). These have involved such diverse physical processes
as turbulent fragmentation, disc fragmentation, the ejection
of stellar embryos and the photo-evaporation of a collapsing
prestellar core. These different mechanisms all rely in form-
ing a hence high gas density in the prestellar cores and thus
a low Jeans mass, and/or in the halting of accretion once a
low-mass fragment has formed.
Turbulent fragmentation (Padoan & Nordlund 2002)
envisions that strong, magnetic shocks produce high den-
sity post-shock gas which will have low Jeans masses and
therefore form brown dwarfs directly from the turbulence
⋆ E-mail: [email protected]
(Padoan & Nordlund 2004). One of the potential difficulties
with this model is that a straight shock only compresses the
gas in 1-D. This does not affect the overall gravitational ra-
dius of the gas and thus has a negligible effect of the Jeans
mass as the gravitational and thermal energies are basically
unchanged (Elmegreen & Elmegreen 1978; Lubow & Pringle
1993; Clarke 1999). Several roughly coincident shocks are re-
quired in order to get 3-D compression and a reduction in
the Jeans mass. The turbulent compression model also ne-
glects any residual internal turbulence generated from the
shocks (Clark & Bonnell 2005).
Disc fragmentation (Bate, Bonnell & Bromm 2002;
Whitworth & Stamatellos 2006; Goodwin & Whitworth
2007; Stamatellos et al. 2007) occurs when a massive cir-
cumstellar disc is unstable to gravitational fragmentation,
potentially induced by a stellar fly-by (but see Clarke et al.
2007). The disc provides the high density material such that
the Jeans mass is necessarily low. Post formation accretion
from the massive disc onto the brown dwarf could increase
the mass significantly while forming brown dwarfs in single
systems is more difficult.
Ejection of newly formed fragments in multiple systems
2
I. A. Bonnell et al.
(Reipurth & Clarke 2001; Bate, Bonnell & Bromm 2002;
Bate & Bonnell 2005) can halt any post formation accretion
such that the fragment maintains a low mass. This still re-
quires that the Jeans mass at the point of fragmentation is
of order a brown dwarf mass such as occurs in circumstellar
discs and infalling filaments. One of the potential difficulties
in this mechanism is that the ejection will tend to truncate
discs at radii of order 10 AU, and likewise disrupt any bina-
ries with separations comparable or larger to this.
One final mechanism is the photoevaporation of collaps-
ing cores (Whitworth & Zinnecker 2004). This mechanism
envisions that a collapsing core in the proximity of an O star
will be photo-eroded before it can fully collapse. The outer
layer of the core will be ionised and unbound while adding
a pressure term onto the inner collapsing core ensuring that
a lower-mass object is formed. The primary difficulty with
this mechanism is that it would need to be finely tuned such
that the radiation field is strong enough to have an effect but
not so strong such that it would completely destroy the core.
Furthermore, it requires the presence of an ionising source
of radiation such that it could only explain brown dwarfs in
the presence of O-stars.
In this paper, we present some recent results on the for-
mation of low-mass stars and brown dwarfs due to the pres-
ence of stellar clusters. We show that brown dwarfs form in
an analogous manner to low-mass stars, due to the gravi-
tational fragmentation of high density gas as it infalls into
a stellar cluster. The primary differences with our previous
work (Bate et al. 2002) is that we stress how the low frag-
mentation is reached due to gravitational compression, and
that subsequent accretion is limited due to the high virial
velocities in clusters such that ejections are not needed. Our
calculations are presented in §2 while the primary results
are presented in §3. Section 4 dissects the role of the stel-
lar cluster in forming brown dwarfs while §5 presents some
observational signatures of the process described here.
2 CALCULATIONS
The results presented here are based on a large-scale
Smoothed Particle Hydrodynamics (SPH) simulation of a
cylindrical 104 M⊙ molecular cloud 10 pc in length and 3
pc in cylindrical diameter. We have chosen an elongated
cloud rather than the more standard spherical cloud as most
molecular clouds are non-sperhical and commonly elongated
(e.g. Orion A). This also allows for the physical properties to
be varied along the cloud in a straightforward manner. The
cloud has a linear density gradient along its major axis with
maximum/minimum values, at each end of the cylinder, 33
percent high/lower than the average gas density. The gas
has internal turbulence following a Larson-type P (k) ∼ k−4
power law and is normalised such that the total kinetic en-
ergy balances the total gravitational energy in the cloud.
The density gradient then results in one end of the cloud
being over bound (still super virial) while the other end of
the cloud is unbound.
The cloud is populated with 15.5 million SPH particles
on two levels, providing high resolution in regions of inter-
est. We initially performed a lower resolution run with 5
million SPH particles producing an average mass resolution
of 0.15M⊙. Upon completion of this low resolution simula-
tion, we used three criteria to identify the regions that re-
quired higher resolution. This included the particles which
formed sinks, and those that were accreted onto sinks. It
also included particles which attained sufficiently high den-
sity such that their local Jeans mass was no longer resolved
in the low-resolution run. All of these particles were iden-
tified and from the initial conditions of the low resolution
run, they were spit into 9 particles each to create the initial
conditions for the high resolution simulations. This particle
splitting was performed on the initial conditions to ensure
that the physical quantities of mass, momentum, energy and
the energy spectrum were preserved. Note that the particle
splitting does not introduce finer structure in the turbulent
energy spectrum. This produced a mass resolution for the
regions involved in star formation of 0.0167M⊙, sufficient to
resolve the formation of brown dwarfs, equivalent to a total
number of 4.5 × 107 SPH particles. The equation of state
(below) was specified in order to ensure that the Jeans mass
in the higher resolution run did not descend below this mass
resolution.
This simulation was rerun from the beginning to en-
sure that the particle splitting did not affect the ongoing
evolution. Particle splitting results in a marked increase in
resolution without unmanageable computational costs (Kit-
sionas & Whitworth 2002, 2007). Note however some of the
unsplit particles, which in the low resolution run neither ex-
ceeded their Jeans mass limit nor became involved in the
star formation, did get accreted by the additional stars in
the high resolution run. This is to be expected as their are
now additional locations of star formation not present in the
low resolution run and these additional sinks will necessarily
accrete unsplit particles.
The simulation follows a barotropic equation of state of
the form:
P = kργ
where
γ = 0.75;
γ = 1.0;
γ = 1.4;
γ = 1.0;
ρ ≤ ρ1
ρ1 ≤ ρ ≤ ρ2
ρ2 ≤ ρ ≤ ρ3
ρ ≥ ρ3,
(1)
(2)
and ρ1 = 5.5 × 10−19g cm−3, ρ2 = 5.5 × 10−15g cm−3, ρ3 =
2 × 10−13g cm−3.
The initial cooling part of the equation of state mimics
the effects of line cooling and ensures that the Jeans mass
at the point of fragmentation is appropriate for characteris-
tic stellar mass (Larson 2005; Jappsen et al. 2005; Bonnell,
Clarke & Bate 2006). The γ = 1.0 approximates the effect
of dust cooling (Larson 2005) while the γ = 1.4 mimics the
effects of when the collapsing core is optically thick to IR
radiation, although its location at ρ = 5.5 × 10−15gcm−3, at
lower densities than is typical, is in order to ensure that the
Jeans mass is always fully resolved and that a single self-
gravitating fragment is turned into a sink particle. A higher
critical density for this optically-thick phase where heating
occurs would likely result in an increase in the numbers
of brown dwarfs formed. the physical processes described
would be unchanged. The final isothermal phase of the equa-
tion of state is simply in order to allow sink-particle forma-
tion to occur, which requires a subvirial collapsing fragment.
Star formation in the cloud is modelled through the in-
Formation of brown dwarfs
3
Figure 1. The final distribution of the gas and stars in the simulation is shown in the 8 by 8 pc image. The stars are indicated by
the yellow filled circles while the brown dwarfs are indicated by the filled blue circles with white edges. The brown dwarfs are located
primarily in clustered regions as these provide the necessary physical properties to form low mass objects. The gas column densities are
also plotted from 0.01 (black) to 100 (white) g cm−2.
troduction of sink-particles (Bate, Bonnell & Price 1995).
Sink-particles formation is allowed once the gas density of
a collapsing fragment reaches ρ ≥ 6.8 × 10−14 g cm−3
although the equation of state ensures that this requires
ρ ≥ 2. × 10−13gcm−3. The neighbouring SPH particles need
be within a radius of 1. × 10−3 pc and that fragment must
be subvirial and collapsing. Once created, the sinks accrete
bound gas within 1. × 10−3 pc and all gas that comes within
2.×10−4 pc. The sinks have their mutual gravitational inter-
actions smoothed to 2. × 10−4 pc or 40 au. No interactions
including binary or disc disruptions can occur within this
radius.
3 BROWN DWARF FORMATION IN
TURBULENT MOLECULAR CLOUDS
The simulation was followed for 1.02 free-fall times or
≈ 6.6 × 105 years and ≈ 3.9 × 105 years after the first
stars formed. During this time, 2542 stars were formed with
masses between 0.017 and 30 M⊙. Of these, ≈ 23 per cent
have masses below 0.08M⊙ but only ≈ 10 per cent (243
objects) have m ≤ 0.08M⊙ and have stopped accreting. A
further 3 per cent are likely to maintain m ≤ 0.08M⊙ given
their final accretion rates and assuming that this accretion
is sustained over the next free-fall time. This gives an ex-
pected final number of brown dwarfs of 342 or 13 per cent
of the stars formed. Figure 1 shows the spatial distribution
of the brown dwarfs at the end of the simulation. They are
primarily in or around forming stellar clusters. We will in-
vestigate the process by which they form in the following
sections.
Most formation mechanisms for brown dwarfs envision
that the physical conditions in the pre-fragmented gas are
such that the Jeans mass, the minimum mass to be gravita-
tionally bound, be of order a brown dwarf mass (Elmegreen
2004; Whitworth et al. 2007; Bonnell et al. 2007). Figure 2
plots the median gas density, and respective Jeans mass,
within 0.05 pc (a typical Jeans length) of where the sink-
particle will form. These values are calculated from the gas
distribution just prior (within 2300 years) to the formation
of the sink-particle. The gas densities and Jeans masses are
plotted against the final masses that these sinks attain by
the end of the simulation. We see that the moderate mass
sinks (0.5 − 2M⊙) form in low density gas where the Jeans
mass is of order 0.1 to almost 1.0 M⊙. In contrast, the lower-
mass sinks form from higher density gas and thus from low
Jeans masses. Note that there is not a perfect one-to-one cor-
respondence between the Jeans masses and the final masses
4
I. A. Bonnell et al.
Figure 2. The gas density (left) and Jeans mass (right) are plotted against the final stellar masses for the 2542 sinks formed. Both the
gas density and Jeans mass are calculated within a radius of 0.05 pc from the incipient sink particle within ≈ 2300 years prior to the
sink formation. The gas densities are higher, and the Jeans masses lower, for the low-mass sinks formed as is expected when the physical
conditions of the gas determine the fragment masses.
even for small masses due to the somewhat arbitrariness of
evaluating the physical conditions within a fixed radius of
0.05 pc. What is important is that the physical properties
in the gas are appropriate to give very low-mass fragments
such that forming brown dwarfs in these regions is natural.
Subsequent accretion can also increase the final masses from
the low value generated at the point of the fragmentation.
This is more evident for the high-mass sinks where the vast
majority of their final masses comes from the subsequent
competitive accretion (Bonnell et al. 2001; Bonnell, Vine &
Bate 2004; Bonnell & Bate 2006).
The first conclusion we can make is that the brown
dwarfs form as lower-mass stars do, from the fragmenta-
tion of a gas cloud where the thermal Jeans mass is of order
the mass of the object formed. The next question is what
drives the gas to such densities that brown dwarfs can form.
One possibility is that turbulent compression leads to the
formation of dense cores and thus brown dwarfs (Padoan
& Nordlund 2004). The spatial distribution of the brown
dwarfs in figure 1 argues against this as the brown dwarfs
are not located randomly throughout the cloud but are lo-
cated in the vicinity of stellar clusters. Furthermore, as SPH
is Lagrangian, we can trace the particles that form individ-
ual sinks backwards in time. We can estimate the fraction of
an individual sink's mass that is brought to the point of for-
mation by the turbulent flows. This shows that the turbulent
flows are responsible for transporting only of order one per
cent of the fragment's mass to within 0.04 pc of where the
sink eventually forms. Instead additional acceleration, such
as occurs when the gas enters the gravitational potential of
the cluster, are necessary. We can therefore conclude that,
in this simulation, turbulence does not lead to the formation
of brown dwarfs.
Figure 3. The mass density of stars within 0.25 pc of the form-
ing sink-particle is plotted as a function of the final mass in the
sink-particle. The moderate and high-mass sinks form in regions
of low stellar mass density whereas the low-mass stars and brown
dwarfs appear to require the presence of a stellar cluster and its
strong gravitational potential in order to form. The gravitational
potential of the cluster generates the high gas densities needed
for forming low-mass objects. The stellar mass densities are cal-
culated within ≈ 2000 years prior to the sink formation. Regions
completely devoid of stars prior to the sink formation are given a
stellar mass density of 1 M⊙ pc−3. Note that although the high-
mass stars form in regions of low stellar mass density, they end
up in the centre of the clusters which form subsequently around
them.
4 BROWN DWARF FORMATION IN
STELLAR CLUSTERS
Gravity provides an alternative to turbulence in generating
the high gas density conditions conducive to forming brown
dwarfs. Gravity has the distinct advantage over turbulence
in that it is intrinsically convergent and therefore better able
to compress gas into small volumes. It also has the ten-
dency of compressing a three-dimensional volume into two-
dimensional sheets, one dimensional filaments and eventu-
ally to point sources (Larson 1985; Bonnell 1999). The stellar
clusters forming in the simulation provide a strong gravita-
tion potential into which flowing gas can be compressed to
much higher densities. Figure 3 shows the stellar mass den-
sity within 0.25 pc, a typical size-scale of a stellar cluster, of
the forming sink-particles just prior to sink formation. We
see that while the moderate and high-mass sinks form in re-
gions where the mass density of stars is low, low-mass stars
and brown dwarfs form in regions of high stellar mass den-
sity. This occurs as the moderate to higher-mass stars form
first while the eventual low-mass stars and brown dwarfs
only form as gas infalls into pre-existing stellar clusters. It
is therefore the presence of the stellar cluster which is driv-
ing the formation of these objects. The higher-mass stars
form the basis for the forming cluster and thus also end up
inside a region of high stellar mass density.
Figure 4 shows the spatial distribution of the gas and
stars within 0.05 pc of a forming brown dwarf. We see a large
filament formed through the amplification of structure in the
infalling gas due to the combination of the cluster potential
and its self-gravity. Several knots of gas are forming along
the filament, some of which go on to form brown dwarfs while
the others collapse to form low-mass stars. Such an occur-
rence of a filamentary structure infalling into the cluster is
relatively common, although often some degree of tangential
motion and thus angular momentum result in the filament
being wound up around the protocluster as it infalls and
forms low-mass objects.
Figure 4 also plots the associated position-velocity di-
agram of the infalling filament centred on the position and
velocity of one of the forming brown dwarfs. The negative
slope shows that the filament is falling into the cluster lo-
cated at ≈ −0.035 pc and ≈ +2 km s−1. The gas is accel-
erated into the cluster such that material at negative (pos-
itive) positions relative to the forming brown dwarf have
negative (positive) relative velocities, respectively. This di-
vergent flow, as well as the large virial velocity the forming
brown dwarf receives due to the cluster potential, restricts
any subsequent accretion and thus maintains the low-mass of
the object. On smaller scales, gravitational collapse reverses
this pattern in the position velocity-diagram. Material at
negative positions has positive relative velocities while ma-
terial at positive positions has negative relative velocities
indicating collapse.
This tidal shearing of the filament acts to limit the mass
of the object. Gas that would normally fall into the forming
fragment is now pulled away by the large-scale potential. As
a result, only the very high gas density fragments, and hence
with low Jeans masses, form in such a shear flow. Figure 5
plots the Jeans mass of the 2542 sinks at the point of forma-
tion as a function of the stellar mass density within 0.05 pc.
We see that the Jeans masses decrease as the cluster stellar
Formation of brown dwarfs
5
Figure 4. In the top panel we show a column density image of
a gaseous filament which is forming brown dwarfs as it accretes
onto a rich cluster. The position (0,0) in the co-ordinates shows
the formation site of one of the new brown dwarfs. The column
density scale runs from 0.5 to 50 g cm−2. The bottom panel
shows a position-velocity diagram for the same region where the
x-positions are centred on the same brown dwarf as in the upper
panel. The velocities are calculated projected along the vector
joining the brown dwarf at position x = 0 to the centre of mass of
the cluster, and are again centred on the brown dwarf. Note that
the colours in the bottom panel scale from 5 to 500 g cm−1 km−1
s. The diagonal form of the filament in position-velocity diagram
reveals that the gas is accelerating into the cluster while being
tidally sheared away from the objects that are forming within.
Relative to the formation site of the brown dwarf, the vast ma-
jority of the filament is moving away: gas at negative positions
have negative relative velocities, while gas a positive positions
have positive relative velocities. Only in the immediate region
surrounding the still-forming brown dwarf can one see a reversal
of this velocity signature, which denotes the gas falling on to the
new object. This feature can be seen in several other points along
the filament, showing the formation sites of other brown dwarfs
or low-mass stars.
6
I. A. Bonnell et al.
Figure 5. The Jeans mass within 0.05 pc of a proto-sink is plotted
against the stellar mass density in the same region for the 2542
sinks formed. Note that the Jeans mass declines with increasing
stellar mass density. This occurs as the gas density is higher in
stellar clusters due to the effects of the gravitational potential and
also that lower gas density fragments would be tidally disrupted
and hence not form any sinks. The solid line indicates the Jeans
mass assuming a gas density that is the same as the stellar mass
density. This tidal limit provides an upper limit on the possible
Jeans masses in the stellar cluster. Both the stellar mass densities
and Jeans masses are calcualted within ≈ 2300 years prior to the
sink formation. The Jeans mass is based on the median gas density
in the region.
mass density increases indicating the prevalence for forming
low-mass objects in dense stellar clusters. Also plotted in
figure 5 is the Jeans mass if the gas density was equal to
the stellar mass density. This is the tidal limit for forming
a fragment and provides an upper limit to the distribution
of Jeans masses of the forming fragments. Higher-mass frag-
ments cannot form as their high Jeans masses and low gas
densities would result in their being tidally disrupted before
they could collapse.
The growth of a stellar cluster, acts to continually de-
crease the Jeans mass and thus the fragmentation mass in
the infalling gas. As more stars form or fall into the cluster,
the increase in the stellar density increases the tidal shear-
ing and thus necessitates higher gas densities in order for
the fragments to be bound. Thus the fragment mass should
decrease with time and with the growth of a cluster. Subse-
quent accretion onto some of these objects will significantly
increase their masses.
5 CLUSTER DYNAMICS AND ACCRETION
In the previous section we saw that the low-mass sinks form
as the gas infalls into an already formed stellar cluster. This
has important implications as to why they remain low-mass
objects of the typical mass of a brown dwarf rather than
accreting from the abundant reservoir of gas in the stellar
Figure 6. The relative velocity of the 2452 sinks at formation
relative to the centre of mass velocity within 0.25 pc is plotted
against the final mass of each sink. The sinks that attain mod-
erate or high masses form with low relative velocities to their
environment while those that remain low-mass objects form with
high velocities relative to the centre of mass of their environments.
This occurs as the sinks that attain moderate or high masses form
earlier and constitute the initial stars in the cluster. The low-mass
sinks form from gas falling into a already formed stellar cluster
and thus have a high relative velocity. This high velocity also
limits any subsequent accretion ensuring they remain low-mass
objects.
cluster. Figure 4 shows the gas being accelerated to high
(negative) velocities as it infalls into the cluster's gravita-
tional potential. Thus any newly formed fragment will enter
the cluster at high velocity and therefore have difficulty in
accreting from the reservoir of gas in the cluster. Figure 6
shows the relative velocity of each forming sink compared to
its environment within 0.25 pc. The sinks that remain low-
mass have high relative velocity, of order several km s−1 at
the point of formation. In contrast, the sinks that ultimately
attain moderate and high mass form with low relative ve-
locities. This difference occurs as the low-mass sinks form
from gas infalling into an already formed cluster whereas
the sinks that attain high mass form before any stellar clus-
ter is present. They are the first stars around which the
cluster is built and due to their low relative velocity, they
are able to accrete significant amounts of gas and become
high-mass objects. The low-mass sinks cannot accrete signif-
icantly due to their high relative velocity and thus remain
low-mass objects. Estimated accretion rates for a low mass
star (m ≤ 0.1M⊙) travelling at a high velocity (v ≥ 2 km
s−1) in a gas reservoir of 10−17 g cm−3 is ≈ 5 × 10−8 M⊙
yr−1, which is too small to significantly alter the star's final
mass over accretion times of tacc <
∼ 106 years (c.f. Bonnell &
Bate 2006).
In this scenario, the forming brown dwarfs and low-
mass stars do not require any subsequent interactions or
ejections to terminate their accretion processes. It is the
acceleration due to the cluster potential which ensures that
Formation of brown dwarfs
7
The velocities of the brown dwarfs and other low-mass
objects are also similar to those of the other stars as both
simply reflect the overall gravitational potential. Although
the brown dwarfs form at higher relative velocities than do
the higher mass objects, this is simply due to the presence or
absence of a well defined gravitational potential at the point
of formation. The velocities of the higher mass objects are
initially low as they form before the cluster develops. Their
velocities increase over time due to the infall of other stars
into the growing cluster. Virial equilibrium ensures that they
all have similar velocity dispersions.
One significant observable feature is that the brown
dwarfs, and for that matter the very low-mass stars, re-
quire a gravitational potential to compress the gas to suffi-
ciently high densities to attain low Jeans masses. This means
that they form in regions of high stellar densities as shown
in figure 3. There is some subsequent dynamical evolution
and brown dwarfs, like low-mass stars, are more likely to
be ejected in any dynamical event. Nevertheless, the brown
dwarfs formed are more commonly found in a clustered envi-
ronment. Figure 7 shows the fractional abundance of brown
dwarfs as a function of stellar density. The stellar density is
calculated from the distance to the ten nearest neighbours of
each sink-particle. The abundance of brown dwarfs is plotted
for all sinks that have masses in the range a brown dwarf
masses and also for those that given their final accretion
rates are likely to remain brown dwarfs over the following
free-fall time, ≈ 6.5 × 105 years. We see that the frequency
of brown dwarfs is significantly higher in the clustered than
in the non-clustered regions, with a peak abundance of ≈ 25
per cent at high stellar densities which decreases to <
∼10 per
cent in isolated or regions of low stellar densities.
In general, the brown dwarf mass sinks in the lower den-
sity bins are those that have been ejected from their natal
clusters. This small number is thus likely to evolve with time
from basically zero at the point of formation, and should in-
crease somewhat over our final value as more low-mass clus-
ter members are ejected from the clusters. Nevertheless, we
expect that an observable signature of the process described
here is that the fractional abundance of brown dwarfs should
increase with stellar density, and should be lowest amongst
the isolated population of young stars. It is worth noting
that the fractional abundance of brown dwarfs actually de-
creases in the highest stellar density bins as the stars in the
cores of the stellar clusters are predominantly higher mass
stars due to the ongoing (competitive) accretion there.
7 CONCLUSIONS
We have investigated the formation of brown dwarfs in a
numerical simulation of a self-gravitating turbulent molecu-
lar cloud. We find that the brown dwarfs form, as do low-
mass stars, due to the fragmentation of high-density gas
that arises as it infalls into the gravitational potential of
a stellar cluster. Approximately 23 per cent of the objects
formed have final masses in the brown dwarf mass range al-
though only ≈ 10 per cent have brown dwarf masses and
have stopped mass accretion.
The turbulent velocities present in the cloud do not con-
tribute directly to the formation of the brown-dwarf mass
fragments. Instead, these fragments form through the gravi-
Figure 7. The fractional abundance of brown dwarfs is plot-
ted against their local stellar density based on their tenth clos-
est neighbour. The dashed line is for all brown-dwarf-mass sink-
particles at the end of the simulation while the solid line only con-
siders brown dwarf-mass sink-particles with low accretion rates
such that they would remain brown dwarfs if they continued to
accrete over the following free-fall time (6.5 × 105 years). The
fractional abundance of brown dwarfs is low at low stellar densi-
ties and peaks at high stellar densities inside clusters. The highest
stellar densities also appear to have lower fractional abundances
of brown dwarfs.
they have low accretion rates and remain low-mass objects.
Brown dwarfs and low-mass stars thus form in the same way
from low Jeans mass fragments which subsequently accrete
little mass. If ejections are not required, then the disc and
binary properties of these objects need not be affected and
should form a continuous distribution with slightly higher-
mass stars.
6 OBSERVATIONAL SIGNATURES
The brown dwarfs that form in the scenario described here
form in an analogous manner as do very low-mass stars
(Elmegreen 2004). There is nothing specific to being a brown
dwarf that makes their formation distinct from stars. They
form from gas which has low Jeans masses due to its com-
pression as it enters a stellar cluster, and subsequent accre-
tion is generally low due to the high infall velocity imparted
from the cluster potential. There is no need for close interac-
tions or ejections to ensure their low mass. This implies that
the properties of brown dwarfs will generally resemble those
of low-mass stars. As such, their circumstellar disc and mul-
tiplicity properties should form a continuum with low-mass
stars. This process can occur in any strong potential such
as that of a small-N cluster where the Jeans mass becomes
small and that the virial velocity limits subsequent accretion
(Bonnell et al.1997, 2001; Bate et al.2002; Bonnell & Bate
2006). Ejections can occur but are not fundamental to the
process.
Clarke C. J., 1999, MNRAS, 307, 328
Clarke C. J., Harper-Clark E., Lodato G., 2007, MNRAS, 381,
1543
Elmegreen B. G., 2004, MNRAS, 354, 367
Elmegreen B. G., Elmegreen D. M., 1978, ApJ, 220, 1051
Goodwin S. P., Whitworth A., 2007, A&A, 466, 943
Kitsionas S., Whitworth A. P., 2002, MNRAS, 330, 129
Kitsionas S., Whitworth A. P., 2007, MNRAS, 378, 507
Larson R. B., 1985, MNRAS, 214, 379
Larson R. B., 2005, MNRAS, 359, 211
Lubow S. H., Pringle J. E., 1993, MNRAS, 263, 701
Luhman K. L., Joergens V., Lada C., Muzerolle J., Pascucci I.,
White R., 2007, Protostars and Planets V, eds B. Reipurth
et al., 443
Padoan P., Nordlund A., 2002, ApJ, 576, 870.
Padoan P., Nordlund A., 2004) ApJ, 617, 559.
Reipurth B., Clarke C., 2001, AJ, 122, 432.
Scholz A., Jayawardhana R., Wood K., 2006, ApJ, 645, 1498
Stamatellos D., Hubber D. A., Whitworth A. P., 2007, MNRAS,
382, L30
Whitworth A., Bate M. R., Nordlund A., Reipurth B., Zinnecker
H., 2007, Protostars and Planets V, eds B. Reipurth et al.,
459
Whitworth A. P., Stamatellos D., 2006, A&A, 458, 817
Whitworth A. P., Zinnecker H., 2004, A&A, 427, 299
8
I. A. Bonnell et al.
tational compression of gas as it infalls into a stellar cluster.
This intrinsically 3-D compression produces high gas den-
sities and thus low thermal Jeans masses in the range of
brown dwarfs and low-mass stars. The tidal shear of the
cluster, and the velocity imparted on the fragment from the
cluster potential act to limit any subsequent mass increase
due to accretion. There is no need for any subsequent ejec-
tions to halt the accretion implying that the circumstellar
disc and binary properties of brown dwarfs should form a
continuum with low-mass stars.
Brown dwarfs formed through this mechanism should
be preferentially located in regions of high stellar density.
The fractional abundance of brown dwarfs in stellar clusters
is of order 25 per cent in highly clustered regions whereas
it decrease to of order 10 per cent in isolated regions. This
fraction is likely to increase somewhat due to subsequent
ejections of brown dwarfs and low mass stars from the clus-
ters.
ACKNOWLEDGMENTS
We acknowledge the contribution of the U.K. Astrophysi-
cal Fluids Facility (UKAFF) and SUPA for providing the
computational facilities for the simulations reported here.
P.C.C. acknowledges support by the Deutsche Forschungsge-
meinschaft (DFG) under grant KL 1358/5 and via the Son-
derforschungsbereich (SFB) SFB 439, Galaxien im fruhen
Universum. MRB is grateful for the support of a Philip Lev-
erhulme Prize and a EURYI Award. This work, conducted
as part of the award The formation of stars and planets:
Radiation hydrodynamical and magnetohydrodynami- cal
simulations made under the European Heads of Research
Coun- cils and European Science Foundation EURYI (Eu-
ropean Young Investigator) Awards scheme, was supported
by funds from the Par- ticipating Organizations of EURYI
and the EC Sixth Framework Programme. We would like to
thank Chris Rudge and Richard West at the UK Astrophys-
ical Fluid Facility (UKAFF) for their tireless assistance and
enthusiasm during the completion of this work. Finally, we
thank the referee for some useful comments which helped
clarify the text.
REFERENCES
Bate M. R., Bonnell I. A., 2005, MNRAS, 356,1201
Bate M. R., Bonnell I. A., Bromm V., 2002a, MNRAS, 332, 65
Bate M. R., Bonnell I. A., Price N. M., 1995, MNRAS, 277, 362.
Bonnell I. A., 1999, The physics of star and planet formation, eds.
Lada & Kyfalis, 479
Bonnell I. A., Bate M. R., 2006, MNRAS, 370, 488
Bonnell I. A., Bate M. R., Clarke C. J., & Pringle J. E., 1997,
MNRAS, 85, 201
Bonnell I. A., Bate M. R., Clarke C. J., & Pringle J. E., 2001,
MNRAS, 323, 785
Bonnell I. A., Clarke C. J., Bate M. R., 2006, MNRAS, 368, 1296
Bonnell I.A., Larson R.B., Zinnecker H., 2006, in Protostars and
Planets V, eds B. Reipurth et al., p. 149
Bonnell I. A., Vine S. G., Bate M. R., 2004, MNRAS, 349, 735
Burgasser A. J., Reid I. N., Siegler N., Close L., Allen P.,
Lowrance P., Gizis J., 2007, Protostars and Planets V, eds
B. Reipurth et al., 427
Clark P. C., Bonnell I. A., 2005, MNRAS, 361, 2
|
astro-ph/0311446 | 1 | 0311 | 2003-11-19T10:25:05 | Rapid N_H Changes in NGC 4151 | [
"astro-ph"
] | We have analyzed two long BeppoSAX observations of the bright Seyfert galaxy NGC 4151, searching for short timescale (10-200 ksec) X-ray spectral variability. The light curve of a softness ratio, chosen as most sensitive to pinpoint changes of the column density of the absorbing gas along the line of sight, shows significant variations. We try to model these variations by performing a detailed, time resolved, spectral analysis. We find significant, large (factors of 1.5-6) variations of the absorber column densities on time scales of 40-200 ksec. These values are 10-100 times shorter than those found by Risaliti et al. 2002 in a sample of Seyfert 2 galaxies, and provide strong constraints on the geometry of the obscuring medium. | astro-ph | astro-ph |
RAPID NH CHANGES IN NGC 4151
S. Puccettia b, G. Risalitic, d, F. Fiorea, M. Elvisd, F. Nicastrod, G. C. Perolae, M. Capalbif.
aUniversit´a degli Studi di Roma "Tor Vergata", Roma, Italy
bINAF-Osservatorio Astronomico di Roma, Monteporzio Catone, Italy
cINAF-Osservatorio Astrofisico di Arcetri, Firenze, Italy
dHarvard-Smithsonian Center for Astrophysics, Cambridge, USA
eUniversit´a degli Studi di Roma Tre, Roma, Italy
f ASI Science Data Center, c/o ESA-ESRIN, Frascati, Italy
We have analyzed two long BeppoSAX observations of the bright Seyfert galaxy NGC 4151, searching for short
timescale (∼10-200 ksec) X-ray spectral variability. The light curve of a softness ratio, chosen as most sensitive to
pinpoint changes of the column density of the absorbing gas along the line of sight, shows significant variations.
We try to model these variations by performing a detailed, time resolved, spectral analysis. We find significant,
large (factors of 1.5-6) variations of the absorber column densities on time scales of 40-200 ksec. These values are
10-100 times shorter than those found by Risaliti et al. 2002 in a sample of Seyfert 2 galaxies, and provide strong
constraints on the geometry of the obscuring medium.
1. Introduction
2. Light curves analysis
The geometry of the obscuring matter in Ac-
tive Galactic Nuclei, predicted by unified models
[1], is still unclear. The study of variability of
the absorbing column density (NH ) in these ob-
jects, provides invaluable informations on the size
and location of the absorber. Recently, Risaliti et
al.
[2] found NH variability on time scales from
∼ 2 month to ∼ 5 years in nearly all the objects
belonging to a sample of bright, Compton-thin
Seyfert 2 galaxies with multiple X-ray observa-
tions. This variability suggests that the absorber
could be much closer to the nucleus than usually
assumed.
To further investigate this issue, we have an-
alyzed two BeppoSAX long observations of the
brightest Seyfert galaxy, NGC 4151, and present
here some preliminary results.
The two BeppoSAX observations were per-
formed on 2001 December, 18 and 1996 July, 6
with the Narrow Field Instruments: LECS (0.1-
10 keV), MECS (1.3-10 keV), and PDS (13-200
keV).
The total absorbing column density mea-
sured in NGC4151 by HEAO1, EXOSAT, Ginga,
ROSAT and ASCA [4,5,6,7,8,9,10] ranges from a
few 1022 cm−2 to a few 1023 cm−2. Since the
corresponding photoelectric cut-off lies in the 2-4
keV energy range, we have chosen to analyze the
light curves of the count rate in the bands 2-4 keV
(MECS), 6-10 keV (MECS) and of their softness
ratio (see Figures 1 and 2), to investigate whether
NH variations might have occurred within each
observation. We also analyzed the light curve
of the count rate in the 15-200 keV band (PDS)
to monitor the continuum. From the analysis of
these light curves we have selected some time in-
tervals where we find indications that NH might
have undergone significant changes. These inter-
2
vals are shown and labelled in figures 1 and 2.
To quantify the spectral variations we have then
performed a detailed spectral analysis of these
stretches of data.
A
B
C
M
MECS 2.-4. keV
MECS 6.-10. keV
2.-4. keV/6.-10. keV
1
0.8
0.6
0.4
0.2
4
3
2
1
0
1.1
1
0.9
0.8
0.7
0.6
2.2
2
1.8
1.6
1.4
1.2
1
0.8
0.6
0.4
0.2
1.4
1.2
1
0.8
0.6
0.4
A
B
C
MECS 2.-4. keV
MECS 6.-10. keV
2.-4. keV/6.-10. keV
PDS 15.-200. keV
Time[s]
0.6
0.5
0.4
0.3
4
3.5
3
2.5
2
0
Time[s]
PDS 15.-200. keV
Figure 2. Same panels of figure 1 but for the 1996
July 6 observation.
Figure 1. Light curves of 2001 December 18
observation. From top to bottom: MECS 2-4
keV count rate, MECS 6-10 keV count rate, 2-
4 keV/6-10 keV softness ratio, PDS 15-200 keV
count rate. The capital letters at the top of the
figure, indicate the time intervals selected for a
time-resolved spectral analysis.
3. Spectral analysis
Standard data reduction was performed using
the SAXDAS software package version 2.0 fol-
lowing Fiore, Guainazzi & Grandi [11]. Spec-
tral fits were performed using the XSPEC 11.2.0
software package and public response matrices as
from the 1999 December release. The PDS data
were reduced using the "variable risetime thresh-
old" technique to reject particle background [11].
LECS and MECS spectra were rebinned follow-
ing two criteria:a) to sample the energy resolution
of the detectors with three channels at all ener-
gies whenever possible, and to obtain at least 20
counts per energy channel. Constant factors have
been introduced in the fitting models in order to
take into account the intercalibration systematics
between the instruments [11]. The model adopted
for the analysis has the following components:
1. A power-law with an exponential high-
energy cut-off;
2. a neutral Compton reflection component
(modelled with PEXRAV [3]);
3. a narrow iron Kα emission line with energy
set to 6.38 keV in the observer frame;
4. two components for the soft excess emis-
sion: a thermal component, and a power
law, reflection component.
5. two components for the intrinsic absorber,
both assumed neutral: NH1 which covers
the nucleus totally; NH2 which covers the
nucleus only in part, with a covering factor
CV [4,5,6,7,8,9,10]. The absorber is com-
pleted with the line of sight column den-
sity through the Galaxy (NHgal =2.17×1020
cm−2[12]).
Figures 3 and 4 illustrate the results for the
two observations. In the 2001 December, 18 ob-
servation, we find that from A to B to C, both
absorbers and the photon index do not show sta-
tistically significative changes. From C to M the
two absorbers NH1 and NH2 change at a confi-
dence level ∼ 90% by a factor of 2 and 3.3 re-
spectively, while the photon index is constant at
a confidence level ∼ 99%. If the covering factor
CV is kept constant throughout the fit the ampli-
tude of the variations of NH1 and NH2 is reduced
to a factor 1.4 and 2 respectively. The variations
occur on a timescale of 100-150 ksec.
In the 1996 July, 6 observation, we find that
from A to B both absorbers undergo changes at a
confidence level ∼ 90%, while from A to C the ab-
sorbers change at a confidence level > 99%. From
A to B to C, the photon index is again statisti-
cally consistent with a constant. The amplitude
of the variation of NH1 and NH2 is a factor 6 and
5 respectively. If the covering factor CV is kept
constant throughout the fit the amplitude of the
variations of NH1 and NH2 is reduced to a factor
4 and 3 respectively. The variations occur on a
timescale of ≈ 30 ksec from A to B and ≈ 200
ksec from A to C.
4. Conclusion
We have performed a detailed time-resolved
spectral study of two long BeppoSAX observa-
tions of the Seyfert galaxy NGC4151. We find
strong evidences for spectral variability which we
model in terms of variations of two absorbing
screens. The timescales of these variations are be-
tween 40 ksec and 200 ksec, 10-100 times shorter
than the typical timescales found by Risaliti et
al.[2] in a sample of Compton thin Seyfert 2 galax-
ies. If we associate these timescales to a crossing
time, the resulting linear size would be of about
3
Figure 3. From top to bottom: 68, 90, 99% χ2
confidence contours for NH2 versus NH1 , NH2 ver-
sus photon index Γ and NH1 versus Γ, for the
spectra corresponding to the intervals A, B, C
and M in the 2001 December, 18 observation.
4
one light day, which is the size of the innermost
part of the Broad Line Region, as inferred from
reverberation mapping in NGC 4151. This inter-
pretation is supported by the argument, used by
Risaliti et al.[2], which assume that the absorb-
ing medium might be made up of spherical clouds
moving with Keplerian velocities around the cen-
tral black hole. This requires that the cloud gas
density is of the order of 109 cm−3, typical of
Broad Line Region clouds [13].
REFERENCES
1. Antonucci, R. 1993 ARA&A, 31, 47
2. Risaliti, G., Elvis, M., Nicastro, F. 2002 ApJ,
571, 234
3. Magdziarz, P. & Zdziarski, A. 1995, MNRAS,
273, 837
4. Holt, S. S., Mushotzky, R. F., Boldt, E. A. et
al. 1980, ApJ, 241L, 13
5. Perola, G.C., Piro, L., Altamore, A. et al.
1986, ApJ, 306, 508
6. Fiore, F., Perola, G. C., Romano, M. 1990,
MNRAS, 243, 522
7. Weaver, K.A., Yaqoob, T., Holt, S.S. et al.
1994ApJ,436L, 27
8. Piro, L., Nicastro, F., Feroci, M. 1998,
axrs.symp, 481
9. Wang, J.-X., Zhou, Y.-Y., Wang, T.-G. 1999,
ApJ, 523L, 129
10. Zdziarski, A., Leighly, K.M., Matsuoka, M. et
al. 2002, ApJ, 573, 505
1999,
Handbook
for
11. Fiore, F., Guainazzi, M. & Grandi,
Bep-
analysis,
P.
poSAX
ftp://ftp.asdc.asi.it/pub/sax/doc/software docs/saxabc v1.2.ps.gz
or
http://heasarc.gsfc.nasa.gov/docs/sax/abc/saxabc/
spectral
NFI
12. Murphy, E.M., Lockman, F.J., Laor, A. et al.
1996 ApJS, 105, 369
13. Blandford, R., Netzer, A. and Woltjer, L.
1990: Active Galactic Nuclei, Springer-Verlag
edt.
Figure 4. From top to bottom: 68, 90, 99% χ2
confidence contours for NH2 versus NH1 , NH2 ver-
sus photon index Γ and NH1 versus Γ, for the
spectra corresponding to the intervals A, B and
C in the 1996 July, 6, observation.
|
astro-ph/0510550 | 1 | 0510 | 2005-10-19T07:36:31 | Dark energy with polytropic equation of state | [
"astro-ph"
] | Equation of state parameter plays a significant role for guessing the real nature of dark energy. In the present paper polytropic equation of state $p=\omega\rho^n$ is chosen for some of the kinematical $\Lambda$-models viz., $\Lambda \sim (\dot a/a)^2$, $\Lambda \sim \ddot a/a$ and $\Lambda \sim \rho$. Although in dust cases ($\omega=0$) closed form solutions show no dependency on the polytropic index $n$, but in non-dust situations some new possibilities are opened up including phantom energy with supernegative ($\omega<-1$) equation of state parameter. | astro-ph | astro-ph |
Dark energy with polytropic equation of state
Satyabharati Vidyapith, Kolkata 700 126, North 24 Parganas, West Bengal, India
Utpal Mukhopadhyay
Saibal Ray
Department of Physics, Barasat Government College,
Kolkata 700 124, North 24 Parganas, West Bengal, India
and Inter-University Centre for Astronomy and Astrophysics,
PO Box 4, Pune 411 007, India; e-mail:[email protected]
(Dated: September 24, 2018)
Equation of state parameter plays a significant role for guessing the real nature of dark energy.
In the present paper polytropic equation of state p = ωρn is chosen for some of the kinematical
Λ-models viz., Λ ∼ ( a/a)2, Λ ∼ a/a and Λ ∼ ρ. Although in dust cases (ω = 0) closed form solutions
show no dependency on the polytropic index n, but in non-dust situations some new possibilities
are opened up including phantom energy with supernegative (ω < −1) equation of state parameter.
PACS numbers: 04.20.-q, 04.20.Jb, 98.80.Jk
I.
INTRODUCTION
Ever since the discovery of an accelerating Universe
through SN Ia observations [1, 2], scientists are search-
ing for the cause behind this acceleration. It has been
suspected that some kind of yet unknown energy is re-
sponsible for injecting the right amount of energy to the
Universe for changing it to an accelerating one from a
decelerating phase. Scientific community has coined the
name dark energy for this unknown energy. Various types
of models have been proposed for approximating this
dark energy so far [3, 4].
It is observed that equation of state parameter plays
a crucial role in understanding the actual nature of dark
energy [5]. Till now, most of the models of Λ (the so-
called cosmological constant of Einstein and a represen-
tative symbol for dark energy) with dynamic character
have relied on the barotropic equation of state with wide
range of values of the equation of state parameter ω, viz.,
ω = 0 (for dust filled Universe), ω = 1/3 (for radiation),
ω = −1 (for vacuum), −1 < ω < 0 (for quintessence)
and ω < −1 (for phantom energy). But, all these models
are plagued by some shortcomings. For instance, models
with ω = 0 and 1/3 show an excess [6] or very low [3]
age of the Universe while the stiff-fluid model, in spite of
its nice agreement with the present age of the Universe
[7], is not, in general, accepted as the real nature of the
present Universe. In this circumstances, it is not unrea-
sonable to think of an equation of state different from the
barotropic one. This has prompted us to investigate the
reaction of some of Λ models when polytropic equation
of state is chosen.
Polytropic equation of state has been used in various
astrophysical situations such as in the case of Lane-
Emden models [8, 9]. In the present investigation it is
used in cosmological realm. As test models we have
selected Λ ∼ ( a/a)2, Λ ∼ a/a and Λ ∼ ρ models -- the
same three kinematical Λ-models which were chosen
in one of our previous works [10] under the barotropic
equation of state. Here Sections 2 and 3 deal with the
field equations and their solutions under three different
Λ-models while various physical
implications of the
present work are discussed in Section 4.
II. EINSTEIN FIELD EQUATIONS
The Einstein field equations are
Rij −
1
2
Rgij = −8πG(cid:20)T ij −
Λ
8πG
gij(cid:21)
(1)
where cosmological constant Λ is assumed as a function of
time, viz., Λ = Λ(t) and the velocity of light c in vacuum
is unity when expressed in relativistic units.
For the spherically symmetric Friedmann-Lemaitre-
Robertson-Walker (FLRW) metric
ds2 = −dt2 + a(t)2(cid:20) dr2
1 − kr2 + r2(dθ2 + sin2θdφ2)(cid:21) (2)
where a is the scale factor and k, the curvature constant
−1, 0, +1 respectively for open, flat and close models of
the Universe, the Einstein field equations (1) take the
forms as
a(cid:19)2
(cid:18) a
=
8πG
3
ρ +
Λ
3
,
a
a
= −
4πG
3
(ρ + 3p) +
Λ
3
.
(3)
(4)
As various observational results [11] and inflation theory
indicate that the Universe is flat so we have assumed
here k = 0.
Let us choose the polytropic equation of state
p = wρn
(5)
where equation of state parameter w can take the
constant values 0, 1/3, −1 and +1 respectively for the
dust, radiation, vacuum fluid and stiff fluid and n is the
polytropic index.
III. PARTICULAR SOLUTIONS
A. Λ ∼ ( a/a)2
If we use the ansatz
a(cid:19)2
Λ = 3α(cid:18) a
(6)
where α is a constant, then using equation (6) we get
from equation (3)
a(cid:19)2
3(cid:18) a
a(cid:19)2
− 3α(cid:18) a
= 8πGρ
(7)
which takes the form for density as follows:
1. Dust case (ω = 0)
2
For pressureless dust, ω = 0 = A and equation (13)
becomes,
dy
da
+
(2 − 2n)B
a
y = 0.
Solving equation (14) we get,
y = C1a(2n−2)B
C1 being an integration constant.
This immediately yields
da
dt
= C2a−B
(14)
(15)
(16)
1/(2−2n).
where C2 = C1
Integrating equation (16) and using the initial condition
that a = 0 for t = 0 we get,
a(t) = C3(cid:20) 3(1 − α)
2
(cid:21)2/3(1−α)
t2/3(1−α)
(17)
ρ =
3(1 − α)
a(cid:19)2
8πG (cid:18) a
.
(8)
where C3 = C2
which provides the Hubble parameter as
1/(B+1) being an integration constant,
Using equation (5) and equation (6) we get from equation
(4)
H(t) =
a
a
=
2
3(1 − α)
1
t
.
(18)
(19)
(20)
3(cid:18) a
a(cid:19)2
a(cid:19) = −12πGωρn − 4πGρ + 3α(cid:18) a
(9)
which, on simplification, yields
a2n−1 d2a
dt2 + A(cid:18) da
dt(cid:19)2n
dt(cid:19)2
+ Ba2n−2(cid:18) da
= 0
(10)
where A = [3nω(1 − α)n]/[2n(4πG)n−1] and B = (1 −
3α)/2.
Let us substitute, z = da/dt for which the equation
(10) reduces to
z1−2n dz
da
+ Aa1−2n + z2−2n B
a
= 0.
(11)
If we put, z2−2n = y, then differentiating it with respect
to the scale factor a, we get
z1−2n dz
da
=
1
(2 − 2n)
dy
da
.
(12)
Hence equation (11), by use of equation (12) becomes,
dy
da
+
(2 − 2n)B
a
y = −(2 − 2n)Aa1−2n.
(13)
Now let us solve equation (13) in some particular cases
for getting insight into the physical situations.
Therefore from (8), we get
ρ(t) =
1
6πG(1 − α)
1
t2 .
Also, from equation (6), we obtain
Λ(t) =
4α
3(1 − α)2
1
t2 .
Thus,
it is interesting to note that in dust case, ex-
pressions for ρ(t) and Λ(t) as well as t dependent part
of a(t) are independent of n. Moreover, a(t), ρ(t) and
Λ(t) follow the same power law with t as those in the
dust case with barotropic equation of state (i.e. n = 1)
[10]. But, surprisingly, equation (17) becomes undefined
for n = 1 since C3 contains the factor (2 − 2n) in the
denominator.
2. Non-dust case (ω 6= 0)
When ω 6= 0, then A 6= 0 and hence equation (11) is a
linear equation. Multiplying equation (11) by integrating
factor a(2−2n)B and solving the resulting equation we get,
y = −
Aa(2−2n)B
B + 1
(21)
which yields after substituting the value of y
B. Λ ∼ ρ
da
dt
= −
A
(B + 1)
a.
(22)
For this model we use the ansatz
3
(31)
(32)
(33)
Λ = 8πGγρ
where γ is a constant.
Then, from equation (3) we get
ρ =
3
a(cid:19)2
8πG(γ + 1) (cid:18) a
.
Using (5) we get from (4)
3(cid:18) a
a(cid:19) = −12πGωρn − 4πGρ + 8πGγρ
By solving equation (22) we get our solution for the scale
factor in the form
a(t) = C4exp[(−ω)1/2(1−n)
8πG
3(1 − α)
]1/2t
(23)
where C4 is integration constant.
It is clear from equation (23) that no real value of a(t) is
possible if ω is positive. So, ω must be negative. Putting
specifically, ω = −1 in equation (23) we obtain
a(t) = C4exp[8πG/3(1 − α)]1/2t.
(24)
Now, ω = −1 means a vacuum fluid. So, for vacuum
fluid, we get an exponential solution independent of n.
But, if ω < −1, then scale factor depends on n. Now,
ω < −1 corresponds to a supernegative equation of state
and hence the idea of phantom energy comes into the pic-
ture. Also, during inflation, the Universe underwent an
exponential expansion. So, for non-dust case our solution
reflects an inflationary scenario which may be due to ei-
ther vacuum energy or quintessence or phantom energy.
Also, from equation (24) we get,
H = (cid:20) 8πG
3(1 − α)(cid:21)1/2
.
From equation (8) and (6) we get respectively
ρ(t) = 1,
Λ(t) =
8πGα
1 − α
.
(25)
(26)
(27)
Therefore, the cosmic matter and vacuum energy densi-
ties are given by
Ωm =
8πGρ
3H 2 = 1 − α,
ΩΛ =
Λ
3H 2 = α.
(28)
(29)
So that, equations (29) and (30), immediately provide
Ωm + ΩΛ = 1.
(30)
Thus, we find that although polytropic equation of state
yields interesting result by invoking the idea of phantom
energy for non-dust case, yet it presents us some unac-
ceptable situations like constant H (equation (25)) and
constant ρ (equation (26)). A possible explanation of
this will be discussed afterwards. However, the cosmic
matter and vacuum energy density parameters satisfy
the well-known relation Ωm + ΩΛ = 1 for flat Universe
(Ωk = 0).
which, on simplification, yields the differential equation
a(2n−1) d2a
dt2 + D(cid:18) da
dt(cid:19)2n
dt(cid:19)2
− Ea(2n−2)(cid:18) da
= 0 (34)
where D = ω3n/[2n(γ + 1)n(4πG)n−1] and E = (2γ −
1)/2(γ + 1).
Substituting da/dt = u and then v = u2−2n equation
(34) becomes
1
(2 − 2n)
dv
da
−
E
a
v = −Da1−2n.
(35)
1. Dust case (ω = 0)
For pressureless dust case D=0, so that equation (35)
reduces to
dv
da
+ (2n − 2)
E
a
v = 0.
(36)
Solving equation (36) in the same manner as in the pre-
vious model, we get our solution set as
a(t) = C5(cid:20)
3
2(γ + 1)(cid:21)2(γ+1)/3
t2(γ+1)/3
(37)
where C5 is a constant.
H(t) =
2(γ + 1)
3
1
t
,
ρ(t) =
(γ + 1)
6πG
1
t2 ,
Λ(t) =
4γ(γ + 1)
3
1
t2 .
(38)
(39)
(40)
Also, we can calculate for the cosmic and vacuum energy
densities which, respectively, are Ωm = 1/(γ + 1) and
ΩΛ = γ/(γ + 1). Therefore, Ωm + ΩΛ = 1 and hence
γ = ΩΛ/Ωm. Thus, we find that for this model also a(t),
ρ(t) and Λ(t) follow the same relationship with time as
in the barotropic case [10] and γ is related to matter and
vacuum energy densities by the same relation as in the
barotropic case [10].
2. Non-dust case (ω 6= 0)
When ω 6= 0, then D 6= 0. In this case, following the
same procedure as in the Λ ∼ ( a/a)2 model, it is easy to
show that no real solution is possible for ω > 0.
For ω = −1, we have
a(t) = C6exp(cid:20)
3
8πG(γ + 1)(cid:21)1/2
t
where C6 is a constant.
H = (cid:20)
3
8πG(γ + 1)(cid:21)1/2
,
ρ(t) = (cid:20)
3
8πG(γ + 1)(cid:21)2
,
Λ(t) =
9γ
8πG(γ + 1)2 .
(41)
(42)
(43)
(44)
4
Here we find that in non-dust case, solutions are possible
for vacuum fluid (ω = −1), quintessence (−1 < ω <
0) and phantom energy (ω < −1). For vacuum fluid,
solution is independent of n whereas for quintessence and
phantom energy solution depends on n. In this case also,
γ is found to be the ratio of vacuum energy density and
matter energy density. For physical reality, γ > −1 to be
imposed on the solutions.
C. Λ ∼ a/a
If we use the supposition Λ = β(a/a) = β( H + H 2),
then from equation (3) we get, 3H 2 = 8πGρ+β( H +H 2).
Therefore
ρ =
1
8πG
[(3 − β)H 2 − β H].
(45)
from (4) we have, 3( H + H 2) =
Then, using (5),
−12πGωρn − 4πGρ + β( H + H 2) which on simplification
becomes
3
2 h(3 − β)H 2 + (2 − β) Hi = −
12πGω
(8πG)n [(3 − β)H 2 − β H]n.
(46)
(52)
(53)
Λ(t) =
β(β − 2)
(β − 3)2
1
t2 ,
H(t) = (cid:18) β − 2
β − 3(cid:19) 1
t
.
For simplifying equation (46), let us investigate in some
particular cases.
If we specifically choose n = 0 (i.e.
p = ω), then equation (46) reduces to
(2 − β) H + (3 − β)H 2 = −8πGω.
(47)
1. Dust case (ω = 0)
For pressureless dust (ω = 0) equation (47) becomes
(2 − β)
dH
dt
+ (3 − β)H 2 = 0.
(48)
By solving equation (48) and using the initial condition
that a = 0 when t = 0 we get
da
dt
= a(cid:18) 2 − β
3 − β(cid:19) 1
t
.
(49)
Then from equation (49) we finally get our solution set
as
a(t) = C7t(β−2)/(β−3)
(50)
where C7 is a constant.
Thus, for n = 0, the solution set is same for a(t), ρ(t)
and Λ(t) as obtained by Ray and Mukhopadhyay [10] in
the dust case with barotropic equation of state.
2. Non-dust case (ω 6= 0)
In non-dust case, equation (47) can be written as
(2 − β)
dH
dt
+ (3 − β)H 2 = −8πGω.
(54)
By solving equation (47) and remembering that H = 0
when t = 0 we get
da
dt
= νa tan(µt)
(55)
ρ(t) =
1
4πG (cid:18) β − 2
β − 3(cid:19) 1
t2 ,
(51)
where µ = [8πGω(3 − β)/(β − 2)]1/2 and ν = [8πGω/(3 −
β)]1/2.
After solving equation (55) we have
a(t) = C8(cid:20)sec
{8πGω(3 − β)}1/2
(β − 2)
t(cid:21)
{8πGω/(3−β)}1/2
,(56)
5
ρ(t) =
(8πGω)3/2(3 − β)1/2
8πG(β − 2)2
(cid:20){8πGω(3 − β)}1/2tan2 {8πGω(3 − β)}1/2
(β − 2)
t − βsec2 {8πGω(3 − β)}1/2
(β − 2)
t(cid:21) ,
(57)
Λ(t) =
(8πGω)3/2
(β − 2)2 β(cid:20)(3 − β)1/2sec2 {8πGω(3 − β)}1/2
(β − 2)
t + (8πGω)1/2tan2 {8πGω(3 − β)}1/2
(β − 2)
t(cid:21) ,
(58)
H(t) =
8πGω
(β − 2)
tan
{8πGω(3 − β)}1/2
(β − 2)
t.
(59)
From the above solution set, it is clear that if ω > 0, then
for real a(t), ρ(t), Λ(t) and H(t), β must be less than
3. For negative ω (i.e., for vacuum fluid, quintessence
and phantom energy) real values of a(t) and H(t) can
be obtained, but Λ(t) and ρ(t) become imaginary.
If
(β − 3) = 0, then ρ(t), Λ(t) and H(t) are zero but a(t)
is undefined. Since tant and sect are both increasing
functions of t, then a(t), Λ(t) and H(t) increase with
time. Increasing a(t) supports the idea of an accelerating
Universe, but increasing Λ(t) and H(t) are contrary to
the present status of those two parameters.
For t = 0, we get respectively from equations (56) - (59)
the following physical parameters:
a(t) = C8,
(60)
ρ(t) = −β
1
(8πGω)3/2(3 − β)1/2
8πG
(β − 2)2
,
(61)
Λ(t) =
(8πGω)3/2
(β − 2)2 β(3 − β)1/2,
H(t) = 0.
(62)
(63)
Since a(t) assumes a constant value for t = 0, then our
model hints at the existence of some kind of quantum
fluctuation at the time of Big Bang. Now, equation (61)
tells us that for physically valid ρ, ω > 0 and β < 0. For
a negative β we get an attractive Λ. So, as a whole this
model also presents some interesting as well as awkward
cosmological picture.
IV. DISCUSSION
Present investigation reveals that for pressureless
dust, all the three models reflect the same result as
obtained by Ray and Mukhopadhyay [10] in dust case
with barotropic equation of state. But,
in non-dust
cases (i.e., ω 6= 0), Λ ∼ ( a/a)2 and Λ ∼ ρ models show
exponential expansion of the Universe for negative ω
whereas no real situation is possible for ω > 0. Thus,
we can say that for dust-filled Universe, there is no
distinction between barotropic and polytropic equations
of state. This is not at all unexpected because for ω = 0,
we have p = 0, whatever may be the value of n in the
equation of state. Also, non-dust cases for all the three
models present us some unpleasant results in terms of
constant H and ρ. These situations can be explained
if we assume that the non-dust cases reflect the picture
of early Universe when inflation occurred due to the
presence of quintessence (−1 < ω < 0) or vacuum fluid
(ω = −1) or phantom energy (ω < −1). So, using
polytropic equation of state it has been possible to show
that non-dust cases admit the presence of a driving force
behind inflation in the form of either quintessence or
vacuum fluid or phantom energy and in the dust cases
there is no distinction between different equation of
states. Moreover, present models do not depend on any
particular value of n in the polytropic equation of state.
This proves the generality of the present investigation.
Acknowledgments
One of the authors (SR) would like to express his grat-
itude to the authority of IUCAA, Pune for providing him
the Associateship Programme under which a part of this
work was carried out.
6
[1] A G Riess et al Astron. J. 116 1009 (1998).
[2] S Perlmutter et al Nat. 391 51 (1998).
[3] J M Overduin and F I Cooperstock Phys. Rev. D 58
043506 (1998).
[4] V Sahni and A Starobinsky Int. J. Mod. Phys. D 9, 373
(2000).
[5] R P Kirshner Science 300 5627 (2003).
[6] R G Vishwakarma Class. Quan. Grav. 17 3833 (2000).
[7] S Ray and U Mukhopadhyay astro-ph/0411257.
[8] R N Tiwari , J R Rao and R R Kanakamedala Phys. Rev.
D 34 1205 (1986).
[9] S Ray Astrophys. Space Sci. 280 345 (2002).
[10] S Ray and U Mukhopadhyay astro-ph/0407295.
[11] P de Bernardis et al Nat. 404 955 (2000).
|
astro-ph/9705195 | 2 | 9705 | 1997-09-29T18:17:25 | A model independent lower limit on the number of Gamma Ray Burst hosts from repeater statistics | [
"astro-ph"
] | We present a general statistical analysis of Gamma Ray Bursts embedded in a host population. If no host generates more than one observed burst, then we show that there is a model independent lower bound on the number of hosts, $H$, of the form $H > c B^2$, where B is the number of observed bursts, and $c$ is a constant of order one which depends on the confidence level (CL) attached to the bound. An analysis by Tegmark et al. (1996) shows that the BATSE 3B catalog of 1122 bursts is consistent with no repeaters being present, and assuming that this is indeed the case, our result implies a host population with at least H=1.2x10^6 members. Without the explicit assumption of no repeaters, a Bayesian analysis based on the results of Tegmark et al. (1996) can be performed which gives the weaker bound of $H>1.7\times 10^5$ at the 90% CL. In the light of the non-detection of identifiable hosts in the small error-boxes associated with transient counterparts to GRBs, this result gives a model independent lower bound to the number of any rare or exotic hosts. If in fact GRBs are found to be associated with a particular sub-class of galaxies, then an analysis along the lines presented here can be used to place a lower bound on the fraction of galaxies in this sub-class. Another possibility is to treat galaxy clusters (rather than individual galaxies) as the host population, provided that the angular size of each cluster considered is less than the resolution of the detector. Finally, if repeaters are ever detected in a statistically significant manner, this analysis can be readily adapted to find upper and lower limits on $H$. | astro-ph | astro-ph | A model independent lower limit on the number of Gamma Ray
Burst hosts from repeater statistics
Anupam Singh1 and Mark Srednicki2
Department of Physics, University of California, Santa Barbara, CA 93106
ABSTRACT
We present a general statistical analysis of Gamma Ray Bursts embedded in
a host population. If no host generates more than one observed burst, then we
show that there is a model independent lower bound on the number of hosts,
H, of the form H > cB2, where B is the number of observed bursts, and c is a
constant of order one which depends on the confidence level (CL) attached to the
bound. An analysis by Tegmark et al. (1996) shows that the BATSE 3B catalog
of 1122 bursts is consistent with no repeaters being present, and assuming
that this is indeed the case, our result implies a host population with at least
H = 1.2 × 106 members. Without the explicit assumption of no repeaters, a
Bayesian analysis based on the results of Tegmark et al. (1996) can be performed
which gives the weaker bound of H > 1.7 × 105 at the 90% CL. In the light of
the non-detection of identifiable hosts in the small error-boxes associated with
transient counterparts to GRBs, this result gives a model independent lower
bound to the number of any rare or exotic hosts. If in fact GRBs are found
to be associated with a particular sub-class of galaxies, then an analysis along
the lines presented here can be used to place a lower bound on the fraction of
galaxies in this sub-class. Another possibility is to treat galaxy clusters (rather
than individual galaxies) as the host population, provided that the angular size
of each cluster considered is less than the resolution of the detector. Finally,
if repeaters are ever detected in a statistically significant manner, this analysis
can be readily adapted to find upper and lower limits on H.
7
9
9
1
p
e
S
9
2
2
v
5
9
1
5
0
7
9
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
Subject headings: gamma rays: bursts
1E-mail: [email protected]
2E-mail: [email protected]
-- 2 --
1.
Introduction
Recently there have been exciting developments in the observational investigation of
the nature of Gamma Ray Bursts (GRBs) and the hosts in which they may originate.
(Several excellent reviews of GRBs are available; see e.g. Blaes [1994]; Piran [1995].) For
GRB970228, an afterglow was detected in the X-Ray and optical wavebands (Van Paradijs
et al. 1997), and Metzger et al. (1997) have identified absorption lines in the spectra of the
optical transient associated with GRB970508. Their observations put the transient beyond
a redshift of z = 0.835. If confirmed, this would establish that GRBs are cosmological. If
so, then this opens the door to their use as tools for probing cosmological issues such as
structure formation (Piran & Singh 1997).
The latest observations herald a new era of more precise, useful information about
GRBs. The counterparts in the X-ray and optical wavebands allow one to get among other
things a more precise location of the bursting source. Further, one can perform a detailed
and intensive survey of the neighboring region of space to seek out the host population
in which the GRBs are embedded. So far, in spite of exhaustive searches, statistical
analysis, and the accompanying debate (Fenimore et al. 1993; Larson, McLean & Becklin
1996; Schaefer et al. 1997; Blaes et al. 1997; Kolatt & Piran 1996; Hurley et al. 1997),
GRBs have not clearly and unambiguously been associated with any other well-known
population of astronomical objects. In particular, no known or identifiable galaxy has been
associated with either of the two GRBs for which transient counterparts have enabled a
small error-box. Regarding repetitions of GRBs from the same host, while there was an
early claim of evidence for this by Quashnock & Lamb (1993), their analysis was called
into question by Narayan & Piran (1993). More recently, Tegmark et al. (1996) concluded
that the data is consistent with no repetition of GRBs. Thus all of the recent data has not
yet resolved the mystery of the nature of the host population in which GRB sources are
embedded.
It is therefore important to deduce all possible constraints which can be placed on
the host population; model-independent constraints are particularly valuable. Obviously,
there must be at least as many hosts as there have been GRBs, assuming no repeaters.
This is clearly not a significant constraint; the total number of galaxies is about 109,
corresponding to about 10−6 bursts per galaxy per year. It is possible, however, that the
host population consists of exotic or otherwise rare cosmological objects. In this Letter,
we perform a statistical analysis which demonstrates that, if a total of B bursts have been
seen, and we assume that there are no repeaters, then the number of hosts H is likely to
exceed cB2, where c is a numerical constant which depends on the desired confidence level
and the precise form of the statistic (e.g., choice of the prior distribution in a Bayesian
-- 3 --
analysis), but not on any details of the GRB model. If we do not make the assumption of
no repeaters, but instead appropriately weight the possibilities based on observational data,
then the lower bound is reduced to Hmin ∼ B2/hN2i, where hN2i is the expected number of
repeaters in the data. In this analysis, a "host" is any object whose angular size is much
less than the resolution of the detector.
The BATSE data is consistent with no repeaters in the analyzed catalog of 1122 bursts.
An analysis by Tegmark et al. (1996) shows that the number of repeaters N2 is less than 10
at the 95% CL. If we assume that there are in fact no repeaters in this catalog, then we can
conclude that the the population of hosts from which the GRBs may be originating should
contain at least H = 1.2 × 106 members. This implies that about 0.1% of the total number
of galaxies could be hosts, if GRBs are embedded in galaxies. This could be a significant
constraint if GRBs are found to be associated with a specific sub-class of galaxies. This
constraint would also be significant if ongoing searches for potential objects within the error
boxes of the transients fails to turn up identifiable normal galaxies or other well-known and
abundant astronomical objects. Furthermore, if repeaters are ever detected in a statistically
significant manner, our analysis can be readily adapted to find upper and lower limits on
the number of hosts H.
We now turn to the statistical analysis of this problem which allows us to draw the
advertised conclusions.
2. The General Statistical Analysis
We suppose that there is a population of H hosts, each of which is producing burst
bursts at an individual rate of γi, i = 1, . . . , H, and that each of these hosts has been
observed for an amount of time ti. Then the number of bursts which have been observed
from each host is expected to be Ai = γiti. Call the actual number of observed bursts from
each host Bi, and the total number of observed bursts B ≡ PH
i=1 Bi.
We assume that the bursts from each host are uncorrelated, so that the probability of
observing Bi bursts from host i is given by a Poisson distribution
P (BiAi) =
ABi
i exp(−Ai)
Bi!
.
(1)
Then the probability that a total of B bursts are observed is
P (BA, H) =
B
XB1=0
. . .
B
XBH =0
δB1+...+BH ,B P (B1A1) . . . P (BHAH)
-- 4 --
=
(AH)B exp(−AH)
B!
,
(2)
where A is the average value A ≡ (1/H)PH
i=1 Ai.
We wish to calculate that probability that B bursts are seen, but that there are no
repetitions; this is given by
P (B, no rep{Ai}, H) =
1
XB1=0
. . .
1
XBH =0
δB1+...+BH ,B P (B1A1) . . . P (BHAH)
(3)
To simplify the calculation, we first treat the idealized situation in which Ai = A for each
host. In this case, eq. (3) becomes
P (B, no repA, H) =
H!
B!(H − B)!
AB exp(−AH) ,
(4)
The probability that no repetitions are seen, given that B bursts are observed, is given by
the ratio
P (no repB, A, H) =
=
P (B, no repA, H)
P (BA, H)
H!
,
H B(H − B)!
(5)
which we see is independent of A. Assuming H ≫ B ≫ 1, and using Stirling's formula for
the factorials, we get
P (no repB, A, H) = exp(−B2/2H) .
(6)
This is our key result. It tells us that we are unlikely to have seen no repetitions among
B bursts if H is significantly smaller that B2. This can be made more precise by using a
Bayesian analysis; see, e.g., Loredo and Wasserman (1995). Assuming that no repetitions
are seen, the probability that the number of hosts is between H and H + dH is proportional
to the likelihood P (no repB, A, H) times a prior distribution; if we make the standard
scale-invariant choice dH/H with an upper cutoff at Hmax = 109, we find that H > B2 with
a confidence level of approximately 90% (the exact value depends weakly on B and Hmax).
If some hosts generate repeat bursts, the analysis above is changed somewhat. Let Nν
be the number of hosts which generated ν bursts each; then H = N0 + N1 + N2 + . . . and
B = N1 + 2N2 + 3N3 + . . .. We assume N0 ≫ N1 ≫ N2,3,.... Then eq. (5) is replaced with
P (N2, N3, . . .B, A, H) =
=
H! B! H −B
(0!)N0N0! . . . (ν!)Nν Nν! . . .
(B2/2H)N2
N2!
exp(−B2/2H) exp(−2N 2
2 /B)δN30δN40 . . . .
(7)
-- 5 --
Terms with Nν 6= 0 for ν ≥ 3 are suppressed by at least one factor of B/H. Henceforth, we
set N3 = N4 = . . . = 0.
Let us now consider what can be deduced from our formulas when they are applied to
the BATSE 3B catalog of B = 1122 bursts. First we will examine the results of Tegmark et
al (1996) concerning repeaters. To gain an analytic understanding of these results, we will
treat the simplified case in which the beam-width σ is the same for every burst; then their
statistic R becomes
B
j−1
R =
1
2B
Xi=1
Xi=1
exp(−θ2
ij/4σ2) ,
(8)
where θij is the angle on the sky between bursts i and j.
We now compute the expected values of R and R2, assuming that the hosts are
uniformly distributed on the sky, that the angular size of each host is much less than σ, and
that each the observed location of each burst is displaced from its true value by a random
angle selected from a gaussian distribution of width σ. (The latter two points are relevant
only for bursts from repeaters.) In the flat-sky approximation (σ ≪ 1), we find
hRi = 1
4Bσ2 + 1
8 f
(∆R)2 ≡ hR2i − hRi2
= 1
16σ2 + 1
96B−1f ,
(9)
(10)
where f ≡ 2N2/B is the fraction of bursts which came from repeaters. If each host has an
angular size φ, then the coefficient of f in eq. (9) is reduced; if the burst probability across
a host has a gaussian profile, then the reduction factor is σ2/(σ2 + φ2). Not including such a
factor in our subsequent analysis implies that each of our hypothetical hosts has an angular
size φ which is much less than σ.
Eqs. (9,10) can be understood heuristically. Roughly speaking, hRi is the expected
number of bursts which lie within an angular distance σ of any one particular burst. There
is a contribution to hRi due to bursts from different hosts; this contribution is of order
Bσ2. There is a second contribution if the particular burst in question is a repeater; the
probability that this is the case is of order f . This explains the structure of eq. (9). To
understand the variance (∆R)2, we note that the Bσ2 term in hRi comes from a Poisson
process, so its variance is also Bσ2. However, averaging over the particular burst in question
reduces the variance by a factor of B. The term of order f in (∆R)2, which in practice is
not important, can be similarly understood.
The probability distribution for R should be approximately gaussian,
P (RN2, B) =
1
√2π∆R
exp[−(R − hRi)2/2(∆R)2] ,
(11)
-- 6 --
where Rav and ∆R are given in terms of N2 = f B/2 and B by eqs. (9,10). For BATSE, in
which each burst has its own value of σ, and in which sky coverage is not uniform, eqs. (9,10)
should be approximately correct, but possibly with different numerical coefficients. For
this case, we turn to the Monte Carlo results of Tegmark et al. (1996). Their Figure 1
0 P (r)dr vs. R for several different values of f . From this, we can read off
shows R R
hRi = 0.316 + 0.15 f and ∆R = 0.0068 for f = 0. This is consistent with our eqs. (9,10)
with a value of σ ∼ 2◦. (We neglect the f dependence of ∆R, which is difficult to extract
accurately, since it will not be important to our conclusions.) The experimental value of
R for BATSE is Rexp = 0.308. Assuming that there are in fact no repeaters, then the
probability that the BATSE data would yield a value of R which is less than Rexp is 0.11.
Eq. (11) is also consistent with the results of Tegmark et al. (1996) that f < 0.018 at 95%
CL and f < 0.049 at the 99% CL.
We now combine eq. (7) and eq. (11) to get
P (RexpB, A, H) =
B/2
XN2=0
P (RexpN2, B)P (N2B, A, H) .
(12)
Again performing a Bayesian analysis for H with the scale-invariant prior distribution
dH/H and an upper cutoff of Hmax = 109, we find H > 1.7 × 105 with a CL of 91%; this is
smaller than B2 by a factor of 7.4. This factor is roughly the expected number of repeaters
in a Bayesian analysis,
hN2i = P∞
N2=0 P (RexpN2, B)N2
P∞
N2=0 P (RexpN2, B)
,
(13)
where P (RexpN2, B) is given by eq. (11), and we have implicitly assumed a uniform prior
distribution for N2; in the present case, eq. (13) yields hN2i = 12.
Turning to future experiments, we first recall that BATSE happened to produce a low
value of Rexp. A larger catalog for BATSE which instead yielded the most likely value when
f = 0, Rexp = hRif =0 = 1
f > 0. Assuming (for example) a catalog with B = 3000 bursts, we find that H > 3.7 × 105
at the 90% CL. More significant improvements in the lower bound on H would require an
instrument with a better angular resolution.
4 Bσ2, would result in P (Rexpf ) ∝ exp[−(0.15 f )2/2(∆R)2] for
We now turn to the more general case of a population of hosts, each with an individual
Ai (the burst production rate times observation time for that host). We will make the mild
assumption that the distribution of Ai's has moments
1
H
H
Xi=1
An
i ≡ cnAn
(14)
-- 7 --
which remain finite in the limit of large H. Note that c1 = 1 and c2 ≥ 1. From here on we
will concentrate on the simplest case of no repetitions.
The probability that we see B bursts with no repetitions is given by eq. (3). We can
write the Kronecker symbol as
δB1+...+BH ,B = I
dz
2πiz
z−BzB1 . . . zBH ,
(15)
where the contour encloses the origin. Each sum in eq. (3) now yields a factor of
exp(−Ai)(1 + zAi), and so we find
P (B, no rep{Ai}, H) = exp(−AH)I
exp(cid:20)
= I
dz
2πiz
dz
2πiz
z−B
H
(1 + zAi)
Yi=1
H
Xi=1
log(1 + zAi) − B log z − AH(cid:21) .
(16)
We now treat both H and B as large, and evaluate the integral by steepest descent. Let
F (z) be the argument of the exponential in the second line of eq. (16); the point z0 of
steepest descent is then given by F ′(z0) = 0, or
0 =
H
Xi=1
Ai
1 + z0Ai −
B
z0
= H[A − z0(c2A2) + z2
0(c3A3) + . . .] − Bz−1
0
.
Taking B ≪ H, this can be solved by power series to yield
2 − c3)(B/H)3 + . . . .
The probability of getting B bursts with no repetitions is now given by
z0A = (B/H) + c2(B/H)2 + (2c2
P (B, no rep{Ai}, H) = [2πF ′′(z0)]−1/2z−1
0 exp F (z0) .
(17)
(18)
(19)
After dividing by P (BA, H) we find that the probability that no repetitions are seen, given
that B bursts are observed, is
P (no repB,{Ai}, H) = exp(−c2B2/2H)
(20)
in the limit H ≫ B ≫ 1. This is qualitatively the same as our previous result, eq. (6),
except that now the relative value of the second moment of the distribution enters as well.
The coefficient c2 must be greater than one, so this correction can only increase the lower
bound on H.
-- 8 --
3.
Implications for the hosts of GRBs and Discussion
We now turn to the implications of the preceding analysis to the specific case of GRBs
and the host population in which the GRBs may be embedded. We have seen that if we
observe B GRBs, which originated from a host population of H hosts, and further if assume
that no host generated more than one burst, then we can conclude that H must be bounded
below by B2, at a confidence level of roughly 90%. If a probability-weighted average over
the number of repeaters is taken instead, then the lower bound on H drops to a number
which is of order B2/hN2i, where hN2i is the expected number of repeaters. This is given
by eq. (13), which would typically yield hN2i ∼ Bσ (and hence Hmin ∼ B/σ), where σ is
the angular resolution of the detector. Our constraint on H can be significant for models in
which the host population for GRBs consists of exotic objects of any kind, such as galaxies
with some particular morphological feature or other specific signature such as a high star
formation rate. Another possibility is to consider clusters of galaxies as hosts (rather than
individual galaxies), provided that the angular size of the clusters considered is much less
than σ.
If repeaters are ever detected statistically through a value of Rexp which is larger than
hRi by several ∆R, this analysis can be readily adapted to find upper and lower limits on
the number of hosts H. The most likely value will be H = B2/hN2i, where hN2i is the most
likely value of the number of repeaters.
Finally, we restate our main point: simply by knowing that there are ∼ 103 bursts with
no sign of repeaters, we can immediately infer a model-independent lower bound on the
number of hosts in which GRBs may be embedded which is larger by more than two orders
of magnitude.
We thank Omer Blaes for many helpful discussions, and the referee for several
important comments which improved the paper. This work was supported in part by NSF
Grant PHY -- 91 -- 16964.
Blaes, O. 1994, ApJS, 92, 643
REFERENCES
Blaes, O., Hurt, T., Antonucci, R., Hurley, K. & Smette, A. 1997, ApJ, 479, 868
Fenimore, E.E. et al. 1993, Nature, 366, 40
Hurley, K., Hartmann, D., Kouveliotou, C., Fishman, G., Laros, J., Cline, T., and Boer, M.
-- 9 --
1997, ApJ, in press.
Kollat, T. & Piran T. 1996, ApJ, 467, L41
Larson, S., Mc Lean, I. & Becklin, E. 1996, ApJ, 460, L95
Loredo, T.J., and Wasserman, I. 1995, ApJS, 96, 261
Metzger, M., Djorgovski, S., Steidel, C., Kulkarni, S., Adelberger, K. & Frail, D. 1997,
IAUC 6655.
Narayan, R. & Piran, T. 1993, MNRAS, 265, L65
Petrosian V. 1993, ApJ, 402, L33
Piran, T. 1995, in Some unsolved problems in astrophysics, eds. Bahcall, J. & Ostriker, J.
(Princeton: Princeton Univ. Press).
Piran, T. & Singh, A. 1997, preprint astro-ph/9607072, to appear in ApJ, July 1997
Quashnock, J. & Lamb, D. 1993, MNRAS, 265, L59
Schaefer, B., Cline, T., Hurley, K. & Laros, G. 1997, preprint astro-ph/9704278.
Tegmark, M., Hartmann, D., Briggs, M., Hakkila, J. & Meegan, C. 1996, ApJ, 466, 757
This preprint was prepared with the AAS LATEX macros v4.0.
|
astro-ph/0112543 | 1 | 0112 | 2001-12-26T04:29:45 | On the Maximal Quantity of Processed Information in the Physical Eschatological Context | [
"astro-ph"
] | An estimate of the maximal informational content available to advanced extraterrestrial or future (post)human civilizations is presented. It is shown that the fundamental thermodynamical considerations may lead to a quantitative estimate of the largest quantity of information to be processed by conceivable computing devices. This issue is interesting from the point of view of physical eschatology, as well as general futurological topics, like the degree of confidence in long-term physical predictions or viability of the large-scale simulations of complex systems. | astro-ph | astro-ph |
ON THE MAXIMAL QUANTITY OF PROCESSED INFORMATION
IN THE PHYSICAL ESCHATOLOGICAL CONTEXT
MILAN M. ´CIRKOVI ´C
Astronomical Observatory, Volgina 7
11000 Belgrade, Yugoslavia
MARINA RADUJKOV
Petnica Science Center, P. O. Box 118
14000 Valjevo, Yugoslavia
Abstract
An estimate of the maximal informational content available to advanced ex-
traterrestrial or future (post)human civilizations is presented. It is shown that the
fundamental thermodynamical considerations may lead to a quantitative estimate
of the largest quantity of information to be processed by conceivable computing
devices. This issue is interesting from the point of view of physical eschatology,
as well as general futurological topics, like the degree of confidence in long-term
physical predictions or viability of the large-scale simulations of complex systems.
1
Introduction
The problem of intelligent information processing in the ever-expanding universe has
recently been investigated by several authors (Krauss and Starkman 1999; ´Cirkovi´c
and Bostrom 2000). These considerations have entered a new phase, building upon
advances in observational cosmology and the strong foundations of the pioneering studies
in cosmological prediction by Rees (1969), Dyson (1979), Frautschi (1982), Tipler (1986)
and Barrow and Tipler (1986). In the same time, interest in the fundamental physical
limits of computation has grown, largely motivated by tremendous advances in computer
science (e.g. Lloyd 2000). Continuation of such trend of technological progress will lead
humanity, sooner or later, into a stage of highly advanced galactic civilization, in recent
years conventionally dubbed the posthuman era. Apart from the epochal importance of
any such possible development for biological and social sciences, it has a strong bearing
on the nascent astrophysical discipline of physical eschatology, for the reasons suggested
boldly by Dyson (1979), when he wrote:
It is impossible to calculate in detail the long-range future of the universe
without including the effects of life and intelligence.
It is impossible to
calculate the capabilities of life and intelligence without touching, at least
peripherally, philosophical questions.
Certainly, the most important property of intelligence is its capacity for information
processing. As it has been recognized for a long time, this form of information process-
ing is still entirely within limits of physical, specifically thermodynamical, laws. This
1
conclusion does not necessarily entail reductionism, since consciousness may still con-
tain uncomputable and therefore irreducible elements in addition to the conventionally
established ones (e.g. Penrose 1989). And thermodynamics in general is, in the limit
of very long timescales, determined by properties of the universe as a whole, i.e. by
astrophysics and cosmology. Our goal in this note is to consider very simple cases of
maximal informational resources available to advanced civilizations, either extraterres-
trial or posthuman. We do not enter into any possible distinction between the two, since
the ages of humanity, Earth and our Galaxy, as well as the pace of chemical evolution,
suggest that there may be intelligent communities much older than ours even in our
cosmological neighbourhood (Livio 1999).
2 An estimate of the maximal quantity of processed
information
Let us consider various energy fields as potential sources of energy for information pro-
cessing, using the assumption ("Cosmic Sum rule") that Ω = ΩΛ + Ωb + ΩCDM = 1,
where the first term corresponds to the vacuum energy, second to the baryonic matter,
and the third to the (non-baryonic) cold dark matter (for the general cosmological con-
siderations, see Peebles 1993). The best present estimates approach values ΩΛ ≈ 0.7,
ΩCDM ≈ 0.25 and Ωb ≈ 0.05 (e.g. ´Cirkovi´c and Bostrom 2000, and references therein).
One cannot do anything useful with the vacuum energy.1 As far as CDM is concerned, it
could conceivably be used as an energy source, since the annihilation of these cosmions
and anticosmions (present in approximately equal numbers according to the standard
theory) would produce potentially usable energy (for some consequences of annihilation,
see Kaplinghat, Knox and Turner 2000). However, depending on the mass spectrum
of cosmions, their galactic density is rather small, and since their interactions are by
definition very weak, their gathering and separation for annihilation will pose huge en-
gineering problems. If we consider the model of "posthumanity soon", than only usable
matter field is the baryonic matter, in the local environment concentrated in the form
of planetary systems.
We may use the Brillouen's (1962) equality for the upper limit of processable quantity
of information (in bits):
Imax =
∆E
kBT ln 2
= 1.05 × 1016 ∆E
T
.
(1)
Here, kB is the Boltzmann constant. Both ∆E and T are functions of time, as well as of
cosmological parameters Ω, Λ and H0. There are several important consequences of the
Brillouen's inequality for cosmology. The amount of information available for processing
in any causally connected region of finite proper size with trivial topology is necessarily
finite.
As for the working temperature of posthuman computers, we may wish to consider
two somewhat extreme models: (I) the first with T = TCMB = 2.730 ± 0.014 K (Staggs
1On the contrary, it plausibly harms the energy budget by preventing, through the so-called cos-
mological "no-hair" theorem, the development and subsequent exploitation of the cosmological shear
during the Hubble expansion (Gibbons and Hawking 1977).
2
et al. 1996) -- indicating quick reaching of posthuman stage; additional cooling would re-
quire diverting the precious energy resources from computing itself -- and (II) the second
with
T = Tvac =
¯hc
kBs Λ
12π2 = 3.2985 × 10−30h(cid:18) ΩΛ
0.7(cid:19)
1
2
K
(2)
indicating posthumanity in the asymptotic limit of physical eschatology. In this equa-
tion, Λ is the vacuum energy-density corresponding to the dimensionless cosmological
density fraction ΩΛ, and h is the dimensionless Hubble constant (H0 ≡ 100 h km s−1
Mpc−1).
Now we can write for the expendable energy:
∆E = ZV
q(Ωbρcrit)c2 dV,
(3)
where V is the proper volume available to posthuman civilization and q is the energy
extraction efficiency (between 0 and 1, and optimistically over 0.5). The usage of Ωb
is particularly useful when we consider possibility of posthuman civilizations of truly
intergalactic size. Currently there is some confusion about the exact value of Ωb, since
the primordial nucleosynthesis inferences apparently conflicts with conclusions drawn
from the microwave background anisotropy observations (e.g. Kaplinghat and Turner
2001). It is to be expected that this problem will be solved soon, but in the meantime
we note that the exact value is inessential for our purposes. If we restrict ourselves to
the model (I), that is, "posthumanity soon", we may neglect spatial variation of q and
Ωb (which would be present in other cases, say q is certainly different in intergalactic
space from the one within galaxies), and using the definitional relation for Ωb we get
∆E ≈ qc2ZV
ρb dV.
Now, plugging this into (1), we obtain
Imax =
qc2
kB ln 2RV
ρb dV
T
= 3.5 × 1036qZV
ρb dV,
(4)
(5)
in bits. If we wish to consider truly short-term posthuman civilization, we may state
that the value of the integral is equal to
ρb dV = 2 × 1033
ZV
¯M
M⊙
n grams,
(6)
where n = 1, 2... is the number of planetary systems controlled by our prototype posthu-
man civilization, and ¯M is the average mass of a planetary system in the Galaxy. The
fraction in (6) is likely to be less or about 1.5, when mass of the hidden matter (like
comets in the Oort cloud whose total mass is still uncertain; see Weissman 1983) in
our planetary system is taken into account, as well as the possibility of harvesting some
interstellar matter between systems. Taking all this into account, we reach the estimate
of
bits.
(7)
¯M
M⊙
Imax = 7 × 1069qn
3
An analogous estimate can be obtained for the case (II) of posthumans in the far future.
This case will certainly be prone to much larger uncertainties of not only quantitative,
but also qualitative nature. Therefore, we shall here give just a sketch and postpone the
detailed analysis to a forthcoming study. To this end, we may generalize the expression
(5) taking into account the changes in both resources and methods available to advanced
civilizations during long cosmological scales
Imax(t) =
q(t) c2
kB ln 2
t
Zt0
dt
T (t) ZV (t)
ρb dV.
(8)
Here, V (t) denotes the physical (proper) volume of space available to the advanced
civilization under consideration, and T (t) is the evolution of the cosmic temperature
(T (t0) = TCMB = 2.730 ± 0.014 K). In this case, it is important to notice that the evolu-
tion of the equilibrium temperature of the universe in the context of physical eschatology
is a problem not exactly solved to this day (approximate results for Einstein-de Sitter
model are presented in Adams and Laughlin 1997), mainly because at very late epochs
controversial sources (like the proton decay and Hawking radiation of an uncertain num-
ber of decaying black holes) come into play. Fortunately, we may be in position to know
the asymptotic limit of the process of cooling of the universe, since it is determined by
the temperature associated with the cosmological event horizon (Gibbons and Hawking
1977) in eq. (2). Since it can be shown that -- if current estimates of the cosmological
constant are correct -- we have already entered the exponentially expanding (quasi)de
Sitter phase ( ´Cirkovi´c and Bostrom 2000), we may use the simplest approximation of
temperature decreasing as T (t) = T (t0) exp[−H(t − t0)], leading to the formal equalisa-
tion with (2) in t = t0 +(1/H0) ln[T (t0)/Tvac] ≈ t0 +69.58/H0, i.e. in about 1140 Gyr (for
h = 0.6). As shown by Adams and Laughlin (1997; although only for the Einstein-de
Sitter case), this is too pessimistic, since other photon sources will replace CMB as the
determinants of the temperature of the universal heat bath. It will take probably many
orders of magnitude larger time for the universe to reach the de Sitter temperature,
but what is important is that there may be no further cooling. Thus, Dyson's (1979)
idea about using a special form of hybernation for the expression in eq. (1) to diverge
is unfeasible. That said, we leave considerations of the detailed solutions of eq. (8) to
a subsequent work, which will benefit from our increased understanding of the future
thermal history of the universe.
3 Discussion
We have estimated (within an order of magnitude) the information processing power
of advanced extraterrestrial or future posthuman communities. Our estimate is conser-
vative in the sense that we have ignored the complicated and not yet fully understood
issue of dissipationless computation (see, for instance, Porod et al. 1984) and assumed
classical limit of kT log 2 dissipation per logical step. In addition, we have neglected the
important issues of information transmission, noise and error correction (for a prelim-
inary treatment of these, see the pioneering study of Sandberg 2000). All these issues
should be covered in future, more realistic treatments.
It is worth noticing that numerical estimates reached above are quite conservative in
comparison to the proposed fundamental information bounds, such as the holographic
4
bound (e.g. Sussking 1995)
I h
max =
A
P l ln 2
4L2
,
(9)
(where LP l ≡ qG¯h/c3 ≈ 2 × 10−33 cm is the Planck length) or the Bekenstein (1973,
1981) bound
I B
max =
2πRE
c¯h ln 2
,
(10)
max
max
where R is the radius and E the total mass-energy of the information cache. For instance,
the equivalent of a holographic bound for the case of entire planetary systems from eq.
≈ 9.8 × 1076 (n ¯M/M⊙)2 bits, and the application of the Bekenstein bound
(7) gives I h
≈ 5.13 × 1071 q(n ¯M /M⊙)(R/1 cm) bits (where one should
to the same case yields I B
keep in mind that in order for this bound to be meaningful, the size of the memory
R has to be larger from its gravitational radius, which is about 3 × 105 cm per Solar
mass). Parenthetically, we note that most of the treatments of cosmological limitations
on computation to be found in the literature (e.g. Tipler 1986; Sandberg 2000) and on
the Internet use the Bekenstein bound, which is certainly more realistic and practical,
but does not look entirely sound from the conceptual point of view. The reason for
such a disadvantage of the Bekenstein bound compared to the holographic bound is
that the Newtonian gravitational constant G figures explicitly in the latter and not
in the former. Since we believe that quantum gravitational degrees of freedom may
contain a huge amount of information (and could be potentially exploited by advanced
communities), a liberal approach would favor the usage of the holographic bound.
There are several reasons for pursuing this problem further. One is related to the
confidence we may have in predictions of nonlinear dynamical systems by future com-
puting devices. It is a well-known fact that (apparent) complexity of nonlinear systems
increase at an exponential rate with time. No matter how advanced computing devices
are employed in order to predict the behaviour of such systems, the prediction will break
down at a particular timescale. One of the main factors determining this timescale is
certainly the maximal amount of information which can be processed whatsoever during
the computations necessary for prediction. But probably the most interesting issue to
be solved by calculations such as these is the question of information cost of running
large-scale detailed simulations of human environment. There are reasons to believe
that advanced posthuman civilizations will run such simulations, which are sometimes
aptly called "ancestor simulations" (Prof. Nick Bostrom, manuscript in preparation).
It is clear, however, that the depth (or complexity) of such a simulation would be very
sensitive to the computing power and information resources of a posthuman "director".
In order to assess viability of this scenario, one should attempt to reasonably predict
those quantities. On the other hand, it remains for computer and cognitive scientists,
as well as sociologists, to answer the deep question of minimal informational cost of any
realistic simulation of human consciousness and society.
It is a pleasure to express gratitude to Prof. Nick Bostrom, whose deep insights into
all sorts of futurological and philosophical questions has largely motivated the study of
this topic. Srdjan Samurovi´c is acknowledged for invaluable technical help.
References
5
Adams, F. C. and Laughlin, G.: 1997, Rev. Mod. Phys. 69, 337.
Barrow, J. D. and Tipler, F. J.: 1986, The Anthropic Cosmological Principle (New York:
Oxford University Press).
Bekenstein, J.: 1973, Phys. Rev. D 7, 2333.
Bekenstein, J.: 1981, Phys. Rev. Lett. 46, 623.
Brillouin, L.: 1962, Science and Information Theory (New York: Academic Press).
´Cirkovi´c, M. M. and Bostrom, N.: 2000, Astrophys. Space Sci. 274, 675.
Dyson, F.: 1979, Rev. Mod. Phys. 51, 447.
Frautschi, S.: 1982, Science 217, 593.
Gibbons, G. W. and Hawking, S. W.: 1977, Phys. Rev. D 15, 2738.
Kaplinghat, M., Knox, L., and Turner, M. S.: 2000, Phys. Rev. Lett. 85, 3335.
Kaplinghat, M. and Turner, M. S.: 2001, Phys. Rev. Lett. 86 385.
Krauss, L. M. and Starkman, G.: 2000, Astrophys. J. 531, 22.
Livio, M.: 1999, Astrophys. J. 511, 429.
Lloyd, S.: 2000, Nature 406, 1047.
Peebles, P. J. E.: 1993, Principles of Physical Cosmology (Princeton: Princeton Uni-
versity Press).
Penrose, R.: 1989, The Emperor's New Mind (Oxford: Oxford University Press).
Porod, W., Grondin, R. O., Ferry, D. K. and Porod, G.: 1984, Phys. Rev. Lett. 52,
232.
Rees, M. J.: 1969, Observatory 89, 193.
Sandberg, A.: 2000, Journal of Transhumanism 5 (available at
http://transhumanist.com/volume5/Brains2.pdf).
Staggs, S. T., Jarosik, N. C., Meyer, S. S., and Wilkinson, D. T.: 1996, Astrophys. J.
473, L1.
Susskind, L.: 1995, J. Math. Phys. 36, 6377.
Tipler, F. J.: 1986, Int. J. Theor. Phys. 25, 617.
Weissman, P. R.: 1983, Astron. Astrophys. 118, 90.
6
|
0808.3771 | 2 | 0808 | 2009-01-07T09:48:47 | Chandra X-ray spectroscopy of the focused wind in the Cygnus X-1 system. I. The non-dip spectrum in the low/hard state | [
"astro-ph"
] | We present analyses of a 50 ks observation of the supergiant X-ray binary system Cygnus X-1/HDE 226868 taken with the Chandra High Energy Transmission Grating Spectrometer (HETGS). Cyg X-1 was in its spectrally hard state and the observation was performed during superior conjunction of the black hole, allowing for the spectroscopic analysis of the accreted stellar wind along the line of sight. A significant part of the observation covers X-ray dips as commonly observed for Cyg X-1 at this orbital phase, however, here we only analyze the high count rate non-dip spectrum. The full 0.5-10 keV continuum can be described by a single model consisting of a disk, a narrow and a relativistically broadened Fe Kalpha line, and a power law component, which is consistent with simultaneous RXTE broad band data. We detect absorption edges from overabundant neutral O, Ne and Fe, and absorption line series from highly ionized ions and infer column densities and Doppler shifts. With emission lines of He-like Mg XI, we detect two plasma components with velocities and densities consistent with the base of the spherical wind and a focused wind. A simple simulation of the photoionization zone suggests that large parts of the spherical wind outside of the focused stream are completely ionized, which is consistent with the low velocities (<200 km/s) observed in the absorption lines, as the position of absorbers in a spherical wind at low projected velocity is well constrained. Our observations provide input for models that couple the wind activity of HDE 226868 to the properties of the accretion flow onto the black hole. | astro-ph | astro-ph | THE ASTROPHYSICAL JOURNAL, 690:330 -- 346, 2009 January 1
Preprint typeset using LATEX style emulateapj v. 2009 January 1
doi:10.1088/0004-637X/690/1/330
CHANDRA X-RAY SPECTROSCOPY OF THE FOCUSED WIND IN THE CYGNUS X-1 SYSTEM
I. THE NONDIP SPECTRUM IN THE LOW/HARD STATE
MANFRED HANKE1,2, JÖRN WILMS1,2, MICHAEL A. NOWAK3, KATJA POTTSCHMIDT4,5,6, NORBERT S. SCHULZ3, AND JULIA C. LEE7
Received 2008 May 28; accepted 2008 August 27; published 2008 December 1
ABSTRACT
We present analyses of a 50 ks observation of the supergiant X-ray binary system Cygnus X-1/HDE 226868
taken with the Chandra High Energy Transmission Grating Spectrometer (HETGS). Cyg X-1 was in its spec-
trally hard state and the observation was performed during superior conjunction of the black hole, allowing for
the spectroscopic analysis of the accreted stellar wind along the line of sight. A significant part of the observation
covers X-ray dips as commonly observed for Cyg X-1 at this orbital phase, however, here we analyze only the
high count rate nondip spectrum. The full 0.5 -- 10 keV continuum can be described by a single model consisting
of a disk, a narrow and a relativistically broadened Fe Kα line, and a power-law component, which is consistent
with simultaneous RXTE broad band data. We detect absorption edges from overabundant neutral O, Ne, and Fe,
and absorption line series from highly ionized ions and infer column densities and Doppler shifts. With emission
lines of He-like Mg XI, we detect two plasma components with velocities and densities consistent with the base
of the spherical wind and a focused wind. A simple simulation of the photoionization zone suggests that large
parts of the spherical wind outside of the focused stream are completely ionized, which is consistent with the low
velocities (<200 km s- 1) observed in the absorption lines, as the position of absorbers in a spherical wind at low
projected velocity is well constrained. Our observations provide input for models that couple the wind activity of
HDE 226868 to the properties of the accretion flow onto the black hole.
Key words: accretion, accretion disks -- stars:
individual (HDE 226868, Cyg X-1) -- stars: winds, outflows --
techniques: spectroscopic -- X-rays: binaries
1. INTRODUCTION
Cygnus X-1 was discovered in 1964 (Bowyer et al. 1965) and
soon identified as a high-mass X-ray binary system (HMXB)
with an orbital period of 5.6 d (Murdin & Webster 1971;
Webster & Murdin 1972; Bolton 1972). It consists of the su-
pergiant O9.7 star HDE 226868 (Walborn 1973; Humphreys
1978) and a compact object, which is dynamically constrained
to be a black hole (Gies & Bolton 1982). The detailed spectro-
scopic analysis of HDE 226868 by Herrero et al. (1995) gives
a stellar mass M⋆ ≈ 18 M⊙, leading to a mass of MBH ∼10 M⊙
for the black hole, if an inclination i ≈ 35◦ is assumed. Note
that Ziółkowski (2005) derives a mass of M⋆ = (40 ± 5) M⊙
from the evolutionary state of HDE 226868, corresponding to
MBH = (20 ± 5) M⊙, while Shaposhnikov & Titarchuk (2007)
claim MBH = (8.7 ± 0.8) M⊙ from X-ray spectral-timing rela-
tions.
Cyg X-1 is usually found in one of the two states that are dis-
tinguished by the soft X-ray luminosity and spectral shape, the
timing properties, and the radio flux (see, e.g., Pottschmidt et al.
2003; Gleissner et al. 2004a,b; Wilms et al. 2006): the low/hard
state is characterized by a lower luminosity below 10 keV, a
hard Comptonization power-law spectrum (photon index Γ ∼
1.7) with a cutoff at high energies (folding energy Efold ∼
Electronic address: [email protected]
1 Dr. Karl Remeis-Sternwarte, Astronomisches Institut der Universität
Erlangen-Nürnberg, Sternwartstr. 7, 96049 Bamberg, Germany
2 Erlangen Centre for Astroparticle Physics, University of Erlangen-
Nuremberg, Erwin-Rommel-Strasse 1, 91058 Erlangen, Germany
3 MIT-CXC, NE80-6077, 77 Mass. Ave., Cambridge, MA 02139, USA
4 CRESST, University of Maryland Baltimore County, 1000 Hilltop Circle,
Baltimore, MD 21250, USA
5 NASA Goddard Space Flight Center, Astrophysics Science Division, Code
661, Greenbelt, MD 20771, USA
6 Center for Astrophysics and Space Sciences, University of California at
San Diego, La Jolla, 9500 Gilman Drive, CA 92093-0424, USA
7 Harvard University, Department of Astronomy (part of the Harvard-
Smithsonian Center for Astrophysics), 60 Garden Street, MS-6, Cambridge,
MA 02138, USA
150 keV) and strong variability of ∼30% root mean square
In
(rms). Radio emission is detected at the ∼15 mJy level.
the high/soft state, the soft X-ray spectrum is dominated by a
bright and much less variable (only few % rms) thermal disk
component, and the source is invisible in the radio. Within the
classification of Remillard & McClintock (2006), the high/soft
state of Cyg X-1 corresponds to the steep power-law state rather
than to the thermal state, as a power-law spectrum with pho-
ton index Γ ∼ 2.5 may extend up to ∼10 MeV (Zhang et al.
1997; McConnell et al. 2002; Cadolle Bel et al. 2006). Most
of the time, Cyg X-1 is found in the hard state, but transitions
to the soft state and back after a few weeks or months are
common every few years. Transitional or intermediate states
(Belloni et al. 1996) are often accompanied by radio and/or X-
ray flares. Similar to a transition to the soft state, the spec-
trum softens during these flares and the variability is reduced.
This behavior is called a "failed state transition" if the true soft
state is not reached (Pottschmidt et al. 2000, 2003). Transi-
tional states have occurred more frequently since mid-1999 than
before (Wilms et al. 2006), which might indicate changes in the
mass-accretion rate due to a slight expansion of HDE 226868
(Karitskaya et al. 2006).
HMXBs are believed to be powered by accretion from the
stellar wind. The accretion rate and therefore X-ray luminos-
ity and spectral state are thus very sensitive to the wind's de-
tailed properties such as velocity, density, and ionization. For
HDE 226868, Gies et al. (2003) found an anticorrelation be-
tween the Hα equivalent width (an indicator for the wind mass
loss rate M⋆) and the X-ray flux. Considering the photoioniza-
tion of the wind would allow for a self-consistent explanation
(see, e.g., Blondin 1994): a lower mass loss gives a lower wind
density and therefore higher degree of ionization due to the
irradiation of hard X-rays, which reduces the driving force of
HDE 226868's UV photons on the wind and results in a lower
wind velocity v, leading finally to a higher accretion rate (∝
M⋆/v4, Bondi & Hoyle 1944). However, Gies et al. (2008) find
2
HANKE ET AL.
suggestions that the photoionization and velocity of the wind
might be similar during both hard and soft states. UV obser-
vations allow the photoionization in the HDE 226868 / Cyg X-1
system to be probed: Vrtilek et al. (2008) reported P Cygni pro-
files of N V, C IV, and Si IV with weaker absorption compo-
nents at orbital phase φorb ≈ 0.5, i.e., when the black hole
is in the foreground of the supergiant. This reduced absorp-
tion, which was already found by Treves et al. (1980), is due to
the Hatchett & McCray (1977) effect, showing that those ions
become superionized by the X-ray source. Gies et al. (2008)
model the orbital variations of the UV lines assuming that the
wind of HDE 226868 is restricted to the shadow wind from
the shielded side of the stellar surface (Blondin 1994), i.e., the
Strömgren (1939) zone of Cyg X-1 extends to the donor star.
However, this assumption applies only to the spherical part of
the wind, which might therefore hardly contribute to the mass
accretion of Cyg X-1.
As HDE 226868 is close to filling its Roche lobe (Conti 1978;
Gies & Bolton 1986a,b), the wind is not spherically symmetri-
cal as for isolated stars, but strongly enhanced toward the black
hole ("focused wind"; Friend & Castor 1982). The strongest
wind absorption lines in the optical are therefore observed at the
conjunction phases (Gies et al. 2003). Similarly, X-ray absorp-
tion dips occur preferentially around φorb = 0, i.e., during su-
perior conjunction of the black hole (Bałuci´nska-Church et al.
2000). These dips are probably caused by dense, neutral
clumps, formed in the focused wind where the photoioniza-
tion is reduced, although recent analyses have also suggested
that part of the dipping activity may result from the interac-
tion of the focused wind with the edge of the accretion disk
(Poutanen et al. 2008).
The photoionization and dynamics of both the spherical and
focused winds can also be investigated with the high-resolution
grating spectrometers of the modern X-ray observatories Chan-
dra or XMM-Newton. As none of the previously reported ob-
servations of Cyg X-1 was performed at orbital phase φorb = 0
and in the hard state, when the wind is probably denser and less
ionized than in the soft state, the Chandra observation presented
here allows for the most detailed investigation of the focused
wind to date.
The remainder of this paper is organized as follows: in Sec-
tion 2, we describe our observations of Cyg X-1 with Chandra
and the Rossi X-Ray Timing Explorer (RXTE), and how we
model CCD pile-up for the Chandra-HETGS data. We present
our investigations in Section 3: after investigating the light
curves, we model the nondip continuum and analyze neutral
absorption edges and absorption lines from the highly ionized
stellar wind -- and the few emission lines from He-like ions,
which indicate two plasma components. In Section 4, we dis-
cuss models for the stellar wind and the photoionization zone.
We summarize our results after comparing them with those of
the previous Chandra observations of Cyg X-1.
2. OBSERVATION AND DATA REDUCTION
2.1. ChandraACIS-S/HETGS Observation
Cyg X-1 was observed on 2003 April 19 and 20 by the Chan-
dra X-Ray Observatory, see Table 1. An overview on all its in-
struments is given by the Proposers' Observatory Guide (CXC
2006).
In the first four months of 2003, the RXTE All-Sky
Monitor (ASM; see Doty 1994; Levine et al. 1996) showed the
source's 1.5 -- 12 keV count rate to be generally below 50 ASM-
cps (Figure 1). At the time of our Chandra observation, it was
less than 25 cps (0.33 Crab), typically indicative of the source
being in its low/hard state (Wilms et al. 2006).
Table 1
Observations of Cyg X-1
Satellite /
Instrument
Chandra
MEG±1b
HEG±1c
RXTE
PCAd
HEXTE a+be
Start
(MJD)
52748.70
(52748.70)
(52748.70)
52748.08
52748.74
52748.74
Stop
(MJD)
52749.28
(52749.28)
(52749.28)
52749.18
52748.78
52748.78
Exposurea Count Ratea
(ks)
(47.2)
16.1
16.1
· · ·
3.0
1.1
(cps)
( · · · )
2 × 27
2 × 17
· · ·
1456
2×186
Notes.
a For the Chandra data, the nondip GTIs (see Figure 4) have been used.
For RXTE, the 11th orbit was considered.
b The Chandra-MEG spectra cover ≈ 0.8 -- 6 keV (1 -- 99 % quantiles).
c The Chandra-HEG spectra cover ≈ 1 -- 7.5 keV (1 -- 99 % quantiles).
d The RXTE-PCA data from 4 to 20 keV has been used.
e The RXTE-HEXTE data from 20 to 250 keV has been used.
200
100
50
]
s
p
c
[
e
t
a
r
t
n
u
o
c
M
S
A
V
e
k
2
1
5
1
.
20
10
Jan Feb Mar Apr May Jun
Jul Aug Sep Oct Nov Dec
Date in 2003
the 1.5 -- 12 keV count rate did not exceed 25 cps.
Figure 1. Brightness of Cyg X-1 as seen by the ASM on board RXTE. During
the Chandra observation (marked by a line), the source was still in its low/hard
state:
In spite of the high
intrinsic variability, the high/soft state during June, July, and August can clearly
be distinguished, a result which is also found by spectral analysis (Wilms et al.
2006).
The High Energy Transmission Grating Spectrometer
(HETGS), containing high and medium energy gratings
(HEG/MEG; see Canizares et al. 2005) was used to disperse X-
ray spectra with the highest resolution (CXC 2006, Table 8.1):
∆λHEG = 5.5 mÅ and ∆λMEG = 11 mÅ
(1)
As only half of the spectroscopy array of the Advanced
CCD Imaging Spectrometer (ACIS-S; see Garmire et al. 2003),
namely a 512 pixel broad subarray, was operated in the timed
event (TE) mode, the six CCDs could be read out after expo-
sure times of tframe = 1.7 s and the position of each event is well
determined. Photons from the different gratings can thus easily
be distinguished due to the different dispersion directions of the
HEG and the MEG. Even in its low state, however, Cyg X-1 is
so bright that several photons may pile up in a CCD pixel during
one readout frame. Both events cannot be discriminated and are
interpreted as a single photon with larger energy. As the undis-
persed image would have been completely piled up, only 10%
of those events in a 40× 38 pixel window have been transmit-
ted in order to save telemetry capacity. The first-order spectra
are, however, only moderately affected, which can be modeled
-
CHANDRA X-RAY SPECTROSCOPY OF THE FOCUSED WIND IN THE CYG X-1 SYSTEM.
I.
3
very well (Section 2.2). The alternative to this approach would
have been to use the continuous clocking (CC) mode, where
the ACIS chips are read out continuously in 2.85 ms, but only
the position perpendicular to the readout direction can be de-
termined for every photon event. The CC mode was, however,
avoided due to difficulties in the reconstruction of HEG and
MEG spectra and other calibration issues.
The undispersed position of the source is required for the
wavelength calibration of the spectra. We redetermined it to
R.A. = 19h 58m 21.s67, δ = +35◦ 12′ 5.′′83 from the intersection
of the HEG and MEG arm and the readout streak (Ishibashi
2006). Afterward, the event lists were reduced using the stan-
dard software from the ChandraX-ray Center (CXC), CIAO 3.3
with CALDB 3.2.3.8 Exceptionally narrow extraction regions
had to be chosen as the background spectrum would other-
wise have been dominated by the dispersed extended X-ray-
scattering halo around the source (Xiang et al. 2005). The fur-
ther analysis was performed with the Interactive Spectral Inter-
pretation System (ISIS) 1.4.9 (Houck 2002).9
We use the four first-order MEG and HEG spectra (with two
dispersion directions each, called +1 and - 1 in the following)
which provide the best signal-to-noise ratio (S/N). The "second-
order spectra" are dominantly formed by piled first-order events
which reach the other order sorting window of data extraction
(defined in dispersion-energy space) when the energy of two
first-order photons accumulates. This effect is most evident for
the MEG, whose even dispersion orders are suppressed by con-
struction of the grating bars (CXC 2006).
2.2. Model for Pile-Up in Grating Observations
For the first-order spectra, pile-up causes a pure reduction of
count rate: a multiple event, i.e., the detection of more than
one photon in a CCD pixel during one readout time, which
cannot be separated, is either rejected by grade selection dur-
ing the data processing or migrates to a higher-order spectrum.
The Poisson probability for single events in a 3× 3 pixel event-
detection cell i (see Davis 2002, 2003; CXC 2005) is
(2)
where the expected number of events, Λ(i) = γ0 ·Ctot(i), is given
by the total spectral count rate, Ctot(i), at this position (in units
of counts per Å and s), and where the constant γ0 is
P1(i) = Λ(i) · exp(cid:0) - Λ(i)(cid:1) ,
γ0 = 3 ∆λ· tframe ,
(3)
where ∆λ is the resolution of the spectrometer of Equation (1),
and tframe is the frame time. We therefore describe the pile-up
in the first-order spectra with the nonlinear convolution model
simple_gpile2 in ISIS,10 which exponentially reduces the
predicted count rate C(λ) according to Equation (2):
C′(λ) = C(λ) · exp(cid:0) - γ ·Ctot(λ)(cid:1)
(4)
Here, the scale γ ≈ γ0 is left as fit parameter, and Ctot(λ) also
takes the photons into account which are dispersed in a higher
order m ≤ 3. The count rates are estimated from the correspond-
ing effective areas Am and the assumed photon flux S:
Ctot(λ) = X
3
m=1
Am(λ/m) · S(λ/m)
(5)
simple_gpile2 is based on the simple_gpile model
(CXC 2005; Nowak et al. 2008), which parameterizes the
8 See http://cxc.harvard.edu/ciao3.3/.
9 See http://space.mit.edu/cxc/isis/.
10 The ISIS/S-LANG code for simple_gpile2 is available online at
http://pulsar.sternwarte.uni-erlangen.de/hanke/X-ray/code/simple_gpile2.sl.
HEG−1
MEG+1
model
pile−up corrected data
5
.
0
2
0
.
1
0
.
5
0
0
.
2
0
.
0
]
V
e
k
/
2
m
c
/
s
/
.
h
p
[
x
u
l
F
piled data
× 0.4
1
2
Energy [keV]
5
10
Figure 2. Pile-up in the HEG (black curves) and MEG (gray curves) data:
the lower spectra, which are shifted by a factor of 0.4, show the uncorrected
data. The dashed line shows the model (free of pile-up). The MEG spec-
trum suffers from significant pile-up losses around 2 keV, where the highest
count rate is obtained. The upper spectra show the pile-up corrected spectra.
Note that we show ISIS' (model independent) flux spectra only for illustration;
simple_gpile(2) operates on the count rates predicted by a model.
strength of pile-up by the (maximum) pile-up fraction
p = 1 - exp(cid:0)- γ · max{Ctot}(cid:1), while using the parameter γ of
simple_gpile2 avoids to have a nonlocal model which de-
pends on the flux at the position of the highest pile-up.
The effect of pile-up is stronger in the MEG spectra than in
the HEG spectra due to the lower dispersion and higher effective
area of the MEG. The apparent flux reduction is most significant
near 2 keV where the spectrometer has the largest efficiency and
the highest count rates are obtained (see Figure 2). It can, how-
ever, excellently be modeled with simple_gpile2. When
fitting a spectrum, there is always a strong correlation between
the pile-up scale γ and the corresponding flux normalization
factor, e.g., the relative cross-calibration factor c introduced in
Section 3.2. The best-fit values for the pile-up scales γ found
in our data analysis (see Table 2) are only slightly larger than
expected from Equation (3) (namely γ0,MEG ±1 = 5.6×10- 2 s Å
and γ0,HEG ±1 = 2.8×10- 2 s Å), and the calibration factors c are
consistent with 1, except for the HEG- 1 spectrum, for which
both the largest γ - γ0 and c were found. Given the presence
of the γ -- c correlation, we consider the latter to be a numerical
artifact.
According to the simple_gpile2 model, the MEG+1
spectra suffer from >30 % pile-up for 6 Å ≤ λ ≤ 8.3 Å, peaking
at pMEG+1 = 45 % in the Si XIII f emission line at 6.74 Å. Except
for some emission lines, among them the Fe Kα line, the con-
tinuum pile-up fraction of the HEG spectra is below 17 %. For
the HEG+1 spectrum, the reduction is less than 10 % outside
the ranges 2.09 Å ≤ λ ≤ 4.08 Å and 6.05 Å ≤ λ ≤ 6.93 Å.
2.3. RXTEObservation
While Chandra's HETGS provides a high spectral resolution,
the energy range covered is rather limited. Within the frame-
work of our RXTE monitoring campaign, the broadband spec-
trum of Cyg X-1 was measured regularly, i.e., at least biweekly,
since 1998 (see Wilms et al. 2006, and references therein). The
observation on 2004 April 19 was extended to provide hard X-
ray data simultaneously with the Chandra observation. More
than one day was covered by 17 RXTE orbits of ∼47 min good
(cid:9)
4
HANKE ET AL.
E
T
X
E
H
−
E
T
X
R
5
2
2
0
5
1
5
7
A
C
P
−
E
T
X
R
0
0
0
1
0
0
8
0
0
6
a
r
d
n
a
h
C
0
5
1
0
0
1
0
5
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
# orbit
PCA
HEXTE
Chandra
19.2
19.4
19.6
19.8
20
20.2
day of 2003 April
Figure 3. Coverage of the simultaneous RXTE and Chandra observations. The plot shows the background-subtracted light curves with a time resolution of 64 s on a
separate y axis each. Top: RXTE-HEXTE (20 -- 250 keV). The count rates have been corrected for the detector dead time. Center: RXTE-PCA (4 -- 20 keV), normalized
by the number of active PCUs. Bottom: Chandra-HETGS (0.5 -- 12 keV, only first-order events; see Figure 4 for more details). The numbers label the RXTE orbits,
vertical lines mark the first part of the nondip spectrum (see Section 3.1 and Figure 4), which completely covers RXTE orbit 11.
orbital phase of the binary system
0.94
0.96
0.98
0.00
0.02
120
100
80
60
40
20
0
0.08
0.06
0.04
0.02
]
s
p
c
[
e
t
a
r
t
n
u
o
c
e
t
a
r
y
g
r
e
n
e
h
g
i
h
:
w
o
l
o
i
t
a
r
]
V
e
k
/
2
m
c
/
s
/
h
p
[
x
u
l
F
1
.
0
1
0
0
.
3
−
0
1
4
−
0
1
2
0
2
−
PCA
HEXTE a+b
0
10
20
30
40
50
observation time [ks]
10
Energy [keV]
100
Figure 4. Top: full 0.5 -- 12 keV band Chandra light curve. Bottom: ratio of
0.7 -- 1 keV band and 2.1 -- 7.2 keV band count rates. Absorption dips -- at first
compact, then with complex substructure -- show up with a reduced flux and
spectral hardening. Count rates >82.7 cps define the nondip data (dark).
time, interrupted by ∼49 min intervals when Cyg X-1 was not
observable due to Earth occultations or passages through the
South Atlantic Anomaly (SAA), see Figure 3.
Data in the 4 -- 20 keV range from the Proportional Counter
Array (PCA; Jahoda et al. 1996) and in the 20 -- 250 keV range
from the High Energy X-Ray Timing Experiment (HEXTE;
Gruber et al. 1996) were used. The data were extracted using
HEASOFT 6.3.1,11 following standard data screening proce-
dures as recommended by the RXTE Guest Observer Facility.
Data were only used if taken more than 30 minutes away from
the SAA. For the PCA, only data taken in the top layer of the
proportional counter were included in the final spectrum, and
no additional systematic error was added to the spectrum. Dur-
ing the observation, different sets of Proportional Counter Units
(PCUs) were operative. During the 11th orbit, extensively used
in this work, PCUs 1 and 4 were off.
3. ANALYSIS
3.1. Light Curve
The Chandra observation covers a phase range between
∼0.93 and ∼0.03 in the 5.599829 d binary orbit (Gies et al.
2003, whose epoch is HJD 2451730.449± 0.008). The top
panel of Figure 4 shows the light curve of first-order events
(MEG±1, HEG±1) in the full band accessible with Chandra-
HETGS. During several dip events, the flux is considerably
11 See http://heasarc.gsfc.nasa.gov/lheasoft/.
Figure 5. RXTE 4 -- 250 keV broadband continuum spectrum, as measured with
PCA below 20 keV and above 20 keV with HEXTE, which can be described by
a broken power-law with high-energy cutoff and a weak iron line, see Table 2.
Due to the joint fit with Chandra, the Fe line consists of narrow and a broad com-
ponent, see Figure 6. The HEXTE spectra shown are renormalized to match the
PCA flux, as the absolute calibrations of these two instruments differ by ∼15%.
reduced. The absorption dips distinguish themselves also by
spectral hardening (the bottom panel of Figure 4). We extract
a 16.1 ks nondip spectrum, which is the subject of this paper,
from all times when the total count rate exceeds 82.7 cps; these
are indicated by dark points in Figure 4. The analysis of dip
spectra will be described in a subsequent paper.
The RXTE light curve shows considerable variability as well.
Although the dips are more obviously detected with Chandra
in the soft X-ray band, similar structures are also seen with
RXTE-PCA or even -HEXTE, especially in the last RXTE or-
bits of these observations (Figure 3). We chose to infer the
nondip broadband spectrum from the RXTE data taken during
the 11th orbit, which was performed entirely during the first
part of the nondip phase. Other parts are interrupted by dips,
occultations, or have nonuniform PCU configuration.
3.2. Continuum Spectrum
The broadband spectrum of Cyg X-1 in the hard state can be
described by a broken power-law with exponential cutoff (Fig-
ure 5). Since the parameters of this phenomenological model
are correlated with those of physical Comptonization models
(see, e.g., Wilms et al. 2006, Fig. 11), we are justified in using
the aforementioned simple continuum model for this paper fo-
cusing on the spectroscopy of the wind. We describe the whole
0.5 -- 250 keV spectrum consistently with one broken power-law
model, i.e., there is no need to fit the continuum locally.
D
c
CHANDRA X-RAY SPECTROSCOPY OF THE FOCUSED WIND IN THE CYG X-1 SYSTEM.
I.
5
Fit Parameters of the Continuum Nondip Hard State Spectrum of Cyg X-1
Table 2
Parameter
Unit
Fit to the
Chandra
Joint Fit to Both the
Chandra and RXTE
Spectra Only
Spectra
1021 cm- 2
3.52 ± 0.04
5.4 ± 0.4a
keV
1.51 ± 0.01
· · ·
· · ·
· · ·
s- 1 cm- 2 keV- 1
1.15 ± 0.01
Photoabsorption
NH
(Broken) power-law
Γ1 (HETGS)
Γ1 (PCA)
Ebreak
Γ2
norm
High-energy cutoff
Ecut
Efold
Disk black body
Adisk (Equation 6)
k Tcol
Narrow iron Kα line
E0,narrow
σnarrow
Anarrow
Broad iron Kα line
E0,broad
σbroad
Abroad
keV
keV
103
keV
keV
eV (!)
10- 3 s- 1 cm- 2
keV
keV
10- 3 s- 1 cm- 2
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
1 (fixed)
1.00 ± 0.01
1.04 ± 0.01
1.01 ± 0.01
· · ·
· · ·
5.6 ± 0.1
6.0 ± 0.1
4.5 ± 0.3
3.5 ± 0.3
11745
11274b- 10c
1.04
1.60 ± 0.01
1.73 ± 0.01
9.0+0.3
- 1.5
1.50 ± 0.01
1.33 ± 0.03
24+2- 3
204 ± 9
23+17
- 12
0.25+0.03
- 0.02
6.4001+0.0004
- 0.0092
< 0.1
1.0 ± 0.3
6.4 (fixed)
0.6 ± 0.1
4.3+2.7
- 0.5
1 (fixed)
0.99 ± 0.01
1.04 ± 0.01
1.02 ± 0.01
1.18 ± 0.03
1.03+0.02
- 0.03
5.6 ± 0.1
5.9 ± 0.1
4.6 ± 0.3
3.8 ± 0.3
12180
11274b + 293b - 25c
1.4
7.4
1.06
1.8
24
Relative flux calibration (constant factor)
cMEG- 1
cMEG+1
cHEG- 1
cHEG+1
cPCA
cHEXTE
Pile-up scales
γMEG- 1
γMEG+1
γHEG- 1
γHEG+1
Fit-statistics
χ2
dof
χ2
red
10- 2 s Å
10- 2 s Å
10- 2 s Å
10- 2 s Å
Absorbed flux (pile-up corrected)
S0.5- 10 keV
F0.5- 10 keV
S0.5- 250 keV
F0.5- 250 keV
photons s- 1 cm- 2
10- 9 erg s- 1 cm- 2
photons s- 1 cm- 2
10- 9 erg s- 1 cm- 2
· · ·
· · ·
b
e
H
X
X
a
C
I
2
1
.
a
y
L
X
X
a
C
o
i
t
a
R
1
8
0
.
4
e
H
V
X
X
e
F
5
6
Energy [keV]
7
Figure 6. Chandra-HEG spectrum in the Fe line region. The model includes
both the narrow and the broad Kα emission line, the latter as required by the si-
multaneous RXTE-PCA data, and the Fe XXV and Ca XIX/XX absorption lines.
Taking pile-up into account (Section 2.2), the nondip Chan-
dra spectra alone can already be described quite well by a
weakly absorbed, relatively flat power-law spectrum with a pho-
ton index Γ1 = 1.51 ± 0.01 (Table 2). This result is consistent
with the fact that the break energy of the broadband broken
power-law spectrum is found at Ebreak = 9 keV, i.e., the Chan-
dra data are virtually entirely in the regime of the (steeper)
photon index Γ1. The onset of the exponential cutoff (with
folding energy Efold) is at Ecut = 24 keV and thus also well
above the spectral range of Chandra. As it is known that there
are cross-calibration uncertainties between Chandra and RXTE
(Kirsch et al. 2005), we use constant factors ci for the relative
flux calibration of every spectrum and also separate parame-
ters Γ1(HETGS) and Γ1(PCA), for which we find similar values
within the joint model (see Table 2). In order to describe a weak
soft excess, we add a thermal disk component, which accounts
for ∼9 % of the unabsorbed 0.5 -- 10 keV flux. The disk has a
color temperature of kTcol = 0.25 keV; similar to that found by
Makishima et al. (2008) and Bałuci´nska-Church et al. (1995),
who described the soft excess with a kT = 0.13 ± 0.02 keV
blackbody only, but the temperature Tcol may be too high by
a factor f & 1.7 (Shimura & Takahara 1995). The norm param-
eter of the diskbb model is
d/10 kpc(cid:19)2
Adisk = diskbb . norm = (cid:18) Rcol/km
· cos θ ,
(6)
where d is the distance and θ is the inclination of the disk, which
can deviate from the orbital inclination i ≈ 35◦ (Herrero et al.
1995) as the disk may be precessing with a tilt δ ≈ 37◦
(Brocksopp et al. 1999). Rcol is related to the inner radius of the
disk, Rin = η g(i) f 2· Rcol (with η ≈ 0.6- 0.7 and g(i) ≈ 0.7- 0.8;
Merloni et al. 2000).
In spite of these uncertainties and the
large statistical error of Adisk (more than 50 %), the inner disk
radius can be estimated to 1.5 RS . Rin . 10 RS if a distance
d ≈ 2.5 kpc (Ninkov et al. 1987a) and a Schwarzschild radius
RS ≈ 30 km (Herrero et al. 1995) are assumed. Thus, the disk is
consistent with extending close to the innermost stable circular
orbit (ISCO).
The good S/N afforded by the RXTE-PCA data clearly re-
veals an iron fluorescence Kα line. While the instrumental re-
sponse of the proportional counters does not allow for a res-
olution of the line profile details, i.e., whether it is narrow or
relativistically broadened, the Chandra-HEG spectra (Figure 6)
Unabsorbed luminosity, assuming d = 2.5 kpc (Ninkov et al. 1987a)
L0.5- 10 keV
L0.5- 250 keV
1037 erg s- 1
1037 erg s- 1
0.67
· · ·
0.78
3.1
Notes. Error bars indicate 90% confidence intervals for one interesting
parameter.
a In the joint fit, photoabsorption was described with tbnew model
(Wilms et al. 2000; Juett et al. 2006b) and the best-fitting abundances (see Section 3.3).
b All data have been rebinned to contain ≥ 50 counts bin- 1.
c The model contains actually more parameters, as the absorption lines have
already been included (see Tables 4 -- 6).
a
6
HANKE ET AL.
do resolve a strong narrow component at 6.4 keV. Nevertheless,
since the integrated flux of the Chandra measured line is insuf-
ficient to account for all of the PCA residuals in this region,
we include an additional broad feature to our modeling, which
is also compatible with the Chandra spectrum. Given the rel-
atively low S/N at these energies, we do not model the broad
iron line with a proper physical model such as a relativisti-
cally broadened line, but use a Gaussian with its energy fixed at
6.4 keV, as the latter is hardly constrained by the data. These re-
sults illustrate the synergy of the simultaneous observation with
complementary instruments, as the combination of narrow and
broad line could only be revealed by the analysis with a joint
model.
Our global model for the continuum spectrum now enables us
to address the features of the high-resolution Chandra spectra,
which is the topic in the remainder of this section.
3.3. Neutral Absorption
)
.
u
.
f
5
−
0
1
(
x
u
l
F
)
n
i
b
r
e
p
(
s
t
n
u
o
C
2
1
0
2
0
1
2
0
2
−
17
Fe L2
Fe L3
17.5
Wavelength [Å]
18
Absorbing columns can be measured most accurately from
the discrete edges in high-resolution spectra at the ionization
thresholds. We detect the most prominent L-shell absorption
edge of iron and the K-shell absorption edges of oxygen and
neon. Juett et al. (2004, 2006a) have inferred the fine struc-
ture at those edges from earlier Chandra-HETGS observations
of Cyg X-1 and other bright X-ray binaries: the Fe L2 and L3
edges, due to the ionization of a 2p1/2 or a 2p3/2 electron, re-
spectively, are separately detectable at 17.2 Å and 17.5 Å. The
O K-edge at 22.8 Å is accompanied by (1s→2p) Kα and higher
(1s→ np) resonance absorption lines. The Kα line occurs at
23.5 Å for neutral O I and at lower wavelengths for ionized oxy-
gen. In the case of neon, neutral atoms have closed L-shells,
such that Ne I only shows a (1s→3p) Kβ absorption line close
to the K-edge at 14.3 Å, while ionized neon also shows Kα ab-
sorption lines, e.g., Ne II at 14.6 Å and Ne III at 14.5 Å. Im-
proved modeling of the neutral absorption that takes these fea-
tures into account has recently been included in the photoab-
sorption model tbnew12 (Juett et al. 2006b), an extension of
the commonly used tbvarabs model (Wilms et al. 2000).
As part of the spectral model for the whole continuum, the
tbnew model can be used to describe the absorption edges
detected with the Chandra observation of Cyg X-1 discussed
in this paper. Figure 7 shows the Fe L-edges requiring a
blueshift by ∆λ/λ0 · c = (540 ± 230) km s- 1 of the tbnew
model, which relies on the cross sections of metallic iron mea-
sured by Kortright & Kim (2000). Unlike Schulz et al. (2002),
Miller et al. (2005) and Juett et al. (2006a) have also found that
the Fe L3 edge requires a small shift;
their mean position
of maximum optical depth is λFe L3 = 17.498 Å, but our value
λFe L3 = (17.469± 0.014) Å is still lower. The shift could be
caused by the Doppler effect due to a moving absorber, by a
modified ionization threshold due to chemical bonds, or ioniza-
tion of the iron atoms (van Aken & Liebscher 2002). In an anal-
ysis of the Cyg X-1 high/soft and low/hard state, focused on this
spectral region, J. Lee et al. (2008, in preparation) find that the
Fe L-edges here can likely be modeled by a heterogeneous com-
bination of gas and condensed matter of iron in combination
with oxygen local to the source environment. If, as suggested
by these authors, the magnitude of the shift is due to molecules
and/or dust, this shift is one identifying signature of the compo-
sition and charge state of the condensed state material. Such
direct Chandra X-ray spectroscopic detection of dust via its
12 The code for tbnew is available online at http://pulsar.sternwarte.uni-
erlangen.de/wilms/research/tbabs/ for beta testing and will soon be released.
Figure 7. Spectral region around the Fe L-edges. The top panel shows the
model flux in units of 10- 5 ph. s- 1 cm- 2 Å- 1 (solid line). The dashed lines de-
scribe the unshifted model, see the text. The second panel shows the count rate
of all the four HETGS spectra (MEG±1, HEG±1), which have been combined
with a resolution of 10 mÅ, and the folded model. The bottom panel displays
the residuals χ = (data -- model)/error.
O I 1s−2p
O II
O III
)
.
u
.
f
5
−
0
1
(
x
u
l
F
)
n
i
b
r
e
p
(
s
t
n
u
o
C
5
2
1
5
.
0
0
2
0
1
0
1
0
1
−
22
24
Wavelength [Å]
26
Figure 8. Same as in Figure 7, but for the spectral region around the O K-edge.
Both MEG±1 spectra have been combined with a resolution of 0.1 Å.
without Ne Ne I 1s−3p
Ne III 1s−2p
Ne II 1s−2p
V III
F e X
V III
F e X
V III
F e X
F e X
V III
)
.
u
.
f
5
−
0
1
(
x
u
l
F
)
n
i
b
r
e
p
(
s
t
n
u
o
C
0
1
5
2
1
0
5
0
2
2
0
2
−
14.2
14.4
Wavelength [Å]
14.6
Figure 9. Same as in Figure 7, but for the spectral region around the Ne K-
edge. All spectra have been combined with a resolution of 5 mÅ. The dotted
lines show the absorbed continuum model without additional absorption lines.
The light dashed line shows the model without the absorption of Ne.
c
c
c
Element
12 + log AISM
element
a
Nelement/NFe
Aelement/AISM
element
b
O
Ne
Na
Mg
Al
Si
S
Ar
Ca
Cr
Fe
8.69
7.94
6.16
7.40
6.33
7.27
7.09
6.41
6.20
5.51
7.43
15.4 ± 0.3
3.91 ± 0.18
1.26 ± 0.15
2.16 ± 0.13
< 0.13
1.45 ± 0.17
0.6 ± 0.2
< 0.06
< 0.05
< 0.18
1
1.45 ± 0.01
2.07+0.04
- 0.03
40 ± 2
4.0 ± 0.1
< 1.8
3.8 ± 0.3
2.4+0.5
- 0.3
< 1.0
< 1.3
1.3+2.7
- 1.3
1.75 ± 0.03
c
Nelement
(1017 cm- 2)
38.1+0.4
- 0.3
9.6 ± 0.2
3.1 ± 0.2
5.3 ± 0.2
< 0.2
3.8 ± 0.3
1.6+0.3
- 0.2
< 0.1
< 0.1
0.02+0.05
- 0.02
2.52 ± 0.04
d
Nelement
(1017 cm- 2)
39.2 ± 2.3
9.43 ± 0.32
· · ·
3.7 ± 1.3
· · ·
2.3 ± 1.8
· · ·
· · ·
· · ·
· · ·
e
Nelement
(1017 cm- 2)
63 ± 14
7.1 ± 1.0
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
ObsID 3407
(3.1×1037 erg s- 1)
-- -- -- -- --
e
Nelement
(1017 cm- 2)
24 ± 3
7.4+0.7
- 0.3
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
ObsID 3724
(soft state)
-- -- -- -- --
e
Nelement
(1017 cm- 2)
29 ± 3
8.6 ± 0.7
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
CHANDRA X-RAY SPECTROSCOPY OF THE FOCUSED WIND IN THE CYG X-1 SYSTEM.
I.
7
Column Density N and Abundance A = N/NH of Neutral Absorbers Detected Along the Line of Sight Toward Cyg X-1
Table 3
This work
Schulz et al. (2002)
-- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- --
Juett et al. (2004, 2006a)
Analysis by
ObsID 107
ObsID 107
(L0.5- 10 kev = 1.6×1037 erg s- 1)
-- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- --
(L0.5- 10 kev = 0.7×1037 erg s- 1)
ObsID 3814
1.5 ± 0.1
1.47+0.25
- 0.19
1.06+0.07
- 0.11
0.96 ± 0.09
Notes.
a We use elemental abundances in the interstellar medium, AISM
b The relative elemental abundance, Aelement/AISM
c Nelement is the product of Aelement and NH (see Table 2).
d Schulz et al. (2002) use the cross sections from Verner et al. (1993), and Kortright & Kim (2000) for Fe.
e Juett et al. (2004, 2006a) use the cross section of Gorczyca & McLaughlin (2000) for O, Gorczyca (2000) for Ne, and Kortright & Kim (2000) for Fe.
element according to Wilms et al. (2000), i.e., xspec_abund("wilm"); in ISIS.
element, is a parameter of the tbnew model.
associated edge structure was first suggested for observations
of the Fe L-edge in the active galactic nucleus MCG- 6-30-15
(Lee et al. 2001) and for the observations of the Si K-edge in
the microquasar GRS 1915+105 (Lee et al. 2002). The latter
study associated the observed Si K-edge structure with SiO2,
although the origin -- source environment or Chandra CCD gate
structure -- was unclear in this case.
The blueshift of (100± 130) kms- 1 which has been deter-
mined for all other absorption edges is consistent with zero. The
O K-edge can only be seen in the MEG spectra after heavy re-
binning (Figure 8). Nevertheless, the Kα resonance absorption
line of O I is clearly detected. The region around the Ne edge
(Figure 9) is dominated by Fe XVIII absorption lines, possibly
blending with the Kα absorption lines of Ne II and Ne III. No
other strong edges are clearly visible in the spectrum. The Na
K-edge (at 11.5 Å; Verner & Yakovlev 1995) blends with an ab-
sorption line due to the 2s→3p excitation of Fe XXII. The Mg
K-edge (at 9.5 Å) blends with the Ne X Ly δ absorption line.
The Si K-edge (at 6.7 Å) is strongly affected by pile-up and
blends with the Mg XII Ly β absorption line. The S K-edge (at
5.0 Å) is relatively weak. Neutral absorption from these ele-
ments is nevertheless required within the tbnew model.
The results for the individual abundances, Ai = Ni/NH, and re-
sulting column densities, Ni, are presented in Table 3. The alpha
process elements O, Ne, Mg, Si, and S are overabundant with
respect to the interstellar medium (ISM) abundances as summa-
rized by Wilms et al. (2000) and therefore suggest an origin in
the system itself. The total column densities are also compared
with the values obtained by Schulz et al. (2002) and Juett et al.
(2004, 2006a) from other Chandra observations of Cyg X-1.
Table 3 includes the corresponding source luminosities if they
are reported in the literature (see also Section 4.4). The X-ray
flux was highest during the soft state observation with the ob-
servation identification (ObsID) 3724. The column densities
confirm the conjecture of Juett et al. (2004) that a higher (soft)
X-ray flux ionizes material local to the Cyg X-1 system and re-
duces the neutral abundances.
The inferred hydrogen column density NH (see Table 2) is in
very good agreement with that from ASCA observations dur-
ing the soft state in 1996, namely NH = (5.3± 0.2)× 1021 cm- 2
(Dotani et al. 1997), and also with NH = 6.2×1021cm- 2 obtained
from two other different Chandra observations (Schulz et al.
2002; Miller et al. 2002, see also Section 4.4). We note that
the large column density toward Cyg X-1 found by many on-
line tools, (7.2- 7.8)×1021cm- 2, is obtained from a coarse grid
-- with (0.675 deg)2 pixel size -- of NH measurements at 21 cm
(Kalberla et al. 2005), which does not resolve the strong varia-
tions of NH in the region around Cyg X-1 (Russell et al. 2007).
3.4. Absorption Lines of H- and He-Like Ions
The high-resolution spectra reveal a large number of absorp-
tion lines of highly ionized ions. The 1.5 -- 15 Å range is shown
in Figure 10 as the ratio between the data and the continuum-
model. As the line profiles are not fully resolved, we model
each line l with a Gaussian profile Gl. In terms of the contin-
uum flux model, Fcont, the global model reads
F(λ) = e- τ (λ) · Fcont(λ) = "1 +Xl
Gl(λ)# · Fcont(λ) . (7)
m
a
r
k
t
h
e
c
o
m
m
o
n
l
i
n
e
s
'
r
e
s
t
w
a
v
e
l
e
n
g
t
h
s
.
)
I
r
o
n
L
-
s
h
e
l
l
t
r
a
n
s
i
t
i
o
n
s
(
l
i
n
e
s
o
f
F
e
<
X
X
V
)
a
r
e
s
h
o
w
n
i
n
b
l
u
e
i
n
t
h
e
o
n
l
i
n
e
j
o
u
r
n
a
l
.
r
e
s
o
l
u
t
i
o
n
o
f
1
0
m
Å
,
a
n
d
a
l
l
M
E
G
±
1
a
n
d
H
E
G
±
1
s
p
e
c
t
r
a
h
a
v
e
b
e
e
n
c
o
m
b
i
n
e
d
.
T
h
e
G
a
u
s
s
i
a
n
l
i
n
e
fi
t
s
a
n
d
i
d
e
n
t
i
fi
c
a
t
i
o
n
s
a
r
e
s
h
o
w
n
a
s
w
e
l
l
.
(
T
h
e
l
a
b
e
l
s
Ratio
1
0.5
1.5
2
0.5
Fe XIX
Fe XIX
Fe XIX
Fe XIXFe X
VII
Fe X
VIIIFe X
VIII
1
1
.
6
1
1
.
8
1
2
1
2
.
2
Fe X
VIII
1
2
.
4
Fe X
Fe X
VIII
VIII
Fe XIX
O Ly d
Fe X
VII
1
2
.
6
1
2
.
8
1
3
O Ly g
1
3
.
2
Fe X
VII
Fe X
VII
1
3
.
4
F
i
g
u
r
e
1
0
.
C
h
a
n
d
r
a
(
n
o
n
d
i
p
)
s
p
e
c
t
r
u
m
o
f
C
y
g
X
-
1
,
s
h
o
w
n
a
s
r
a
t
i
o
o
f
d
a
t
a
a
n
d
c
o
n
t
i
n
u
u
m
m
o
d
e
l
(
T
a
b
l
e
2
)
.
F
o
r
v
i
s
u
a
l
c
l
a
r
i
t
y
,
t
h
e
d
a
t
a
h
a
v
e
b
e
e
n
r
e
b
i
n
n
e
d
t
o
a
W
a
v
e
l
e
n
g
t
h
[
Å
]
1
3
6
.
1
3
8
.
1
4
1
4
.
2
1
4
4
.
1
4
.
6
1
4
.
8
1
5
1
5
.
2
1
5
.
4
2
Ratio
1
1.5
Fe X
XIIFe X
Ne He b
VIII
Fe X
X+X
XII
Fe X
XIIFe X
XI
Fe X
Ne Ly a
VII
Fe X
Fe X
VII
Fe X
XIFe X
XII
XI
Fe X
X
Fe X
XFe X
X
Fe X
XFe XIX
Fe X
X
Fe X
XI
Fe XIX
Ne He a
Fe XIXFe XIX
Ratio
1
1.5
Ne Ly d
Ne Ly g
Fe X
XNa Ly a
Fe X
Fe XIX
VII−X
X
Ne Ly b
Fe X
Fe X
Ne He e
VII+X
XIV
XIV
Fe XIX
Fe X
XII+X
Fe X
VII+X
Na He i
Ne He g
XIII
XIII
7
6
.
7
8
.
8
8
.
2
8
.
4
8
.
6
8
.
8
9
Na He f
9
.
2
Fe X
VIII
9
.
4
Fe X
Fe X
XII
XII
9
.
6
9
.
8
1
0
1
0
.
2
1
0
.
4
1
0
.
6
1
0
.
8
1
1
1
1
.
2
1
1
.
4
Ratio
1
1.2
0.8
1.4
0.8
Ratio
1
1.2
1.4
0.8
Ratio
1
1.2
1.4
0.8
5
6
.
5
8
.
Si He b
3
6
.
3
8
.
6
4
Al Ly b
Si Ly a
6
.
2
6
.
4
Al He a
Al He iMg He b
XII
Fe X
Fe X
XIV
Fe X
XIII
Mg Ly a
4
.
2
4
.
4
4
.
6
4
.
8
Fe X
XI
6
.
6
6
.
8
Mg Ly d
Al He b
Si He a
Si He fFe X
Mg Ly g
XIV
Fe X
XII
7
5
Mg He a
Fe X
X+X
Mg He i
XI
Ne Ly zMg He fNe Ly e
7
.
2
7
.
4
Ne Ly d
Mg Ly b
Al Ly a
Mg He d
Mg He g
5
.
2
5
.
4
1
6
.
1
.
8
2
2
.
2
2
.
4
2
.
6
2
.
8
3
3
.
2
3
.
4
S Ly d
Ar Ly a
S Ly g
Ar He a
S He g
S He b
S Ly a
Si Ly g
S He aS He iS He f
Si Ly b
Si He g
8
H
A
N
K
E
E
T
A
L
.
Ratio
1
1.2
1.4
Fe Ly b
?
Fe He b
?
Fe Ly a
Fe He a
?
Fe Ka
Ca He b
Ca Ly a
Ar Ly bCa He i
Ar He g
Ar He b
CHANDRA X-RAY SPECTROSCOPY OF THE FOCUSED WIND IN THE CYG X-1 SYSTEM.
I.
9
Overview on the Detected Lines from H- and He-Like Ions: Theoretical Rest Wavelengths in Å
Table 4
Transition
Hydrogen-like (1 electron)
1s (2S1/2) → 2p (2P3/2,1/2)
1s (2S1/2) → 3p (2P3/2,1/2)
1s (2S1/2) → 4p (2P3/2,1/2)
1s (2S1/2) → 5p (2P3/2,1/2)
1s (2S1/2) → 6p (2P3/2,1/2)
1s (2S1/2) → 7p (2P3/2,1/2)
1s (2S1/2) → 8p (2P3/2,1/2)
1s (2S1/2) → 9p (2P3/2,1/2)
O
VIII
18.97
16.01
15.18
14.82
(14.63)
(14.52)
(14.45)
(14.41)
Ne
X
12.13
10.24
9.71
9.48
9.36
9.29
9.25
( 9.22)
Na
XI
10.03
8.46
8.02
7.83
( 7.73)
( 7.68)
( 7.64)
( 7.61)
Mg
XII
8.42
7.11
(6.74)
6.58
6.50
6.45
(6.42)
(6.40)
Al
XIII
7.17
6.05
5.74
(5.60)
(5.53)
(5.49)
(5.47)
(5.45)
Si
XIV
6.18
5.22
4.95
(4.83)
(4.77)
(4.73)
(4.71)
(4.70)
S
XVI
4.73
3.99
3.78
3.70
(3.65)
(3.62)
(3.60)
(3.59)
Ar
XVIII
3.73
3.15
2.99
(2.92)
(2.88)
(2.86)
(2.85)
(2.84)
Ca
XX
3.02
(2.55)
2.42
(2.36)
(2.33)
(2.31)
(2.30)
(2.29)
Fe
XXVI
(1.78)
1.50
1.43
(1.39)
(1.37)
(1.36)
(1.36)
(1.35)
Ni
XXVIII
1.53
1.29
1.23
(1.20)
1s (2S1/2) → ∞
(14.23)
( 9.10)
( 7.52)
6.32
5.38
(4.64)
(3.55)
(2.80)
(2.27)
1.34
(1.15)
Ly α
Ly β
Ly γ
Ly δ
Ly ǫ
Ly ζ
Ly η
Ly θ
· · ·
limit
Transition
Helium-like (2 electrons)
f [em.] 1s2 (1S0) ← 1s2s (3S1)
i [em.] 1s2 (1S0) ← 1s2p (3P1,2)
r ≡ He α 1s2 (1S0) → 1s2p (1P1)
1s2 (1S0) → 1s3p (1P1)
1s2 (1S0) → 1s4p (1P1)
1s2 (1S0) → 1s5p (1P1)
1s2 (1S0) → 1s6p (1P1)
1s2 (1S0) → 1s7p (1P1)
1s2 (1S0) → 1s8p (1P1)
He β
He γ
He δ
He ǫ
He ζ
He η
· · ·
limit
O
VII
22.10
21.80
21.60
18.63
(17.77)
(17.40)
(17.20)
(17.09)
(17.01)
Ne
IX
(13.70)
(13.55)
13.45
11.54
11.00
10.77
10.64
10.56
(10.51)
Na
X
11.19
11.08
11.00
9.43
8.98
8.79
( 8.69)
( 8.63)
( 8.59)
Mg
XI
9.31
9.23
9.17
7.85
7.47
7.31
(7.22)
(7.17)
(7.14)
Al
XII
7.87
7.81
7.76
6.64
6.31
(6.18)
(6.10)
(6.06)
(6.03)
Si
XIII
6.74
(6.69)
6.65
5.68
5.40
(5.29)
5.22
(5.19)
(5.16)
S
XV
5.10
5.07
5.04
4.30
4.09
4.00
3.95
(3.92)
(3.90)
Ar
XVII
(3.99)
3.97
3.95
3.37
3.20
(3.13)
(3.10)
Ca
XIX
(3.21)
3.19
3.18
2.71
(2.57)
(2.51)
Fe
XXV
(1.87)
(1.86)
1.85
1.57
1.50
1.46
Ni
XXVII
1.60
1.59
(1.35)
(1.28)
(1.25)
1s2 (1S0) → 1s ∞
16.77
10.37
8.46
(7.04)
5.94
5.09
(3.85)
(3.01)
2.42
1.40
(1.20)
Notes. Lines with (wavelengths in parentheses) are not detected in our Chandra-HETGS observation of Cyg X-1, while lines indicated with bold wavelengths are
clearly detected and those with underlined wavelengths are detected as two components. The wavelengths of the lines are taken from the CXC atomic database
ATOMDB and the table of Verner et al. (1996), those of the series limits (= K-ionization thresholds) are from Verner & Yakovlev (1995).
Results from the Detected Absorption Lines from H- and He-Like Ions: Velocity Shifts v = (λ - λ0)/λ0 · c in km s- 1
Table 5
Ar
XVIII
- 322+712
- 617
- 750+710
- 451
Ca
XX
164+838
- 402
· · ·
· · ·
· · ·
· · ·
· · ·
Fe
XXVI
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
H-like
Ly α
Ly β
Ly γ
Ly δ
Ly ǫ
Ly series
He-like
He α
He β
He γ
He δ
He ǫ
O
VIII
- 718+1716
- 281
- 90+187
- 185
- 39+85
- 81
- 109+1109
- 198
· · ·
- 41+67
- 72
O
VII
· · ·
507+492
- 1020
· · ·
· · ·
· · ·
He series
- 291 ± (>1000)
Ne
X
- 128+35
- 27
- 72+50
- 47
7+64
- 64
- 192+46
- 46
288+213
- 177
- 97+28
- 20
Ne
IX
- 203+85
- 91
- 36+47
- 0
- 507+106
- 123
· · ·
- 93+54
- 59
- 158+35
- 56
Na
XI
284+81
- 81
· · ·
· · ·
· · ·
· · ·
207+104
- 101
Na
X
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
Mg
XII
- 34+26
- 34
17+27
- 73
· · ·
3+922
- 1401
· · ·
- 38+27
- 28
Mg
XI
- 71+35
- 49
- 65+44
- 0
77+81
- 71
· · ·
· · ·
- 72+25
- 28
Al
XIII
- 171+88
- 104
344+431
- 448
· · ·
· · ·
· · ·
- 183+140
- 117
Al
XII
286+187
- 574
59+54
- 46
· · ·
· · ·
· · ·
135+205
- 234
Si
XIV
- 60+32
- 33
75+263
- 238
308+333
- 193
· · ·
· · ·
- 54+33
- 32
Si
XIII
- 30+23
- 46
18+165
- 195
- 19+320
- 234
· · ·
· · ·
- 86+26
- 41
S
XVI
- 43+114
- 95
· · ·
- 347+393
- 399
420+536
- 489
· · ·
- 90+74
- 73
S
XV
84+127
- 138
609+785
- 1556
702+478
- 658
· · ·
· · ·
47+37
- 52
· · ·
· · ·
· · ·
- 327+41
- 128
Ar
XVII
- 142+250
- 233
72+440
- 306
655+283
- 470
· · ·
· · ·
2+8- 12
Ca
XIX
1124+98
- 1069
- 74+616
- 422
· · ·
· · ·
· · ·
· · ·
Fe
XXV
- 116+501
- 531
· · ·
· · ·
· · ·
· · ·
- 21+0- 379
Notes. A negative velocity indicates a blue shift, due to the absorbing material moving toward the observer. Rows labeled with "Ly/He series" show the results from
modeling the complete absorption line series of the corresponding ion at once with a single physical model (Section 3.5).
10
HANKE ET AL.
Results from the Detected Absorption Lines from H- and He-Like Ions: Column Densities in 1016 cm- 2
Table 6
H-like
Ly α
Ly β
Ly γ
Ly δ
Ly series
He-like
He α
He β
He γ
He series
O
VIII
3+3- 1
15+3- 5
20+5- 8
11+13
- 20
36+18
- 12
O
VII
· · ·
4+3- 6
· · ·
0+4- 0
Ne
X
3.5+0.2
- 0.3
10.8+1.2
- 1.3
51+5- 4
82+9- 7
42+10
- 4
Ne
IX
1.5 ± 0.2
2.8+0.8
- 0.4
3.0 ± 1.5
6.3+1.3
- 1.6
Na
XI
· · ·
· · ·
· · ·
· · ·
2.3+0.4
- 0.5
Na
X
· · ·
· · ·
· · ·
· · ·
Mg
XII
5.2+0.2
- 0.3
8.0+1.6
- 1.8
· · ·
8+18
- 8
Al
XIII
1.8 ± 0.5
3+3- 2
· · ·
· · ·
Si
XIV
7.1+0.3
- 0.4
18+5- 6
19 ± 13
· · ·
7.2 ± 0.6
1.4 ± 0.5
10.1 ± 0.8
Mg
XI
1.57+0.13
- 0.15
5.6 ± 0.9
8 ± 2
5+2- 1
Al
XII
0.5+0.4
- 0.1
3.8+1.2
- 0.6
· · ·
0.8+0.2
- 0.4
Si
XIII
1.94+0.11
- 0.18
9+2- 3
8+5- 6
12 ± 3
S
XVI
5.0+0.6
- 1.1
· · ·
18+19
- 15
29+34
- 29
15+12
- 5
S
XV
2.3+0.4
- 0.5
16+6- 9
23+11
- 14
15+2- 3
Ar
XVIII
2.2+1.2
- 1.8
7+9- 7
· · ·
· · ·
11+1- 6
Ar
XVII
1.3 ± 0.6
4 ± 4
21+11
- 10
8+1- 4
Ca
XX
2 ± 2
· · ·
· · ·
· · ·
· · ·
Ca
XIX
1.0+0.9
- 0.8
11+5- 7
· · ·
· · ·
Fe
XXVI
· · ·
· · ·
· · ·
· · ·
· · ·
Fe
XXV
15+3- 4
· · ·
· · ·
146+80
- 0
Notes. The column densities for the single lines have been calculated using Equation (11), assuming that the line is on the linear part of the curve of growth, which
underestimates the column density for saturated lines. Rows labeled with "Ly/He series" show the results from modeling complete absorption line series (Section 3.5).
From a Gaussian's centroid wavelength λ and the rest wave-
length λ0 of the identified line, the radial velocity
v = (λ - λ0)/λ0 · c
(8)
of the corresponding absorber can be inferred. With the defini-
tion in Equation (7), the norm of Gl is just the equivalent width:
Wλ,l
:= Z (cid:2)1 - Fl(λ)/Fcont(λ)(cid:3) dλ = Z Gl(λ) dλ
(9)
Wλ,l is related to the absorber's column density, as a bound --
bound transition i → j (with the rest frequency ν0 and oscillator
strength fi j) in an absorbing plasma with column density Ni cre-
ates the following line profile (see Mihalas 1978, Section9-2):
e- τ (ν) = exp(cid:26)- Ni
fi j √π e2
me c ∆νD
H(cid:18) Γ
4π ∆νD
,
ν - ν0
∆νD (cid:19)(cid:27) (10)
Assuming pure radiation damping, the damping constant Γ
equals the Einstein coefficient A ji. The Doppler broadening
∆νD = ν0 · ξ0/c is given by ξ0
turb, i.e., is due to
the thermal and turbulent velocities of the plasma. For optically
thin lines (with τ(ν) ≪ 1), the equivalent width is independent
of ∆νD, such that the absorbing column density can be inferred
(Spitzer 1978, Eq. 3-48; Mihalas 1978, §10-3):
2 = 2kT /mion + v2
Ni =
mc2 ·Wλ
π e2 · fi j λ2
0
=
1.13×1017 cm- 2
fi j
- 2
Å(cid:19)
·(cid:18) λ0
·(cid:18) Wλ
mÅ(cid:19) (11)
If the lines are, however, saturated, Wλ depends on ∆νD as well,
and one has to construct the full "curve of growth" with several
lines from a common ground state i in order to constrain Ni (see,
e.g., Kotani et al. 2000).
We have therefore performed a systematic analysis of absorp-
tion line series: H-like ions are detected by their 1s → n p Ly-
man series and He-like ions by their 1s2 → 1s n p resonance
absorption series, see Table 4. For those lines that are clearly
detected and not obviously affected by blends, the measured ve-
locity shifts (Equation 8) are shown in Table 5. Most of the lines
are detected at rather low projected velocity (v <200 km s- 1).
Note that the systemic velocity of Cyg X-1/HDE 226868 is
(- 7.0± 0.5) km s- 1 (Gies et al. 2003), and that the radial veloc-
ity of both the supergiant and the black hole vanishes at or-
bital phase φorb = 0, while the radial component of the focused
stream should be maximal (likely to be up to 720 km s- 1, see
Section 4.1). The column densities in Table 6 are calculated us-
ing Equation (11), assuming that the line is on the linear part
of the curve of growth. As the strongest lines are, however,
often saturated, Equation (11) predicts too small column densi-
ties from them. For weak lines, however, the equivalent width
is most likely to be overestimated such that the quoted values
may rather be upper limits. The properties of the lines (Einstein
A-coefficients and quantum multiplicities, which determine the
oscillator strength, as well as the rest wavelengths) are taken
from Verner et al. (1996) and ATOMDB13 version 1.3.1.
3.5. Line Series of H-/He-Like and Fe L-Shell Ions
As an alternative to the curve of growth, we chose to develop
a model which implements the expected line profiles (Equa-
tion 10) directly for all transitions of a series from a common
ground state i. The model contains Ni, ξ0, and the systemic
shift velocity v (Equation 8) as fit parameters, and thus avoids
the use of equivalent widths at all. This approach allows for
a systematic treatment of the iron L-shell transitions as well,
which are often blended with other lines such that the different
contributions to a line's equivalent width can hardly be sepa-
rated when only single Gaussians are used. As an example,
Figure 11 shows the Fe XIX complex between 12.8 Å and 14 Å.
Lines from different Fe XIX transitions overlap, and so does the
strong Ne IX r-line. Furthermore, the absorption features are of-
ten rather weak and no prominent lines can be fitted, whereas
the line series model can still be applied. Although a disad-
vantage of this approach is the larger computational effort, it
usually allows us to constrain the parameters of a line series
more tightly.
The model with physical absorption line series fits the data
hardly worse than the model with single Gaussian lines: χ2 of
12812 instead of 12180 before (see Table 2) is obtained. The
results are presented in the last row for the whole Ly/He series
of Tables 5 and 6 for the H-/He-like ions, and in Table 7 for the
13 See http://cxc.harvard.edu/atomdb/.
CHANDRA X-RAY SPECTROSCOPY OF THE FOCUSED WIND IN THE CYG X-1 SYSTEM.
I.
11
e
l
i
f
o
r
p
e
n
i
l
d
e
z
i
l
a
m
r
o
N
1
5
0
.
0
13
r
I
X
g
M
r1
(r2)
XI ?
X ?
Fe X
Fe X
1
5
.
0
e
t
a
r
t
n
u
o
c
d
e
z
i
l
a
m
r
o
N
i1
i2
f1
f2
G
E
H
G
E
M
e I X f)
( N
i
I
X
g
M
h
y
L
e
N
y
L
e
N
f
I
X
g
M
14
9.15
9.2
9.25
9.3
9.35
e I X i)
N
e I X r
( N
13.5
Wavelength [Å]
Figure 11. Absorption line series of Fe XIX between 12.8 Å and 14 Å. The
Ne IX r line which blends with the complex at 13.4 -- 13.6 Å is also shown, and
the expected positions of the Ne IX i and f emission lines are indicated.
Parameters of the Absorption Line Series for Fe L-Shell Ions
Table 7
Ion
Fe XXIV
Fe XXIII
Fe XXII
Fe XXI
Fe XX
Fe XIX
Fe XVIII
Fe XVII
v
(km s- 1)
21+31
- 30
120+68
- 94
- 27+33
- 35
- 156+30
- 29
- 283+33
- 48
- 116+46
- 41
- 60+51
- 54
- 277+56
- 42
N(ion)
(1016 cm- 2)
4.9+0.7
- 1.0
2.2+0.4
- 0.2
2.3+0.2
- 0.4
2.5+0.2
- 0.4
3.0+0.3
- 0.2
3.1+0.2
- 0.3
1.8+0.3
- 0.2
1.4 ± 0.3
ξ0
(km s- 1)
82+21
- 22
410+68
- 140
193+61
- 64
260+59
- 68
333+57
- 92
251+80
- 68
203+72
- 56
78+26
- 25
Fe L-shell ions. The column densities inferred from the series
model are generally in good agreement with the values derived
from the single Gaussian fits for the higher transitions of H-
or He-like ions (Table 6), while the α -- sometimes even β --
lines are saturated. These measurements can be used as input
for wind or photoionization models, although a few columns
are rather badly constrained if the thermal velocity ξ0 is left as
a free parameter. As the line profiles are not resolved, there
is a degeneracy between Doppler-broadened lines and narrow,
but saturated lines. The notable fact that no strong wavelength
shifts are observed is, however, independent of this degener-
acy: almost all line series are consistent with a velocity between
- 200 km s- 1 and +200 km s- 1. Lower ionization stages of the
same element are usually seen at higher blueshift, like in the
sequence Fe XXIII -- XXII -- XXI -- XX.
3.6. Emission Lines from He-Like Ions
The transitions between the 1s2 ground state and the 1s2s or
1s2p excited states of He-like ions lead to the triplet of for-
bidden (f), intercombination (i), and resonance (r) line, see Ta-
ble 4. These lines provide a density and temperature diagnos-
tics of an emitting plasma via the ratios R(ne) := F(f) / F(i)
and G(Te) := [F(i) + F(f)] / F(r) of the fluxes F in the r-, i-,
and f-line (see, e.g., Gabriel & Jordan 1969; Porquet & Dubau
2000). In this observation, the dipole-allowed resonance transi-
tions are seen as absorption lines, as an absorbing plasma is
Wavelength [Å]
Figure 12. Mg XI triplet of resonance (r), intercombination (i), and forbidden
(f) lines in the Chandra-HETGS spectrum. The MEG data are shown with
an offset of - 0.5. While the ion/line labels indicate the rest wavelengths, the
abbreviations rn, in, and fn label the actual line positions (see Table 8).
Parameters of the Lines of the Mg XI Triplet
Table 8
Line
Component
Mg XI r
Mg XI r?
Mg XI i
Mg XI i
Mg XI f
Mg XI f
r1
(r2)
i1
i2
f1
f2
λ
(Å)
9.167+0.001
- 0.002
9.190+0.004
- 0.003
9.232 ± 0.003
9.261+0.002
- 0.003
9.311+0.004
- 0.001
9.328+0.009
- 0.003
v
(km s- 1)
- 70+35
- 49
`700+120
- 90 ´
14+107
- 105
968+52
- 110
- 110+130
- 34
454+293
- 111
Wλ
(mÅ)
- 8.6+0.8
- 0.7
- 3.5 ± 0.9
+1.5+1.5
- 0.6
+1.7+1.1
- 0.9
+2.5+0.9
- 1.2
+1.3+0.9
- 1.0
Notes. The shift velocity v and equivalent width Wλ are given by Equations (8)
and (9). Note that the (r2) component is likely to be caused by absorption from
blueshifted transitions of Fe XXI, Fe XX, and Ne X.
detected in front of the X-ray source. For the same reason,
we cannot use the Fe L-shell density diagnostics (Mauche et al.
2005) as, e.g.,
the Fe XXII emission lines at 11.77 Å and
11.92 Å used by Mauche et al. (2003) are both seen in absorp-
tion. But we can still use the detected He-i and -f emission
lines to estimate the density via the R-ratio, noting the caveat
that the densities are systematically overestimated in the pres-
ence of an external UV radiation field, as photoexcitation of
the 1s2s(cid:0)3S1(cid:1) → 1s2p(cid:0)3P(cid:1) transition depopulates the upper
level of the f-line in favor of the i-line and leads to a lower
R-ratio (see, e.g., Mewe & Schrijver 1978; Kahn et al. 2001;
Wojdowski et al. 2008).
The i- and f-lines of the Mg XI triplet are seen as two distinct
components each -- one with almost no shift, which is consistent
with the r-absorption line, and the other one at a redshift of 400 --
1000 km s- 1 (see Figure 12 and Table 8). Given these two emis-
sion components and their Doppler shifts, one could be tempted
to identify the absorption feature at 9.19 Å with a second red-
shifted Mg XI r line, but our model for the complete absorption
line series (Section 3.5) predicts that the blueshifted 2s → 4p
transitions from the ground states of Fe XXI and Fe XX, as well
as the Ne X 1s → 10p transition, account for most of the absorp-
tion seen at 9.19 Å. Table 9 shows the R-ratios obtained for the
two pairs of lines as well as the corresponding densities accord-
z
12
HANKE ET AL.
Electron Densities Corresponding to R = F(f)/F(i) Ratios
Table 9
Solutions to Equation (15) for the Model of Equation (13)
Table 10
Component
First pair
of i and f lines
(unshifted)
Second pair
of i and f lines
(redshifted)
Measurement
R1(Mg XI) = 1.6+0.7
- 0.6n
R2(Mg XI) = 0.8 ± 0.5n
2.3
1.6
1.0
1.3
0.8
0.3
Note. a A temperature between 0.3 and 8 MK is assumed
and UV-photoexcitation is not considered.
Model Calculations by
Porquet & Dubau (2000)a
R(Mg XI)
ne
(1012 cm- 3)
1.0. . . 4.5
3.5. . . 10
8 . . . 22
5. . . 15
12. . . 28
43. . . 91
ing to the calculations of Porquet & Dubau (2000, Fig. 8e), ne-
glecting the influence of the UV radiation of HDE 226868 on the
R-ratio. The unshifted lines would then be caused by a plasma
with an electron density ne,1 ≈ 5×1012 cm- 3; the redshifted pair
of lines seems to stem from another plasma component with
ne,2 ≫ ne,1. A more detailed discussion of these densities is
presented in Section 4.1.
The S/N of the spectrum above 20 Å is not high enough to
describe the O VII triplet. The Ne IX triplet blends with sev-
eral Fe XIX lines (Figure 11). Na X i- and f-lines are likely to
be present with several components, but those cannot be distin-
guished clearly. Similar as for Mg, there are also two i-lines
of Al XII at shifts of - 477+120
- 228 km s- 1, re-
spectively, but no f-line is detected due to a blend with the
strong Mg XI He β absorption line and especially an Fe XXII
1s22s2p(2s→5p) absorption line at 7.87 Å. Similarly, the Si XIII
i-line is not detected as it overlaps with absorption features
probably due to nearly neutral Si K-edge structures. (Further-
more, the flux of the Si XIII f-line might be underestimated as it
blends with the Mg XII Ly γ line.) The i- and f-lines of He-like
sulfur are detected with R(S XV) = 0.61+1.31
- 0.60, but S XV has not
been modeled by Porquet & Dubau (2000). Below 4 Å, the S/N
below is again not high enough to resolve further triplets from
heavier elements such as argon, calcium, or iron.
- 240 km s- 1 and +333+156
4. DISCUSSION AND CONCLUSIONS
In the following, we discuss our results and derive constraints
on the stellar wind in the accretion region.
4.1. Velocity and Density of the Stellar Wind
While a spherically symmetric model for the stellar wind
in the HDE 226868/Cyg X-1 system can be excluded by ob-
servations (see, e.g., Gies & Bolton 1986b; Gies et al. 2003;
Miller et al. 2005; Gies et al. 2008), a symmetric velocity law
v(r) = v∞ · f (r/R⋆)
(12)
is usually assumed to obtain a first estimate of the particle den-
sity in the wind. The fraction f of the terminal velocity v∞ is
often parameterized by (Lamers & Leitherer 1993, Equation 3)
f (x) = f0 + (1 -
f0)·(cid:0)1 - 1/x(cid:1)β
with f0 := v0/v∞, where v0 is the velocity at the base of the
wind (x = 1). The simple model for the radiatively driven
wind of isolated stars by Castor et al. (1975) is obtained for
(for x ≥ 1)
(13)
nH(x)
(1010 cm- 3)
440
110
110
110
22
22
22
4.4
4.4
4.4
d(x)
200
f0
= v0/v∞
0.005
50
50
50
10
10
10
2
2
2
0.011
0.011
0.011
0.011
0.011
0.011
0.011
0.011
0.011
β
x
= r/R⋆
2100 · f (x)
for β = βmax
< ∞
1
≤ 1 ≤ 1.01
≤ 2 ≤ 1.08
≤ 3 ≤ 1.18
≤ 1 ≤ 1.08
≤ 2 ≤ 1.29
≤ 3 ≤ 1.48
≤ 1 ≤ 1.36
≤ 2 ≤ 1.69
≤ 3 ≤ 1.97
11
≤ 41
≤ 36
≤ 30
≤ 180
≤ 127
≤ 95
≤ 570
≤ 368
≤ 271
Note. The binary separation a = 41 R⊙ corresponds to xa = a/R⋆ = 2.4.
β = 1/2. The photoionization of the wind, however, sup-
presses its acceleration (Blondin 1994), such that a smaller
f (e.g., a larger β within the same model) is required in the
Strömgren zone. Gies & Bolton (1986b) have explained the
orbital variation of the 4686 Å He II emission line profile of
Cyg X-1 with a similar model for the focused wind, where
v∞, R⋆, and β depend on the angle θ from the binary axis.
β was interpolated between 1.60 and 1.05 for θ between 0 and
20◦. The value β(θ = 20◦) = 1.05 is, however, often used for
a spherically symmetrical wind as well (Lachowicz et al. 2006;
Szostek & Zdziarski 2007; Poutanen et al. 2008). Vrtilek et al.
(2008) use f0 = 0.01 to avoid numerical singularities and fit
the (relatively low) value β ≈ 0.75 to their models for UV
v0 is likely to be of the order of the thermal veloc-
lines.
ity of H atoms, which is (2kT /mH)1/2 = 23 km s- 1 and corre-
sponds to f0 = 0.011 for HDE 226868's effective temperature
Teff = 32 000 K and v∞ = 2100 km s- 1 (Herrero et al. 1995).
With the mean molecular weight per H atom, µ ≈ 1.4,14
the continuity equation M⋆ = µmH nH(r) · 4πr2 v(r) gives the
following estimate for the hydrogen density profile:
nH(r) =
M⋆/(µmH)
4πR2
⋆v∞ · d(r/R⋆) with d(x) = x- 2/ f (x)
(14)
Using the parameters of HDE 226868 ( M⋆ = 3×10- 6 M⊙ yr- 1,
R⋆ = 17 R⊙; Herrero et al. 1995;
Nowak et al. 1999, Table 1), Equation (14) predicts
also summarized by
nH(r) = 2.2×1010 cm- 3 · d(r/R⋆) .
(15)
0
and
1 ≤ x ≤ p1/( f0d)
Equation (14) can only be solved within the model of Equa-
tion (13) -- or any other model for f
in Equation (12) with
f (x) ≥ f0 -- for
d < f - 1
(16)
For f0 ≈ 0.01, the value d ≈ 190 -- which would be required to
explain the density15 nH,1 ≈ 4.2×1012 cm- 3 obtained from the
unshifted Mg XI triplet (Section 3.6) -- can never be reached
within our model for the continuous spherical wind. This result
shows that the density is likely to be overestimated by an R-ratio
analysis which ignores the strong UV-flux of the O9.7 star. Ta-
ble 10 lists therefore some solutions to Equation (15) for much
lower densities as well. Due to the additional constraint that
the radial velocity is less than 100 km s- 1 (Table 8), we suggest
.
14 The helium abundance per H atom is ≈10% (Wilms et al. 2000).
15 A plasma of fully ionized H and He contains ≈1.2 electrons per H atom.
CHANDRA X-RAY SPECTROSCOPY OF THE FOCUSED WIND IN THE CYG X-1 SYSTEM.
I.
13
that the first emission component stems from close to the stel-
lar surface. Although the simple wind model of Equation (13)
may not be appropriate in this region, the results of Table 10 are
rather insensitive to the assumptions on the velocity law, as a
wide range of β values was considered, but mostly depend on
the wind's initial velocity v0, for which we have used a reason-
able estimate. Note that f0 and d also depend on v∞, but their
product, which is important in Equation (16), does not.
In spite of the systematic errors of the (absolute) density anal-
ysis with the R-ratio, we infer that the second plasma com-
ponent is much denser relative to the first one. As it is seen
at a larger redshift of 400 -- 1000 km s- 1, we favor its identifi-
cation with the focused wind. The two emission components
could also be caused coincidentally by dense clumps in the stel-
lar wind (which are common for O-stars; e.g., Oskinova et al.
2007; Lépine & Moffat 2008), but the interpretation as a slow
base of the wind close to the stellar surface and a focused wind
between the accreting black hole and its donor star provides a
consistent description of both emission components: an undis-
turbed wind (with v∞ and f0 as above, and β = 1.05 → 1.6)
would reach a velocity of v(a) ≈ 900 ← 1200 km s- 1 in the dis-
tance of the black hole. The focused wind in the orbital plane
would then be detected with a projected velocity v(a) sini =
500 ← 700 km s- 1 at φorb = 0. While photoionization reduces
the efficiency of acceleration for the spherical wind, the denser
focused wind is less strongly affected due to self-shielding and
reaches the expected velocity. It is an observational fact that
the focused wind has another ionization structure: optical emis-
sion lines (Hα, He II λ4686) from ions which only exist at a
low ionization parameter have been observed in the focused
stream (see, e.g., Gies & Bolton 1986b; Ninkov et al. 1987b;
Gies et al. 2003). For the spherical wind, however, Gies et al.
(2008) have conjectured that the part between Cyg X-1 and the
donor star might be completely ionized in the soft state. Our
observations show that the situation in the hard state may be
similar.
4.2. Modeling of the Photoionization Zone
We use the photoionization code XSTAR 2.1ln7b16
(Kallman & Bautista 2001) to model the photoionization zone.
As the latter is quite complex (due to the inhomogeneous wind
density, which is strongly entangled with the X-ray flux), we do
not claim to describe the photoionized wind self-consistently,
but only want to derive a first approximation. For an optically
thin plasma, the relative population of a given atom's ions is
merely a function of the ionization parameters
ξ(r) =
L
nH r2 =
L37
n13 r2
12
erg cm s- 1 ,
(17)
where L37 is the ionizing source luminosity above 13.6 eV
in 1037 erg s- 1, n13 is the hydrogen density in 1013 cm- 3 and
r12 is the distance from the source in 1012 cm.
For Cyg X-1 as observed in the hard state, extrapolation of
the unabsorbed model obtained by our analysis (Table 2) gives
L37 ≈ 3.5. Although there are obviously strong variations in
the density, the XSTAR calculation has been performed with
constant n13 = 0.01, which is the average of nH(r) from Equa-
tion (15) for R⋆ ≤ r ≤ a. Figure 13 shows the resulting popu-
lation of Fe ions. The ionization parameters at which the pop-
ulation distributions of some ions of interest peak, as well as
the corresponding full width at half maximum (FWHM), are
presented in Table 11. We have also included Si IV, N V, and
16 See http://heasarc.gsfc.nasa.gov/lheasoft/xstar/xstar.html.
K−shell ions
L−shell ions
I
V
X
X
e
F
V
X
X
e
F
I
V
X
X
e
F
X
X
e
F
I
I
I
X
X
e
F
I
X
X
e
F
I
I
X
X
e
F
I
X
X
e
F
6
.
0
n
o
i
t
a
l
u
p
o
p
e
v
i
t
a
l
e
R
4
.
0
2
.
0
I
I
I
I
I
V
X
e
F
V
X
e
F
0
3
2.5
log x
2
Figure 13. Relative population of the ionization stages of iron as a function of
the ionization parameter ξ = L/(nH r2), as calculated with XSTAR for L37 = 3.5
and n13 = 0.01.
Ionization Parameters for Peak Ion Populations
Table 11
Ion
Fe XXVI
Fe XXV
Fe XXIV
Fe XXIII
Fe XXII
Fe XXI
Fe XX
Fe XIX
Fe XVIII
Fe XVII
Ca XX
Ca XIX
Ar XVIII
Ar XVII
S XVI
S XV
Si XIV
Si XIII
Si IV
Mg XII
Mg XI
Ne X
Ne IX
O VIII
O VII
N V
C IV
log ξ
2.91
2.75
2.65
2.58
2.51
2.44
2.32
2.12
2.01
1.97
2.69
2.39
2.55
2.22
2.38
2.08
2.21
1.96
1.09
2.06
1.82
1.94
1.74
1.80
1.62
1.51
1.48
FWHM(log ξ)
r12 = `L37/n13/10log ξ´1/2
0.27
0.32
0.30
0.28
0.31
0.36
0.44
0.40
0.23
0.14
0.56
0.54
0.65
0.50
0.69
0.43
0.65
0.39
0.18
0.58
0.31
0.40
0.24
0.26
0.22
0.01
0.19
0.6 -- 0.8
0.7 -- 1.0
0.8 -- 1.1
0.8 -- 1.1
0.9 -- 1.3
0.9 -- 1.4
1.0 -- 1.7
1.2 -- 1.9
1.6 -- 2.0
1.8 -- 2.1
0.7 -- 1.2
0.9 -- 1.6
0.7 -- 1.5
1.0 -- 1.8
0.8 -- 1.7
1.2 -- 2.0
0.9 -- 2.0
1.5 -- 2.4
4.7 -- 5.7
1.2 -- 2.3
1.8 -- 2.6
1.6 -- 2.5
2.1 -- 2.8
1.9 -- 2.6
2.5 -- 3.2
3.25 -- 3.30
3.3 -- 4.1
Notes. We list the range in ionization parameter and distance for the maximum
ion population (Figure 13) obtained in an XSTAR simulation with L37 = 3.5
and n13 = 0.01. Note that the binary separation in units of 1012 cm is a12 = 2.9.
C IV for a comparison with the work of Vrtilek et al. (2008)
and Gies et al. (2008). The H- and He-like ions detected with
Chandra only appear at considerable distance from the X-ray
source. Taking into account that the photons emanating from
Cyg X-1 first propagate through a lower density wind and that
the (eventually stalled) wind of higher density is only reached
in the vicinity of the star, the actual distances will be even larger
than the r12 values quoted in Table 11.
14
0.7
0.8
0.5
0.6
0.9
1.0
Figure 14. Gray-scale image of vrad,BH(rrr) of Equation (18), i.e., projected wind
velocity against the black hole. A spherically symmetric velocity law given by
Equations (12) and (13) with v∞ = 2100 km s- 1,
f0 = 0.01, and β = 1.05 is
assumed. The star and the black hole are shown as filled black circles; the size
of the latter is the accretion radius 2GMBH/v(a)2. On the black circle passing
through the center of the star and through the black hole vrad,BH(rrr) is 0 since
α(rrr) = 90◦. Positive velocities (+) are seen only within this circle; a lighter gray
means a larger redshift. Negative velocities (- ) can likewise only occur outside
of this circle; a lighter gray here means a larger blueshift. The dashed gray lines
of constant vrad,BH are shown from - 1400 km s- 1 to +1000 km s- 1 in steps of
200 km s- 1. The two highlighted sectors contain all observable lines of sight
toward the black hole (after rotation in this plane) for the inclination i = 35◦.
The labels show the corresponding orbital phases.
4.3. Origin of Redshifted X-Ray Absorption Lines
Additional constraints on the accretion flow can be derived
by considering the Doppler shifts observed in absorption lines
(Sections 3.4 and 3.5). Models for the wind velocity like that of
Equations (12) and (13) predict a velocity vvv(rrr) at the position rrr
in the stellar wind. But only the projection of vvv against the black
hole, vrad,BH, can eventually be observed as radial velocity in an
X-ray absorption line. With the angle α(rrr) between vvv(rrr) and the
direction from rrr toward the black hole, vrad,BH(rrr) is
vrad,BH(rrr) = cos α(rrr) · vvv(rrr) .
(18)
Assuming a velocity field with radial symmetry with respect
to the star, α(rrr) is just the angle at rrr between the star and the
black hole. For example, α is 90◦ (and vrad,BH is therefore 0) on
the sphere containing the center of the star and the black hole
diametrically opposed. Redshifted absorption lines can only be
observed from wind material inside this sphere, while the part
of the wind outside of it is always seen at a blueshift -- a fact
which is independent of the assumed velocity field as long as it
is directed radially away from the star.
For a spherically symmetrical wind model, any line of sight
can be rotated to an equivalent one in a half-plane limited by
the binary axis. Only a sector with half-opening angle i can be
observed unless the system's inclination is i = 90◦. Figure 14
shows the projected velocity vrad,BH(rrr) for a simple wind model
and two sectors containing all observable lines of sight toward
Cyg X-1 for an inclination of i = 35◦.
We now investigate the region where the projected wind ve-
locity (Figure 14) is compatible with the observed Doppler
HANKE ET AL.
−
+
0.5
0.4
0.3
0.2
0.1
orb = 0.0
]
s
p
c
[
e
t
a
r
t
n
u
o
c
M
S
A
E
T
X
R
V
e
k
2
1
−
5
1
.
0
7
0
6
0
5
0
4
0
3
0
2
# 3724
# 3407
# 107
# 2415
# 1511
0.7
0.8
orbital phase f
0.9
orb
# 3814
1
Figure 15. Orbital phase and mean ASM count rate of the previously reported
Chandra observations of Cyg X-1 (ObsID 3814 is presented in this paper).
shifts (Tables 5 and 7). We are confident that this method
allows for the identification of the absorption regions, as the
low observed velocities are always found close to the α = 90◦
sphere -- independent of the wind model. For most of the in-
vestigated ions (Ne X, Mg XII, Si XIV and XIII, S XV, Ar XVII,
Fe XIX and XVIII) the inferred distance from the black hole
agrees with the predictions of Table 11, e.g., the projected ve-
locity - 117 km s- 1 ≤ vNe X ≤ - 69 km s- 1 measured for Ne X is
(during 0.93 ≤ φorb ≤ 1) obtained at r12 = 1.78 ± 0.07. For
many other ions, both results are still very similar: e.g., Ne IX
with - 214 km s- 1 ≤ vNe IX ≤ - 123 km s- 1 is expected at r12 =
1.95± 0.07 from the vrad,BH model and at r12 = 2.1 -- 2.8 from
the photoionization model. Only for the highly ionized iron
lines, the small distances are -- as already anticipated in Sec-
tion 4.2 -- underestimated by the XSTAR simulation run with
constant average density, which overestimates the wind den-
sity close to the black hole. For example, the velocity range
- 9 km s- 1 ≤ vFe XXIV ≤ 52 km s- 1 measured for Fe XXIV corre-
sponds to r12 = 1.51 ± 0.07, while the population of this ion
peaks at r12 = 0.8 -- 1.1 within the model presented in Table 11.
4.4. Previous ChandraObservations
In previous Chandra-HETGS observations of Cyg X-1, the
stellar wind was seen under different viewing angles, or the X-
ray source was in other spectral states (see Figure 15), which
probably means that the properties of the wind were also dif-
ferent. Schulz et al. (2002) have analyzed the 14 ks observation
ObsID 107 performed in 1999 October (at φorb ≈ 0.74),17 when
Cyg X-1 was in a transitional state. They derive a neutral col-
umn density of NH = 6.2×1021 cm- 2 from prominent absorp-
tion edges (see also Table 3) and detect some emission and ab-
sorption lines with indications of P Cygni profiles. Miller et al.
(2005) have investigated the focused wind with the 32 ks ob-
servation ObsID 2415 of Cyg X-1 in an intermediate state, per-
formed in 2001 January (at φorb ≈ 0.77). They report absorp-
tion and emission lines of H- and He-like resonance lines of
Ne, Na, Mg, and Si with a mean redshift of ≈100 km s- 1, as
density is 6.2×1021 cm- 2 (Miller et al. 2002). Marshall et al.
(2001) describe Ly α and He α absorption lines, redshifted by
(450 ± 150) km s- 1 from the 12.6 ks observation ObsID 1511
well as some lines of highly ionized Fe and Ni. The column
of Cyg X-1 in the hard state, performed in 2001 January (at
17 Note that Schulz et al. (2002) and Miller et al. (2005) quote an erroneous
date and thus the wrong orbital phase φorb = 0.93 6= 0.74 for this observation.
f
CHANDRA X-RAY SPECTROSCOPY OF THE FOCUSED WIND IN THE CYG X-1 SYSTEM.
I.
15
φorb ≈ 0.83). Chang & Cui (2007) report dramatic variability in
the 30 ks observation ObsID 3407 performed in 2001 October
(at φorb ≈ 0.88), when Cyg X-1 reached its soft state. While a
large number of absorption lines (mostly redshifted, but not by a
consistent velocity) is identified in the first part of their observa-
tion, most of them weaken significantly or cannot be detected
at all during the second part. The complete ionization of the
wind due to a sudden density decrement is given as a possible
explanation. Feng et al. (2003) detect asymmetric absorption
lines with the 26 ks observation ObsID 3724 of Cyg X-1 in the
soft state, performed in 2002 July (at φorb = 0). The line cen-
ters are almost at their rest wavelengths, but the red wings are
more extended, especially for the transitions of highest ionized
ions, which they explain by the inflowing focused wind reach-
ing both the highest redshift and ionization parameter closest to
the black hole.
The interpretation of the absorption lines presented in the pre-
vious section describes our observation consistent with wind
and photoionization models.
It can, however, not be applied
to all of the other observations; the model of Figure 14 nei-
ther predicts a conspicuously high redshift at φorb ≈ 0.83 and
at a higher soft X-ray flux (ObsID 1511), nor a positive red-
shift at φorb ≈ 0.77 at a still higher flux (ObsID 2415, see Fig-
ure 15). Inhomogeneities (e.g., density enhancements in shield-
ing clumps) and asymmetries of the wind (e.g., due to noniner-
tial forces in the binary system) may therefore play an important
role in these cases.
4.5. Summary
In this paper, we have presented a Chandra-HETGS observa-
tion of Cyg X-1 during superior conjunction of the black hole,
which allows us to detect the X-ray absorption signatures of the
stellar wind of HDE 226868. The light curve near φorb = 0 is
shaped by absorption dips; these are, however, excluded from
this analysis by selecting nondip times of high count rate only.
At a flux of ∼0.25 Crab, we have to deal with moderate pile-
up in the grating spectra, for which can be accounted very
well with the simple_gpile2 model. The continuum of
both Chandra's soft and RXTE's broadband X-ray spectrum
has been described consistently by a single model, consisting
of an empirical broken power-law spectrum with high-energy
cutoff (which is typical for the hard state) and a subordinated
disk component with an inner radius close to the ISCO. The
joint modeling of both spectra reveals the presence of both a
narrow and a relativistically broadened fluorescence Fe Kα line.
Chandra has resolved absorption edges of neutral atoms with an
overabundance of metals, suggesting an origin not only in the
ISM, but also in the stellar wind of the evolved supergiant. The
previously suspected anticorrelation of neutral column densi-
ties and X-ray flux, which is due to photoionization, is con-
firmed. Absorption lines of highly ionized ions are produced
where the wind becomes extremely photoionized. For H- and
He-like ions, Lyman and He-series are detected up to the Ly
or He ǫ line, which we use to measure the column density and
velocity of absorbing ions. For the wealth of Fe L-shell transi-
tions, column densities can best be obtained by directly using
a physical model for the complete line series of an ion. The
nondip spectrum shows almost no Doppler shifts, probably in-
dicating that we have excluded the focused wind by our selec-
tion of nondip times. We have also detected two plasma com-
ponents in emission by two pairs of i- and f-lines of He-like
Mg XI. The first one is roughly at rest and we identify it with
the base of the spherical wind close to the stellar surface. The
second plasma component is denser and observed at a redshift
that is compatible with the focused wind. A simple XSTAR
simulation indicates that most of the observed ions only exist
in a distance of the black hole, where the velocity of a spher-
ical wind, projected against the black hole, is low -- which is
a consistent explanation of the small Doppler shifts observed
in absorption lines. We review the previously reported Chan-
dra observations of Cyg X-1 and find that not all of them can
be described in the same picture, as the wind may be affected
by asymmetries or inhomogeneities. A detailed spectroscopy of
the absorption dips, which might shed more light on the focused
wind, will be presented in a subsequent paper.
We are grateful to A. Juett for her help with the data reduction
and her contributions to tbnew model. We thank A. Young,
J. Xiang, and M. Böck for helpful discussions, J. Houck for
his support on ISIS and T. Kallman for his help with XSTAR.
M. H. and J. W. acknowledge funding from the Bundesminis-
terium für Wirtschaft und Technologie through the Deutsches
Zentrum für Luft- und Raumfahrt under contract 50OR0701.
M. N. is supported by NASA Grants GO3-4050B and SV3-
73016. We thank the International Space Science Institute,
Berne, Switzerland, and the MIT Kavli Institute for Space Re-
search, Cambridge, MA, USA, for their hospitality during the
preparation of this work.
REFERENCES
Bałuci´nska-Church, M., Belloni, T., Church, M.J., & Hasinger, G. 1995,
A&A, 302, L5
Bałuci´nska-Church, et al. 2000, MNRAS, 311, 861
Belloni, T., et al. 1996, ApJ, 472, L107
Blondin, J.M. 1994, ApJ, 435, 756
Bolton, C.T. 1972, Nature, 240, 124
Bondi, H., & Hoyle, F. 1944 MNRAS, 104, 273
Bowyer, S., Byram, E.T., Chubb, T.A., & Friedman, H. 1965,
Science, 147, 394
Brocksopp, et al. 1999, MNRAS, 309, 1063
Cadolle Bel, M., et al. 2006, A&A, 446, 591
Canizares, C.R., et al. 2005, PASP, 117, 1144
Castor, J.I., Abbott, D.C., & Klein, R.I. 1975, ApJ, 195, 157
Chang, C., & Cui, W. 2007 ApJ, 663, 1207
Conti, P.S. 1978 A&A, 63, 225
Chandra X-ray Center 2005, The Chandra ABC Guide to Pileup,
http://cxc.harvard.edu/ciao/download/doc/pileup_abc.ps
Chandra X-ray Center 2006, The Chandra Proposers' Observatory Guide,
http://cxc.harvard.edu/proposer/POG/
Davis, J.E. 2002, in High Resolution X-ray Spectroscopy with XMM-Newton
and Chandra, ed. G. Branduardi-Raymont,
http://adsabs.harvard.edu/abs/2002hrxs.confE..11D
Davis, J.E. 2003, in Proc. SPIE 4851, X-Ray and Gamma-Ray Telescopes and
Instruments for Astronomy, ed. J.E. Trümper, & H.D. Tananbaum
(Bellingham, WA:SPIE), 101
Dotani, T., et al. 1997 ApJ, 485, L87
Doty, J.P. 1994, The All Sky Monitor for the X-ray Timing Explorer, Technical
report, MIT, http://xte.mit.edu/XTE.html
Feng, Y.X., Tennant, A.F., & Zhang, S.N. 2003 ApJ, 597, 1017
Friend, D.B., & Castor, J.I. 1982 ApJ, 261, 293
Gabriel, A.H., & Jordan, C. 1969, MNRAS, 145, 241
Garmire, G.P., et al. 2003, in Proc. SPIE 4851, X-Ray and Gamma-Ray
Telescopes and Instruments for Astronomy, ed. J.E. Truemper, & H.D.
Tananbaum (Bellingham, WA:SPIE), 28
Gies, D.R., & Bolton, C.T. 1982 ApJ, 260, 240
Gies, D.R., & Bolton, C.T. 1986a, ApJ, 304, 371
Gies, D.R., & Bolton, C.T. 1986b, ApJ, 304, 389
Gies, D.R., et al. 2003, ApJ, 583, 424
Gies, D.R., et al. 2008, ApJ, 678, 1237
Gleissner, T., et al. 2004a, A&A, 425, 1061
Gleissner, T., et al. 2004b, A&A 414, 1091
Gorczyca, T.W. 2000, Phys. Rev. A, 61, 024702
Gorczyca, T.W., & McLaughlin, B.M. 2000, J. Phys. B Atom. Mol. Phys.,
33, L859
Gruber, D.E., et al. 1996, A&AS, 120, C641
Hatchett, S., & McCray, R. 1977, ApJ, 211, 552
Herrero, A., et al. 1995, A&A, 297, 556
Houck, J.C. 2002, in High Resolution X-ray Spectroscopy with XMM-Newton
and Chandra, ed. G. Branduardi-Raymont,
http://adsabs.harvard.edu/abs/2002hrxs.confE..17H.
Humphreys, R.M. 1978, ApJS, 38, 309
Ishibashi, K. 2006, Specification for finding zeroth order positions in grating
data analysis, http://space.mit.edu/CXC/docs/memo_fzero_1.3.ps
16
HANKE ET AL.
Jahoda, K., et al. 1996, in Proc. SPIE 2808, EUV, X-Ray, and Gamma-Ray
Instrumentation for Astronomy, Vol. VII ed. O.H. Siegmund, & M.A.
Gummin (Bellingham, WA:SPIE), 59
Juett, A.M., Schulz, N.S., & Chakrabarty, D. 2004, ApJ, 612, 308
Juett, A.M., Schulz, N.S., Chakrabarty, D., & Gorczyca, T.W. 2006a,
ApJ, 648, 1066
Juett, A.M., Wilms, J., Schulz, N.S., & Nowak, M.A. 2006b, BAAS, 38, 921
Kahn, S.M., et al. 2001, A&A, 365, L312
Kalberla, P.M.W., et al. 2005, A&A, 440, 775
Kallman, T., & Bautista, M. 2001, ApJS, 133, 221
Karitskaya, E.A., et al. 2006, Inf. Bull. Variable Stars, 5678, 1
Kirsch, M.G., et al. 2005, in Proc. SPIE 5898 UV, X-Ray, and Gamma-Ray
Space Instrumentation for Astronomy, Vol. XIV, ed. O.H.W. Siegmund
(Bellingham, WA:SPIE), 22
Kortright, J.B., & Kim, S.-K. 2000, Phys. Rev. B, 62, 12216
Kotani, T., et al. 2000, ApJ, 539, 413
Lachowicz, P., et al. 2006, MNRAS, 368, 1025
Lamers, H.J.G.L.M., & Leitherer, C. 1993, ApJ, 412, 771
Lee, J.C., et al. 2001, ApJ, 554, L13
Lee, J.C., et al. 2002, ApJ, 567, 1102
Lépine, S., & Moffat, A.F.J. 2008, AJ, 136, 548
Levine, A.M., et al. 1996, ApJ, 469, L33
Makishima, K., et al. 2008, PASJ, 60, 585
Marshall, H.L., Schulz, N.S., Fang, T., Cui, W., Canizares, C.R., Miller, J.M.,
& Lewin, W.H.G. 2001, in X-ray Emission from Accretion onto
Black Holes, ed. T. Yaqoob, & J.H. Krolik (Baltimore: John Hopkins Univ.),
45
Mauche, C.W., Liedahl, D.A., & Fournier, K.B. 2003, ApJ, 588, L101
Mauche, C.W., Liedahl, D.A., & Fournier, K.B. 2005, in AIP Conf. Ser. 774,
X-ray Diagnostics of Astrophysical Plasmas: Theory, Experiment, and
Observation, ed. R. Smith (New York: AIP), 133
McConnell, M.L., et al. 2002, ApJ, 572, 984
Merloni, A., Fabian, A.C., & Ross, R.R. 2000, MNRAS, 313, 193
Mewe, R., & Schrijver, J. 1978, A&A, 65, 99
Mihalas, D. 1978, Stellar atmospheres (San Francisco, CA: W.H. Freeman)
Miller, J.M., et al. 2002, ApJ, 578, 348
Miller, J.M., et al. 2005, ApJ, 620, 398
Murdin, P., & Webster, B.L. 1971, Nature, 233, 110
Ninkov, Z., Walker, G.A.H., & Yang, S. 1987a, ApJ, 321, 425
Ninkov, Z., Walker, G.A.H., & Yang, S. 1987b ApJ, 321, 438
Nowak, M.A., et al. 1999, ApJ, 510, 874
Nowak, M.A., et al. 2008, ApJ, in press (arXiv: 0809.3005)
Oskinova, L.M., Hamann, W.-R., & Feldmeier, A. 2007, A&A, 476, 1331
Porquet, D., & Dubau, J. 2000, A&AS, 143, 495
Pottschmidt, K., el al. 2000, A&A, 357, L17
Pottschmidt, K., et al. 2003, A&A, 407, 1039
Poutanen, J., Zdziarski, A.A., & Ibragimov, A. 2008, MNRAS, 389, 1427
Remillard, R.A., & McClintock, J.E. 2006, ARA&A, 44, 49
Russell, D.M., Fender, R.P., Gallo, E., & Kaiser, C.R. 2007,
MNRAS, 376, 1341
Schulz, N.S., et al. 2002, ApJ, 565, 1141
Shaposhnikov, N., & Titarchuk, L. 2007, ApJ, 663, 445
Shimura, T., & Takahara, F. 1995, ApJ, 445, 780
Spitzer, L. 1978, Physical processes in the interstellar medium (New York:
Wiley-Interscience)
Strömgren, B. 1939, ApJ, 89, 526
Szostek, A., & Zdziarski, A.A. 2007, MNRAS, 375, 793
Treves, A., et al. 1980, ApJ, 242, 1114
van Aken, P.A., & Liebscher, B. 2002, Phys. Chem. Miner., 29, 188
Verner, D.A., Verner, E.M., & Ferland, G.J. 1996,
At. Data Nucl. Data Tables, 64, 1
Verner, D.A., & Yakovlev, D.G. 1995, A&AS, 109, 125
Verner, D.A., Yakovlev, D.G., Band, I.M., & Trzhaskovskaya, M.B. 1993,
At. Data Nucl. Data Tables, 55, 233
Vrtilek, S.D., et al. 2008, ApJ, 678, 1248
Walborn, N.R. 1973, ApJ, 179, L123
Webster, B.L., & Murdin, P. 1972, Nature, 235, 37
Wilms, J., Allen, A., & McCray, R. 2000, ApJ, 542, 914
Wilms, J., et al. 2006, A&A, 447, 245
Wojdowski, P.S., Liedahl, D.A., & Kallman, T.R. 2008, ApJ, 673, 1023
Xiang, J., Zhang, S.N., & Yao, Y. 2005, ApJ, 628, 769
Zhang, S.N., et al. 1997, ApJ, 477, L95
Ziółkowski, J. 2005, MNRAS, 358, 851
|
astro-ph/9506102 | 1 | 9506 | 1995-06-20T13:17:35 | Gravitational Lensing and Anisotropies of CBR on the Small Angular Scales | [
"astro-ph"
] | We investigate the effect of gravitational lensing, produced by linear density perturbations, for anisotropies of the Cosmic Background Radiation (CBR) on scales of arcminutes. In calculations, a flat universe ($\Omega=1$) and the Harrison-Zel'dovich spectrum ($n=1$) are assumed. The numerical results show that on scales of a few arcminutes, gravitational lensing produces only negligible anisotropies in the temperature of the CBR. Our conclusion disagrees with that of Cay\'{o}n {\it et al.} who argue that the amplification of $\Delta T/T$ on scales $\le 3'$ may even be larger than 100\%. | astro-ph | astro-ph | GRAVITATIONAL LENSING AND ANISOTROPIES OF CBR
ON THE SMALL ANGULAR SCALES
J.M. Liu
J.G. Gao
a
b
International School for Advanced Studies(SISSA),
Via Beirut n.-, Trieste Italy
a
b
I.C.R.A., Dipartimento di Fisica, Universit(cid:19)a \La Sapienza",
Piazzale Aldo Moro, - Roma Italy
Accepted for Publication in: MNRAS
We investigate the e(cid:11)ect of gravitational lensing, produced by linear density per-
turbations, for anisotropies of the Cosmic Background Radiation (CBR) on scales
of arcminutes. In calculations, a (cid:13)at universe ((cid:10) = ) and the Harrison-Zel'dovich
spectrum (n = ) are assumed. The numerical results show that on scales of a few
arcminutes, gravitational lensing produces only negligible anisotropies in the tem-
perature of the CBR. Our conclusion disagrees with that of Cay(cid:19)on et al. who argue
that the ampli(cid:12)cation of (cid:1)T =T on scales (cid:20)
may even be larger than %.
Key words:cosmology - gravitational lensing - anisotropy on CBR.
I. INTRODUCTION
Many authors have computed the e(cid:11)ect of gravitational lensing for anisotropies of the
CBR (Blanchard & Schneider ; Kashlinsky ; Cole & Efstathiou ; Tomita &
Watanabe ; Linder ; Watanabe & Tomita ; Feng & Liu. ; Cay(cid:19)on et al.
). Unfortunately, however, their conclusions are controversial. Roughly speaking, there
exist three di(cid:11)erent kinds of conclusion so far. The (cid:12)rst is that gravitational lensing e(cid:11)ects
strongly erase (cid:13)uctuations of the CBR on scales of a few arcminutes (Kashlinsky ); the
second is that an appreciable, even strong, ampli(cid:12)cation of (cid:1)T =T is possible (Sasaki ;
Linder ; Cay(cid:19)on et al. ); the last is that gravitational lens e(cid:11)ects on the CBR are
negligible (Cole & Efstathiou ; Tomita & Watanabe ).
Recently, Cay(cid:19)on et al. presented calculations of the gravitational lensing e(cid:11)ects, pro-
duced by linear density (cid:13)uctuations on the CBR and got an interesting result. Their work
implied that there should be an appreciable ampli(cid:12)cation, of the order of %, for (cid:1)T =T in
present experiments on the scales of several arcminutes, whereas previous work including the
e(cid:11)ect of nonlinear density (cid:13)uctuations found a negligible ampli(cid:12)cation (Cole & Efstathiou
). This implies that the gravitational lensing e(cid:11)ect on anisotropies of the CBR due
to linear density (cid:13)uctuations overwhelms that due to nonlinear clustering. This result is
surprising and di(cid:14)cult to interpret.
In this paper, we use a new formalism to calculate the e(cid:11)ect of gravitational lensing
on anisotropies of the CBR produced by linear density (cid:13)uctuations. Our calculations show
that the gravitational lensing e(cid:11)ect on scales of a few arcminutes is essentially negligible in
contrast with the results of Cay(cid:19)on et al.
Within the geometrical optics approximation, a gravitational (cid:12)eld is equivalent to an
optical medium with a refractive index di(cid:11)erent from unity, and the de(cid:13)ection of light may
be interpreted in terms of the refractive index and its spatial variation. The amplitude and
phase (cid:13)uctuations, produced by a random gravitational (cid:12)eld, may then be calculated using
the usual methods of random medium optics (Fang ).
In Sec.II, we establish the basic equations for a wave scattered by a gravitational potential
and for the amplitude of the scattered wave. In Sec.III, we present the explicit formulae
for calculating the anisotropies of the CBR. Numerical results and brief conclusions are
summarized in Sec.IV.
II. THE BASIC METHOD
Under some reasonable assumptions (Cole & Efstathiou ), one can prove that there
exists a gauge such that the metric perturbations are characterized by a single potential
(cid:30)(t; x) (cid:28)
i
j
ds
= (cid:0)( + (cid:30))dt
+ ( (cid:0) (cid:30))a
(t)(cid:14)
dx
dx
;
()
ij
where a(t) is the scale factor of the universe. The relationship between the gravitational
potential (cid:30) and the matter density perturbation (cid:14)(cid:26) is with (cid:25)G = c =
(cid:1)(cid:30) =
a
(cid:26)
(cid:14)(cid:26):
()
b
In the background universe for the linear perturbation (which we assume to have (cid:10)
= ),
(cid:30) is time-independent.
For the sake of simplicity in writing the Maxwell equations in a perturbed metric, we
use the conformal time (cid:28) =
t
instead of t. Then
p
=
i
j
ds
= (cid:0)
( + (cid:30))(cid:28)
d(cid:28)
+
( (cid:0) (cid:30))(cid:28)
(cid:14)
dx
dx
:
()
ij
The speed of light in the metric of Eq.() is set to unity. A fundamental problem for treating
light propagating in an inhomogeneous universe is that of how to describe the gravitational
lensing e(cid:11)ect. According to geometrical optics, a gravitational (cid:12)eld is equivalent to a medium
with a refractive index di(cid:11)erent from unity and the de(cid:13)ection of light may be interpreted in
terms of the refractive index and its spatial variation. As a simple version of this analogue,
consider the spatial part of the photon four-momentum k
=
to be directed along the
d(cid:21)
(cid:22)
dx
(cid:22)
x
axis; one then has the following geodesic equation:
i
@ (cid:30)
dk
= (cid:0)
dx
;
i = ; :
()
i
@ x
On the other hand, the change in the direction of a light ray propagating in an inhomoge-
neous medium with refractive index n = + n
is
@ n
i
dk
=
dx
;
i = ; :
()
i
@ x
By comparing of Eqs.() and (), it is obvious that the e(cid:11)ect of the perturbed gravitational
(cid:12)eld is equivalent to a change in the refractive index, i.e.
n
= (cid:0)(cid:30):
()
Therefore, considering the equivalence of the two descriptions of light propagating in an
inhomogeneous matter distribution and in a medium with inhomogeneous refractive index,
the Maxwell equations may be written in the following form
@E
@ (cid:28)
= ( + (cid:30)) (cid:2)H;
()
@H
@ (cid:28)
= (cid:0)( + (cid:30)) (cid:2)E:
()
Formally, adopting Hanni's approach (Hanni ), the Maxwell equations in an universe
with an inhomogeneous matter distribution may be obtained in more rigorous way.
The propagation of an electromagnetic wave in a random inhomogeneous medium is
accompanied by a number of (cid:13)uctuation phenomena including polarization, (cid:13)uctuation of
phase and (cid:13)uctuation of amplitude. In the present paper, we concentrate on (cid:13)uctuation of
amplitude.
In a (cid:13)at universe, because the gravitational potential is time-independent, we can derive
a solution of the Maxwell equations () and () representing monochromatic waves with
(cid:12)xed frequency ! . In other words, we can assume that the electric and magnetic (cid:12)elds have
the form Re(Ee
) and Re(He
) respectively. In the case of (cid:21) = (cid:25)!
(cid:28) l
, (l
is the
(cid:0)i!(cid:28)
(cid:0)i!(cid:28)
(cid:0)
typical length-scale of inhomogeneity for matter in the universe), keeping only (cid:12)rst order
terms, the wave equation can be simpli(cid:12)ed as (Feng and Liu )
(
+ !
)E = (cid:0)!
(cid:30)E :
( )
To solve the equation of wave scattering, in the case of weak (cid:13)uctuations, we apply the
Born approximation to expand E in the series
E = E
+ E
+ E
+ (cid:1) (cid:1) (cid:1)
( )
with E
(cid:29) E
(cid:29) E
(cid:29) (cid:1) (cid:1) (cid:1). Putting Eq.( ) into Eq.( ), we get two (cid:12)rst-order equations:
(
+ !
)E
= ;
()
(
+ !
)E
= (cid:0)!
(cid:30)E
:
()
If the CBR is perfectly uniformly distributed, there is no net gravitational lensing e(cid:11)ect
on anisotropies of the CBR. This is a well-known result in geometrical optics and will be
con(cid:12)rmed below from the point of view of wave scattering. In order to study the gravitational
lensing e(cid:11)ect on anisotropies of the CBR, we focus on how the (cid:13)uctuation part of the CBR
is a(cid:11)ected. To do so, we separate E
into a homogeneous part E
and a (cid:13)uctuation part
h
E
f
with
E
= E
+ E
()
h
f
E
=< E
>;
E
= E
(cid:0) < E
> :
()
h
h
f
f
Here, the averaging is done over all observation directions. As E
and E
propagate
h
f
in an inhomogeneous universe, they interact with the gravitational potential and produce
scattered waves E
and E
respectively. E
and E
and their scattering terms satisfy
h
f
h
f
following equations
(
+ !
)E
= ;
()
h
(
+ !
)E
= (cid:0)!
(cid:30)E
;
()
h
h
(
+ !
)E
= ;
()
f
(
+ !
)E
= (cid:0)!
(cid:30)E
:
()
f
f
After scattering, the outgoing wave is
E = E
+ E
= E
+ E
+ E
+ E
:
( )
h
f
h
f
Because the last scattering surface is far away from us, without loss of generality, we let the
incident waves have the plane wave forms, E
= A
e
and E
= A
e
. The scattered
h
h
f
f
i!(cid:1)x
i!(cid:1)x
waves can then be easily expressed, using Green's function method, in the following forms
Z
i(!x(cid:0)!(cid:1)x)
!
e
E
(^x
) =
A
h
h
(
x)(cid:30)(x)
^
d
x;
( )
(cid:25)
x
Z
i(!x(cid:0)!(cid:1)x)
!
e
E
(^x
) =
A
(
x)(cid:30)(x)
d
x:
()
^
f
f
(cid:25)
x
When Eqs.( ) and () are evaluated, it is su(cid:14)cient to include only the contribution due
to waves scattered through angles not exceeding (cid:18) =
(cid:28) : In other words, the integration
(cid:21)
l
(cid:21)
can be con(cid:12)ned to the part of space which lies within the cone C (d(cid:10)), (cid:18) (cid:20)
, where (cid:18) is
l
the angle between the direction of observation and the direction of the scattering element,
and l
is the typical inhomogeneity scale in the universe. In fact, the integration functions
of Eqs.( ) and () oscillate rapidly outside the cone C (d(cid:10)), so that for a su(cid:14)ciently
smooth variation of (cid:30)(x), integration over the region external to cone C (d(cid:10)) only provides
a negligible contribution. This cone C (d(cid:10)) is much smaller than the angular scale of CBR
inhomogeneities and so we have
E
(^x
) = A
(^x
)
(cid:30)(x)
d
x;
()
h
h
(cid:25)
x
Z
i(!x(cid:0)!(cid:1)x)
!
e
E
(^x
) = A
(^x
)
(cid:30)(x)
d
x;
()
f
f
(cid:25)
x
Z
i(!x(cid:0)!(cid:1)x)
!
e
A
(^x
) is the amplitude of the uniform part of the CBR and A
(^x
) is that of the (cid:13)uc-
h
f
tuation part; as a direct consequence, E
(^x
) is independent of observation direction, but
h
E
(^x
) does depend on the observation direction.
f
III. ANGULAR CORRELATION FUNCTION
In the Rayleigh-Jeans part of blockbody radiation spectrum, T / I (where T and I are
temperature and intensity of the CBR respectively) and I / A
(where A is the amplitude
of the electric vector E). We will not distinguish A and E in the following except when
necessary. We then have
(cid:14)T
(cid:14)E
E (cid:0) hE i
=
=
()
T
E
E
Putting Eq.( ) into Eq.() and using the de(cid:12)nitions of E
and E
, we have
h
f
(cid:14)E
E
+ E
(cid:0) hE
i + E
(cid:0) hE
i
f
h
h
f
f
=
:
()
E
E
h
It is obvious from Eqs.() and () that E
(cid:0) hE
i = and hE
i = . Without loss of
h
h
f
generality, we let E
= and then the above equation simpli(cid:12)es to
h
(cid:14)E
E
= E
+ E
:
()
f
f
Considering the above relations, the angular correlation function of anisotropy of tempera-
ture of the CBR may be obtained
D
E
(cid:14)T (^x
)
(cid:14)T (^x
)
C ((cid:11)) =
(cid:1)
;
()
T
T
where cos((cid:11)) = ^x
(cid:1) ^x
and the averaging is done over all observation directions. Substituting
Eq.() into Eq.(), it follows that
C ((cid:11)) = hj E
(^x
) j
i + hj E
(^x
)E
(^x
) ji + hj E
(^x
) j
i:
()
f
f
f
f
In the right-hand side of Eq.(), the (cid:12)rst term is the angular correlation function of the
primordial (cid:13)uctuation background; the second describes the interaction between the gravita-
tional lensing e(cid:11)ect and the perturbed part of the primordial background (this is called the
angular cross correlation function C
((cid:11)), and determines the lensing e(cid:11)ects on the anisotropy
c
of the primordial CBR); the last is a higher order term and may reasonably be omitted. The
ma jor purpose of this paper is to determine the angular cross correlation function C
((cid:11)).
c
We Fourier decompose the gravitational potential, obtaining
Z
ik(cid:1)x
(cid:30)(x) =
(cid:1)
e
d
k:
( )
((cid:25) )
k
Based on linear perturbation theory and assuming a Gaussian random (cid:13)uctuation (cid:12)eld, we
have
(cid:0)
(cid:1)
=
H
k
(cid:14)
;
( )
k
k
h(cid:14)
(cid:14)
i =j (cid:14)
j
(cid:14)
(k
(cid:0) k
);
()
k
k
k
h(cid:1)
(cid:1)
i =
H
k
j (cid:14)
j
(cid:14) (k
(cid:0) k
);
()
k
k
k
(cid:0)
(cid:0)
h(cid:14)
(cid:1)
i =
H
k
j (cid:14)
j
(cid:14) (k
(cid:0) k
);
()
k
k
k
where j (cid:14)
j
and j (cid:1)
j
are the spectra of perturbations of the density and gravitational
k
k
potential respectively.
Roughly speaking, for adiabatic perturbations, the temperature (cid:13)uctuation of the CBR
on scales of a few arcminutes is (cid:14)T =T = (=)(cid:14)(cid:26)=(cid:26) before the recombination era (Silk
). What are the amplitude and shape of the temperature (cid:13)uctuations after decoupling?
Of course, this depends very much on the assumed recombination history. For the sake
of simplicity, we assume that the time-scale of recombination is extremely short so that
(cid:14)T =T = (=)(cid:14)(cid:26)=(cid:26) remains with the previous amplitude and shape. This assumption is rea-
sonable for our purposes since we are dealing with wave propagation and its interaction with
gravitational lensing between the last scattering surface and the observer. Our assumption,
in fact, just means choosing a convenient initial condition.
Fourier decomposing the (cid:13)uctuation part of the CBR and its scattering term, we obtain
Z
ik(cid:1)x
E
(^x
) =
(cid:14)
e
d
k;
()
f
k
((cid:25) )
Z Z Z
!
E
(^x
) =
(cid:14)
(cid:1)
e
e
cos(!x (cid:0) ! (cid:1) x)d
xd
k
d
k
;
()
f
k
k
ik
(cid:1)x
ik
(cid:1)x
((cid:25) )
where the vector x
points to the last scattering surface and has length H
; the unit
(cid:0)
vector ^x
represents the direction of observation.
For mathematical convenience, it is suitable to use a spherical harmonic analysis, which
is widely used when dealing with anisotropies on large angular scales:
m=+l
X
X
E
(
^
x
) =
A
Y
((cid:10));
()
f
lm
lm
l=
m=(cid:0)l
m=+l
X
X
E
(
^
x
) =
a
Y
((cid:10));
()
f
lm
lm
l=
m=(cid:0)l
and
Z
?
A
=
E
(
x
)Y
((cid:10))d(cid:10);
^
()
lm
f
lm
Z
?
a
=
E
(
x
)Y
((cid:10))d(cid:10):
^
( )
lm
f
lm
Substituting Eqs.() and () into Eqs.() and ( ) respectively and using the Rayleigh
equation
ik(cid:1)x
ikx cos (cid:13)
l
l=
X
e
= e
=
i
(l + )j
(kx)P
(cos (cid:13) )
( )
l
l
l=
and the addition expression
(cid:25)
m=+l
X
?
P
(cos(cid:13) ) =
Y
((cid:10)
)Y
((cid:10));
()
l
lm
k
lm
l +
m=(cid:0)l
where j
and P
are the l-th spherical Bessel function and the l-th Legendre function respec-
l
l
tively, we obtain the following equations:
Z
l
i
(cid:0)
A
=
(cid:14)
j
(kH
)Y
((cid:10)
)d
k;
()
lm
k
l
lm
k
(cid:25)
l
Z Z Z
i
!
(cid:0)
a
=
(cid:14)
(cid:1)
j
(k
H
)j
(k
x)Y
((cid:10)
)Y
((cid:10)
)
lm
k
k
l
l
lm
k
lm
k
(cid:25)
cos(!x (cid:0) ! (cid:1) x)
d
xd
k
d
k
:
()
x
According to random (cid:12)eld theory, for a Gaussian random (cid:12)eld, we have
h(cid:14)
(cid:14)
(cid:14)
(cid:14)
i = j (cid:14)
j
j (cid:14)
j
(cid:14) (k
(cid:0) k
)(cid:14) (k
(cid:0) k
)+ j (cid:14)
j
j (cid:14)
j
(cid:14) (k
(cid:0) k
)(cid:14) (k
(cid:0) k
)
k
k
k
k
k
k
k
k
+ j (cid:14)
j
j (cid:14)
j
(cid:14) (k
(cid:0) k
)(cid:14) (k
(cid:0) k
):
()
k
k
Combining the above expansions and integrating for angular coordinates within the cone
C (d(cid:10)), we (cid:12)nally obtain
?
<j a
j> = <j a
A
j>
lm
lm
lm
(cid:0)
H
Z
Z
H
(cid:0)
=
f[
j (cid:14)
j
j
(kH
)j
(kx)xdxdk ]
k
l
l
(cid:25) l
(cid:0)
(cid:0)
H
H
Z
Z
Z
Z
+
j (cid:14)
j
j (cid:14)
j
j
(k
H
)j
(k
x
)j
(k
x
)x
x
dx
dx
dk
dk
g;
()
k
k
l
l
l
(cid:0)
where the averaging is done over the entire sky and over all observation positions in the
universe. Finally, the angular cross correlation function, generated by interaction between
the gravitational lensing e(cid:11)ect and the primordial perturbed part of the CBR, is as follows:
X
C
((cid:11)) =
(l + ) <j a
j
> P
((cid:11)):
()
c
lm
l
(cid:25)
l=
For a double beam switch experiment, the observable (cid:14)T ((cid:11))=T , produced by gravitational
lensing, is
X
(cid:14)T ((cid:11))=T =
(l + ) <j a
j
>
( (cid:0) P
((cid:11))):
()
lm
l
(cid:25)
l=
IV. NUMERICAL RESULTS AND CONCLUSIONS
We assume that the (cid:13)uctuation spectrum is the Harrison-Zel'dovich spectrum when cal-
culating the gravitational lensing e(cid:11)ect on anisotropies of the CBR. For comparison with
observation, an appropriate normalization is necessary. We use the rms mass (cid:13)uctuation
((cid:14)M=M )
= within a sphere of radius r
= h
M pc as the normalization condition. Of
(cid:0)
course, how to select normalization conditions depends somewhat on how the mass distri-
bution traces the galaxy distribution. At present, this is not clear and so the selection of
the normalization condition may produce an uncertainty in the numerical results. However,
this does not signi(cid:12)cantly a(cid:11)ect our conclusion. With the above normalization condition,
the normalized power spectrum is given by
j (cid:14)
j
= r
k
()
k
In order to cancel the unknown nonlinear e(cid:11)ect, we need to apply a low-pass (cid:12)lter function
to truncate the spectrum. For mathematical convenience, we choose this as an exponential
form e
. Thus, for making calculations, the power spectrum is replaced by
(cid:0)kr
t
(cid:0)kr
t
j (cid:14)
j
= r
ke
:
( )
k
Here r
is a cut-o(cid:11) scale. Integrating Eq.() over k, we obtain
t
t
(H
r
)
r
H
r
+ H
+ x
(cid:0)
ha
i =
(
)
f[
Q
(
dx]
+
lm
l
(cid:0)
(cid:25)
l
H
x)
(cid:0)
H
Z
(cid:0)
(cid:0)
H
H
Z
Z
H
+ r
H
r
+ x
+ x
t
t
Q
(
)
Q
(
dx
dx
g
( )
l
l
x
x
)
with
R
j (cid:14)
j
d
k
k
k
l
=
= r
;
()
t
R
j (cid:14)
j
d
k
k
where Q
is the second kind of Legendre function. We have integrated Eq.( ) numerically
l
in the case of r
= h
M pc, r
= h
M pc and h = :. The behavior of the multipoles as
t
(cid:0)
(cid:0)
a function of the harmonic number l is showed in Fig.. The predicted cross function of the
temperature (cid:13)uctuation of the CBR due to the gravitational lensing e(cid:11)ect for comparison
with double beam switch experiments is showed in Fig.. We have also made numerical
calculations varying the parameters r
, r
and h within acceptable regions. However, the
t
numerical results are not sensitive to changes of these parameters. From Fig., it is obvious
that the e(cid:11)ect of gravitatioal lensing on anisotropies of temperature of the CBR on the scale
of arcminutes are too small to signi(cid:12)cantly amplify or depress the primordial anisotropies of
the CBR. We can then safely conclude that the gravitational lensing e(cid:11)ect on anisotropies
of temperature of the CBR on scales of a few arcminutes are negligible. Our conclusion
disagrees strongly with that of Cayon et al. who argue that gravitational lensing, produced
by linear density perturbations, may enhances (cid:1)T =T by (cid:24) % or even more.
In addition, the problem in the calculation presented by Cay(cid:19)on et al. ( ) could be
that they considered the e(cid:11)ect on the anisotropies of the CBR caused only by de(cid:13)ections
of light lines, but not by convergences and divergences of light beams. In principle, the last
e(cid:11)ect could also produce a considerable change in the (cid:13)uctuation of the intensity of the CBR
compared with that caused by de(cid:13)ections of light lines.
Acknowledgements.
We thank the anonymous referee for his enlightening comments which resulted in im-
portant improvements of the paper. We appreciate Prof. J.Miller for his carefully reading
our manuscript which improved its readability. J.M.L. is grateful to Dr. L.To(cid:11)olatti, Dr.
P.Mazzei, Dr. N.Di.Cicco and especially to Prof. G.De Zotti and Prof. L.Danese for their
hospitality during his stay in Padova. He also acknowledges (cid:12)nancial support from SISSA
and from the Osservatorio Astronomico di Padova where part of this work was done.
REFERENCES
[] Blanchard,a.,& Schneider,J. , A&A, ,
[] Cay(cid:19)on,L., Gonz(cid:19)alez,M.E.& Sanz,L.J. , ApJ, ,
[] Cole,S.& Efstathiou,G., , MNRAS, ,
[] Fang,L.Z., , Scientia Sinica, A,
[] Feng,L.L.& Liu,J.M., , A&A, ,
[] Gaier,T., et al., , ApJ, , L
[] Hanni,R.S., , Phys.Rev., ,
[] Holtzman,J.A., , ApJ.Supp., ,
[ ] Kashlinsky,A., , ApJ, , L
[ ] Linder,E.V., , MNRAS, ,
[] Readhead,A.C.S., Lawrence,C.R., Myers,S.T., Sargent,W.L.W.
& Mo(cid:11)et,A.T., , ApJ, ,
[] Sasaki,M., , MNRAS, ,
[] Silk,J., , Nature, ,
[] Tomita,K.& Watanabe,K., , Prog.Theor.Phys., ,
[] Uson,J.M.& Wilkinson,D.T., , ApJ, , L
[] Watanabe,K.& Tomita,K., , ApJ, ,
|
astro-ph/0605272 | 1 | 0605 | 2006-05-10T20:04:05 | High-Resolution Infrared Imaging of Herschel 36 SE: A Showcase for the Influence of Massive Stars in Cluster Environments | [
"astro-ph"
] | We present high-resolution infrared imaging of the massive star-forming region around the O-star Herschel 36. Special emphasis is given to a compact infrared source at 0".25 southeast of the star. The infrared source, hereafter Her 36 SE, is extended in the broad-band images, but features spatially unresolved Br gamma line emission. The line-emission source coincides in position with the previous HST detections in H alpha and the 2 cm radio continuum emission detected by VLA interferometry. We propose that the infrared source Her 36 SE harbors an early B-type star, deeply embedded in a dusty cloud. The fan shape of the cloud with Herschel 36 at its apex, though, manifests direct and ongoing destructive influence of the O7V star on Her 36 SE. | astro-ph | astro-ph | Draft version August 21, 2017
Preprint typeset using LATEX style emulateapj v. 6/22/04
6
0
0
2
y
a
M
0
1
1
v
2
7
2
5
0
6
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
A SHOWCASE FOR THE INFLUENCE OF MASSIVE STARS IN CLUSTER ENVIRONMENTS 1,2
HIGH-RESOLUTION INFRARED IMAGING OF HERSCHEL 36 SE:
M. Goto,3 B. Stecklum,4 H. Linz,3,4 M. Feldt,3 Th. Henning,3 I. Pascucci,3,5, T. Usuda6
Draft version August 21, 2017
ABSTRACT
We present high-resolution infrared imaging of the massive star-forming region around the O-star
Herschel 36. Special emphasis is given to a compact infrared source at 0.′′25 southeast of the star. The
infrared source, hereafter Her 36 SE, is extended in the broad–band images, but features spatially
unresolved Br γ line emission. The line-emission source coincides in position with the previous HST
detections in H α and the 2 cm radio continuum emission detected by VLA interferometry. We
propose that the infrared source Her 36 SE harbors an early B-type star, deeply embedded in a dusty
cloud. The fan shape of the cloud with Herschel 36 at its apex, though, manifests direct and ongoing
destructive influence of the O7V star on Her 36 SE.
Subject headings: circumstellar matter - dust, extinction - planetary systems: protoplanetary disks
- early-type - stars: formation - stars: individual (Herschel 36, G5.97−1.17)
1. INTRODUCTION
Massive stars are the primary source of radiation, ki-
netic energy, and chemical enrichment in the interstel-
lar medium, playing a pivotal role in galactic evolution.
Because of their remote locations, our understanding of
their formation has been limited by the lack of high
resolution techniques. This challenge has been under-
taken by adaptive optics systems at large-aperture tele-
scopes. The present work is part of a coordinated effort
to understand the formation of high-mass stars by us-
ing state-of-the-art instruments available at the VLT and
other telescopes (Feldt et al. 1999; Henning et al. 2001;
Feldt et al. 2003; Grady et al. 2004; Pascucci et al. 2004;
Puga et al. 2004; Linz et al. 2005; Apai et al. 2005).
1997; but see also Arias et al.
Herschel 36 is located in a high-mass star-forming re-
gion at a distance of 1.8 kpc from us (van den Anker
et al.
2006) near
the center of M 8. The bright central part of M 8 is
called the Hourglass. The Hourglass is a cavity of ion-
ized gas seen through the gaps between the foreground
obscuration (Woodward et al. 1986). Herschel 36, an
O7V star (Woolf 1961),
is responsible for the ioniza-
tion of the gas in the cavity. The inferred dynamical
age of the ionized gas and, therefore, the age of the
Hourglass and Herschel 36, is as small as ∼5×104 yrs
(Chakraborty & Anandarao 1997).
In the present study a special focus is placed on the
infrared source found at a distance of 0.′′25 southeast
of Herschel 36. The extended source called hereafter
Electronic address: [email protected]
1 Based on data collected in the course of NACO and MIDI guar-
anteed time observations (71.C-0143(A) and 73.C-0175(A)) at the
VLT on Cerro Paranal (Chile), which is operated by the European
Southern Observatory (ESO).
2 Based on data collected at Subaru Telescope, which is operated
by the National Astronomical Observatory of Japan.
3 Max Planck Institute for Astronomy, Konigstuhl 17, D-69117
Heidelberg, Germany.
4 Thuringer Landessternwarte Tautenburg, Sternwarte 5, D-
07778 Tautenburg, Germany.
5 The University of Arizona, Department of Astronomy/Steward
Observatory, 993 N Cherry Ave. Tucson, AZ 85721-0065.
6 Subaru Telescope, 650, North A'ohoku Place, Hilo, HI 96720.
Her 36 SE was first recognized by Stecklum et al. (1995)
by means of
lunar occultation measurements. Her-
schel 36 had long been known as a peculiar early-type
star with substantial mid-infrared excess (Woolf 1973;
Allen 1986).
It is only after Her 36 SE was spatially
resolved that we know that this object is actually the
source responsible for the excess infrared emission. Af-
ter the discovery of Stecklum et al. (1995), the possible
identity of Her 36 SE has been discussed, including an
externally ionized protoplanetary disk, or, a proplyd, an
obscured embedded source, and a leftover circumstellar
disk of Herschel 36; however, no solid conclusion was
reached.
In the next section, the observations at the VLT and
supplemental spectroscopy at the Subaru Telescope are
described. The direct consequences of the observations
are summarized in §3. In §4 we will further discuss the
possible nature of Her 36 SE as a deeply embedded early-
type star under the violent influence of the nearby O-star
Herschel 36.
2. OBSERVATION AND DATA REDUCTION
2.1. Thermal Near-infrared Imaging
The infrared imaging at L′ (3.8 µm) and M ′ (4.7 µm)
was carried out at the VLT UT4 on 2003 June 11 with
the adaptive optics imager NACO (Rousset et al. 2000;
Lenzen et al. 2003; Hartung et al. 2003). Herschel 36
(V = 9.1 mag) served as a wavefront reference source
for the visible wavefront sensor in the adaptive optics
system. A short exposure of 180 ms was repeated 27
times in the L′ imaging at each of the 9 positions of the
telescope dithering. Imaging at M ′ was performed the
same way, but with a shorter integration time of 56 ms
repeated 89 times. The total on-source integration time
is 58 s and 60 s at L′ and M ′, respectively. The observing
log is presented in Table 1 including imaging with other
filters and additional spectroscopy.
The imaging data were reduced in the standard man-
ner. After sky-subtraction and flat-fielding,
images
were registered referring to the position of Herschel 36.
The size of the isoplanatic patch was measured using
more than 40 stars inside the entire field of view of
2
Goto et al.
posite image of L′, M ′ and the narrow-band image at
2.17 µm described in the next section is shown in Fig-
ure 1.
2.2. Narrow-band Imaging at Br γ
Narrow-band Br γ images (2.17 µm) were obtained
during the same night together with continuum images
at 2.18 µm. Short exposures of 1 s were repeated 20
times to minimize the saturation of bright stars. The
total integration time on source is 100 s. The sky at
30′′ distance from the field was recorded for background
subtraction. The data were processed in the same way
as it was the case for broadband imaging. Aperture pho-
tometry was performed for Her 36 SE in Br γ and the
continuum at 2.18 µm. Since Herschel 36 is saturated,
Her 36 B was used to establish the correct flux scale
of the images (K = 9.4 mag; KS1 in Woodward et al.
1990, 18006nr766 in Bik 2004). The PSF sampled from
Her 36 B was scaled and subtracted from Herschel 36 to
isolate the extended emission of Her 36 SE. The total
pixel counts were summed up inside a circular aperture
of 1.′′3 centered on Her 36 SE. Note that the Br γ photom-
etry presented in Table 2 is before the continuum sub-
traction. The accuracy of the photometry is ∼0.1 mag
for the images with the underlying continuum emission.
The continuum image was scaled and subtracted so that
the pixel counts of blue stars (with respect to their in-
frared colors) around Herschel 36 are equally canceled in
the line-emission image.
2.3. Mid-infrared Imaging at 8.7 µm and 12.8 µm
Fig. 2.- Left: MIDI source-acquisition images at 12.8 µm for
Her 36 SE (top) and the flux calibration star HD 169916 (bot-
tom). Herschel 36 is no longer visible at this wavelength. While
HD 169916 has multiple diffraction rings around, Her 36 SE is ex-
tended with no trace of a point source. Right: The radial profile
of Her 36 SE and HD 169916. The size of the emitting region at
Her 36 SE is 850 AU in diameter, after deconvolution with the
spatial profile of HD 169916.
Fig. 1.- Color composite image of Herschel 36. M ′ is color coded
in red, L′ in green, and Br γ in blue before continuum subtraction.
Her 36 SE is the red extended emission at 0.′′25 southeast of Her-
schel 36. Her 36 B is an infrared star at 3.′′6 north of Herschel 36.
The coordinates of Herschel 36 are R.A. (J2000) 18h03m40.s20, Dec
(J2000) −24◦22′43.′′0 (Maiz-Apellaniz et al. 2004).
NACO (27′′×27′′). The measurements have been done
at 2.2 µm, since the isoplanatic patch becomes smaller
with the wavelength. The point spread function (PSF) is
found elongated only at the edge of the field of view, and
no significant degradation of PSF is recognized within the
field of view relevant to the following discussion shown in
Figure 1. The full width at half maximum (FWHM) of
point sources are 0.′′11 at L′ and 0.′′13 at M ′ in the fully
reduced images.
The absolute flux calibration was tied to the photom-
etry of Herschel 36 as given in the literature. The zero
point magnitudes were calculated to be consistent with
the photometric magnitudes L′ = 6.3 mag and M ′ =
3.8 mag in Woodward et al. (1990) and Woolf (1973), re-
spectively. The images were convolved with Gaussian fil-
ter at the zero-point calculation to match the spatial res-
olution in the previous observations. The photometry of
Herschel 36 was performed inside a small aperture of 0.′′2
to avoid confusion with Her 36 SE. An aperture correc-
tion was applied by using the PSF sampled at Her 36 B
(see Figure 1). The photometry of Her 36 SE was then
performed in a circular aperture of 1.′′3, after the con-
tribution of Herschel 36 was removed by subtracting the
scaled PSF. The primary error source in the photometry
is the spatial fluctuation of the background emission in
the immediate vicinity of Herschel 36. We restored the
flux compensation function with varying outer bounds
from 0.′′6 to 1.′′4, and found that the amount of aper-
ture correction does not differ more than 15 %. The sky
level sampled at different locations at 0.′′9 to 2.′′6 around
the object does not change the net photometry by more
than 14 %. We therefore quote 0.2 mag as the photomet-
ric accuracy, although the formal error is much smaller
(<0.05 mag). The results are presented in Table 2 with
other photometry obtained in this paper. The color com-
Infrared Imaging of Herschel 36 SE
3
TABLE 1
UT Date
Filter/Band
λ [µm]
∆λa [µm]
Telescope
Instrument
Observation
Spectral Resolution
2003 Jun 11
2003 Jun 11
2003 Jun 11
2003 Jun 11
2004 Jun 3
2004 Jun 3
2004 Jul 29
aIn FWHM.
L′
M ′
Br, γ
IB 218
N8.7
[Ne II]
Br α
3.80
4.78
2.17
2.18
8.64
12.8
4.05
0.62
0.59
0.023
0.060
1.55
0.39
VLT
VLT
VLT
VLT
VLT
VLT
Subaru
NACO
NACO
NACO
NACO
MIDI
MIDI
IRCS/AO
Imaging
Imaging
Imaging
Imaging
Imaging
Imaging
Spectroscopy
R = 10, 000
Mid-infrared images in the N 1 filter (8.7 µm) and in
the [Ne II] filter (12.8 µm) were obtained on 2004 June 3
at the VLT with MIDI (Leinert et al. 2003). The instru-
ment is an interferometer/spectrometer, but was used as
mid-infrared camera in the present observation. The tip-
tilt corrector STRAP was used to stabilize the images.
These thermal infrared data were recorded by using a
chopping throw of 10′′. The total on-source integration
is 80 s at 8.7 µm and 375 s at 12.8 µm. The data reduc-
tion was carried out using the pipeline provided by the
MIDI consortium7. The spatial resolution of the final
image is nearly diffraction limited (Fig. 2).
The emission from Her 36 SE is clearly extended both
in N 1 and [Ne II]. Herschel 36 is no longer visible at
mid-infrared wavelengths, as is expected from its pho-
tospheric spectral energy distribution (SED). It is clear
that the peculiar mid-infrared excess toward Herschel 36
(Dyck 1977) is not from the O-star itself, but is almost
entirely attributed to Her 36 SE. The flux calibration
was performed with respect to the photometric stan-
dard HD 169916 for which the absolute flux density was
taken from Cohen et al. (1999). The photometry was
performed inside a 1.′′8 aperture centered on Her 36 SE.
The size of the aperture is slightly larger than that is
used in the shorter wavelengths. The smaller aperture
at the thermal near-infrared is because there seem to
be two overlapping emission contributions at Her 36 SE;
the compact dusty cloud at Her 36 SE itself, and the
filamentary emission more connected to the diffuse emis-
sion at 2′′ southeast of Herschel 36. Since we will discuss
an internal source inside Her 36 SE below, the diffuse
emission should be excluded not to overestimate its lu-
minosity. On the other hand, the mid-infrared images
by MIDI do not show any clear hints of multiple sources,
we therefore use a safe oversized aperture not to lose the
mid-infrared flux of Her 36 SE for later discussion of its
energetics.
2.4. Medium-Resolution Br α Spectroscopy
Supplemental 4-micron spectroscopy (R = 10,000)
was performed at the Subaru Telescope on 29 July
2004 with IRCS (Tokunaga et al. 1998; Kobayashi et al.
2000). The slit was oriented along a position angle of
110◦ counted from north to east to cover Herschel 36 and
SE at the same time. An adaptive optics system was
used to attain higher spatial resolution (Gaessler et al.
2002; Takami et al. 2004). The sky at 2′ north of Her-
schel 36 was observed for background subtraction after
Fig. 3.- (a) Continuum-subtracted Br γ image of Her 36 SE. The
location of the point source is marked by ticks. The gray contour
is from the narrow-band continuum image at 2.18 µm presented
darker at around Herschel 36 for clarity. No point source is detected
at the continuum wavelength. (b) The H α image retrieved from
the HST archive. The PSF generated by TinyTim is subtracted.
The contour in black is radio continuum emission at 2 cm obtained
with the VLA (see Stecklum et al. 1998 for observational detail).
The unresolved sources at Br γ, H α, and radio wavelengths are all
in positional agreement. The coordinate offsets are with respect to
Herschel 36.
7 http://www.mpia-hd.mpg.de/MIDISOFT/
each on-source integration. The spectroscopic flat field
4
Goto et al.
Fig. 4.- Left: A close-up view of the two–dimensional spectrogram near the Br α emission. The continuum emission at Her 36 SE is
clearly separated from that of Herschel 36, and apparently extended (position A and C). The line emission from the ambient nebulosity is
seen at Br α (position B). The white line traces the wavelength centroids of the Br α emission along the slit. The emission line is slightly
(∼2 km s−1) blueshifted at Her 36 SE with respect to the ambient nebular emission. Right: Cross-cuts of the spectrogram along the slit,
at the continuum (A and C; gray polygon), and the Br α emission (B; thin solid line). The contribution of the nebulosity is subtracted at
the position B. Her 36 SE shows a sharp emission line at Br α on top of the extended continuum emission. The line emission profile, after
subtracting the continuum profile, is close to that of Herschel 36, which means that the Br α emission at Her 36 SE is also compact and
point-like.
was obtained from a halogen lamp exposure at the end
of the night.
One dimensional spectra of Herschel 36 and SE were
extracted with the aperture-extraction package of IRAF8
after the sky-subtraction, flat-fielding, and the interpo-
lation of the bad pixels were applied. The wavelength
calibration was carried out using the atmospheric trans-
mission curve modeled by ATRAN (Lord 1992). The
Br α line-emission was calibrated to the photometry of
Herschel 36 at L′. First, to correct the continuum slope,
the one-dimensional spectrum of Herschel 36 was divided
by the spectroscopic standard star HR 7121 (B2.5V),
and was multiplied by a blackbody spectrum of the tem-
perature corresponding to the effective temperature of a
B2.5V star (Teff =19,000 K; Crowther 2005). The spec-
trum with correct slope was then scaled so that the aver-
aged flux density inside the L′ bandpath (3.49–4.11 µm)
is equal to the L′-band photometry of Herschel 36. The
same conversion factor was used to calibrate the spectral
flux of Her 36 SE reduced in the same way with Her-
schel 36. The Br α luminosity at Her 36 SE is found
to be L(Br α) =(1.4–1.6)×1024 W at the distance of
1.8 kpc, after the continuum and the surrounding diffuse
emission is subtracted. The error interval is given by the
difference of the two sequential measurements, although
it is subject to the uncounted uncertainty associated to
the possible vignetting by the narrow slit of 0.′′3.
3. RESULTS
8 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.
An unresolved source is detected at the location of
Her 36 SE in the continuum-subtracted image at Br γ
(Fig. 3). The diameter of the point-like source is less than
130 AU from the diffraction-limited spatial resolution of
NACO (0.′′072 in FWHM at 2.17 µm). The HST/PC2
image retrieved from the ST-ECF9 archive shows a com-
pact H α emission at the same location (Fig. 3). The
H α emission is unresolved as well. Considering the plate
scale of PC2 (0.′′046 pixel−1), this finding indicates that
the source is less than 100 AU across. Furthermore, ra-
dio interferometric observations have been carried out
with the VLA at a wavelength of 2 cm (see Stecklum et
al. 1998 for the observational detail). A compact radio
source is found at the same location as the position of the
Br γ and H α emission. The radio source is unresolved
with regard to the synthesized beam size of 0.′′16, which
translates to 290 AU.
Another line of evidence for a point-like source comes
from the spectroscopy. The two-dimensional spectro-
gram near Br α is shown in Figure 4. Her 36 SE shows
distinct line emission in Br α, slightly blueshifted from
ambient nebular emission by 2 km s−1, with a spatial
profile apparently narrower than the continuum emission.
The sharp spatial profile is comparable to that of Her-
schel 36, which corroborates the presence of a point-like
source in the hydrogen line emission at the location of
Her 36 SE.
On the other hand, Her 36 SE is clearly extended in
9 The Space Telescope European Coordinating Facility (ST-
ECF), jointly operated by ESA and the European Southern Ob-
servatory (ESO), is the European HST science facility, supporting
the European astronomy community in exploiting the research op-
portunities provided by the Hubble Space Telescope.
Infrared Imaging of Herschel 36 SE
5
the continuum emission at wavelengths from 2 to 13 µm.
If we use the spatial profile of HD 169916 as the in-
strumental PSF, and deconvolve Her 36 SE by inverting
simple square sum, the extent of the emitting source at
Her 36 SE is reduced to 0.′′47 at 12.8 µm, which is 850 AU
in diameter at the distance of the object (Fig. 2). The
SED of Her 36 SE is presented in Figure 5 to character-
ize the extended emission. The color temperature of the
continuum source clearly points to the existence of warm
dust at the location of Her 36 SE. No point–like substruc-
ture is found in the broad–band images of Her 36 SE that
could have corresponded to the unresolved line emission.
4. DISCUSSION – NATURE OF HERSCHEL 36 SE
Here we first discuss the identity of the line emission
source and its possible ionization mechanism, including
external ionization by Herschel 36, an embedded low- to
intermediate-mass star in its active accretion phase, and
an H II region internally ionized by an early-type star.
The Br γ emission is apparently inside the dusty cloud
at Her 36 SE, since it is spatially more confined than
the continuum emission. In addition, there is no hint of
rim-ionization detected in Br γ emission at the side of
Her 36 SE where it faces toward Herschel 36. We found
no solid evidence that Herschel 36 plays a direct role to
externally ionize the unresolved source inside Her 36 SE.
The radio flux at 2 cm (Fν =1.3 mJy) is probably too
high for an accretion signature of an intermediate-mass
star at 1.8 kpc away. Neufeld & Hollenbach (1996) have
calculated the free-free emission arising from an accretion
shock in dependence of (proto)stellar mass and accretion
rate. However, even with their most extreme setup (M
Macc = 10−4M⊙yr−1) they just reach a 3.6 cm
= 10 M⊙,
flux of roughly 3 mJy for a source 100 pc away. Extrap-
olated to λ = 2 cm (by optimistically assuming that the
ionized gas is completely optically thick with Fν ∼ ν2)
and scaled to a distance of 1.8 kpc, the expected 2-cm
flux would be just some 30 µJy, around 40 times less than
the measured value.
It is therefore inferred that a star with an early spectral
type exists inside Her 36 SE that gives rise to a small H II
Fig. 5.- The infrared SED of Herschel 36 and SE, plotted by the
circles and the squares, respectively. The filled symbols are from
the present observations, and the crosses and open symbols are
from the literature. Note that this previous photometry has been
done with much larger apertures. The color temperature of the
dust emission measures 400 K from 3–13 µm photometry, which
is taken as the upper limit of the dust emission temperature of
Her 36 SE.
region responsible for the radio emission. The H II region
is internally ionized, but is kept compact because of the
high density of the enshrouding cloud.
If we take the
VLA beam size as the physical dimension of the H II
region (r ∼140 AU), the number of Lyman continuum
photons required to maintain the ionized gas is 1.6×1045
s−1. The Lyman continuum photon-rate is reproducible
only by a star earlier than B2 if the star is at the zero-
age main sequence (Panagia 1973; Crowther 2005). On
the other hand, in order not to create a parsec-scale H II
region, the hydrogen density has to be as high as nH >
106 cm−3.
We use the K-band extinction toward the hypotheti-
cal early-type star to estimate whether the gas density
is sufficiently high to confine the H II region. A B2 star
at the zero-age main-sequence should have a relative K-
band brightness of 9.9 mag at the distance of Herschel 36
without any attenuation ((V − K)0 = −0.9 mag from
Ducati et al. 2001). The sensitivity of our observation
at 2 µm is 15.7 mag for a 3 sigma detection at the loca-
tion of Her 36 SE. With non-detection of any continuum
point source at this wavelength, the dust extinction must
be larger than 6 mag at 2 µm, which translates to AV >
60 mag after correcting AV ≈ 5 mag for the foreground
extinction toward the Herschel 36 region (Stecklum et al.
1998). The visible extinction can be related to a hy-
drogen column density (e.g., Mathis 1990; Ryter 1996).
Provided that the dusty core of Her 36 SE is spherical,
of constant density, and 850 AU across as is measured
in the 12.8 µm image; the mass in the obscuration is
MSE ≥ 1.7 × 10−2M⊙ with nH ≥ 1.8 × 107 cm−3. Thus,
the gas density should be high enough to keep the H II
region spatially unresolved.
Lν/hν dν ≈ αB/αeff
Br α · L(Br α)/hνBr α.
The Lyman photon rate derived from Br α spec-
troscopy is also consistent with that of an early B
type star. The ionizing flux in an H II region is ob-
tained from Br α line flux by scaling the photon number-
count proportionally to the recombination coefficients,
Rν0
If we take
αeff
Br α = 1.085 × 10−14cm3 s−1 and αB = 2.658 ×
10−13cm3 s−1 from Storey & Hummer (1995) for the
Case B of Te = 104 K and Ne = 107 cm−3, the Lyman
continuum rate turns out (6.9–8.0) ×1044 s−1, which is
reproducible by a B2.5 dwarf (Crowther 2005). The cor-
rection of the foreground extinction needs caution, since
the dust obscuration is increasingly transparent in the
longer wavelength (Rieke & Lebofsky 1985); but if we
use AK = 6 mag as a face value, the intrinsic Lyman
continuum rate is (4.1–4.8) ×1045 s−1, which is still con-
sistent with the ionizing photon rate of a B1–B1.5 dwarf.
The infrared luminosity of Her 36 SE is consistent both
with an internal B2 star, and also with external heating
by Herschel 36. The infrared luminosity from 2 to 40 µm
is calculated from the NACO and MIDI photometry with
the flux density at the longer wavelength extrapolated as
F (λ) = κ(λ) Md d−2 B(Td, λ); where d is the distance to
the object, and κ(λ) is the computed mass absorption co-
efficient for the grains without ice mantles coagulated in
the protostellar cores (Ossenkopf & Henning 1994). The
total infrared luminosity is LIR = 400 L⊙ at an assumed
distance of 1.8 kpc insensitive to the gas density of the
core n = 106 cm−3 to 108 cm−3. It is therefore well re-
producible either by the luminosity of a B2 star at the
6
Goto et al.
TABLE 2
Filter Fλ
Br γ
IB 218
L′
M ′
N8.7
Ne II
Herschel 36
Her 36 SE a
b [10−14 W m−2 µm−1] Fλ
b [10−14 W m−2 µm−1] PSF in FWHM
· · ·
· · ·
5.1
5.2
· · ·
· · ·
2.2 c
2.0
7.9
21
54
12
0.′′072
0.′′071
0.′′11
0.′′13
0.′′35
0.′′37
aThe photometric apertures are 1.′′3 in the mid-infrared bands
(N8.7 and Ne II), and 1.′′8 in all the others.
bThe photometric errors are typically 10% at 2.2 µm, and 20% at
all the other filters.
cContinuum is not subtracted.
zero-age main-sequence (L∗ = 3 × 103 L⊙), or by Her-
schel 36 (L∗ = 105 L⊙) while the solid angle subtended
by Her 36 SE is of the order of unity at the location of
Herschel 36. The dust emitting temperature is 400 K
which should be taken as the upper limit, for the lack of
additional photometry at the longer wavelengths.
We conclude by comparing Her 36 SE with two sim-
ilar cases reported to date in which bright mid-infrared
sources are found in the immediate (projected) vicin-
ity of O-type stars. The infrared source SC3 has been
found at 1.′′8 (810 AU) west of θ1 Ori C, the pri-
mary illumination source of the Orion Trapezium Clus-
ter (Hayward, Houch, & Miles 1994). SC3 is spatially
resolved, measuring 1.′′5 across, however, despite the
apparent proximity to θ1 Ori C (O5.5V), its appear-
ance is barely distorted, almost with a perfect circu-
lar symmetry.
It is thus proposed that SC3 is a pro-
plyd seen face-on, located deep behind θ1 Ori C with
the physical separation much larger than the apparent
projection (Robberto, Beckwith, & Panagia 2002). SC3
is visible in the optical (e.g., Smith et al. 2005), and
at near-infrared wavelengths (McCaughrean & Stauffer
1994), which also lends support to its proplyd nature.
The infrared appearance of SC3 is in strong contrast to
Her 36 SE. We may use the highly distorted dust emis-
sion from Her 36 SE as circumstantial evidence that the
source is actually under the influence of Herschel 36, and
that the physical distance to the O star is not signifi-
cantly larger than it appears.
On the other hand, σ Ori IRS 1, found next to σ Ori-
onis, shares a similar morphology with Her 36 SE. It is
a compact infrared source at 1200 AU away from the
O9.5V star, and features a fan-shaped emission with
σ Orionis at the apex (van Loon & Oliveira 2003), ex-
actly as Herschel 36 is to SE. The mid-infrared spectrum
of σ Ori IRS 1 shows the partly crystalline silicate in
emission. The presence of processed silicates suggests
significant grain growth which is naturally present in a
circumstellar disk. A proplyd is therefore again the most
probable cause of σ Ori IRS 1, especially because a cen-
tral star has been detected recently in the K-band con-
tinuum emission (B. Stecklum, private communication).
The star-forming region around Herschel 36 has many
features in common with the Orion Nebula Cluster.
The local concentration of massive stars, like Her 36 B,
together forms a Trapezium-like cluster around Her-
schel 36.
The presence of known proplyd nearby
at G5.97−1.17 (Stecklum et al. 1998) underscores the
physical similarity as well. The mid-infrared color of
Her 36 SE, and the close vicinity to an O-type star with
a distorted morphology suggestive of radiative influence
of it all point toward Her 36 SE is also a proplyd with a
low-mass star at its center, as is the case for SC3 at
θ1 Ori C and σ Ori IRS 1. However, in addition to
the radio luminosity hardly accounted for by a low-mass
star, and no ionized-front outside the dusty cloud; a pro-
plyd cannot explain the lack of a point source to be de-
tected at the continuum wavelengths that comes from
the photospheric emission of the star.
In the case of
no central star inside, another possibility would be that
Her 36 SE is an evaporating gaseous globule, a failed pro-
plyd without an internal star in formation. These star-
less cores have been detected in numbers toward M 16
(McCaughrean & Andersen 2002). However, a hypothet-
ical starless globule conflicts with the presence of the un-
resolved Br γ emission apparently inside of Her 36 SE.
We therefore propose that Her 36 SE harbors a rela-
tively massive star of early B-type producing a squeezed
H II region inside the dusty cloud (Keto 2003), but the
star itself is completely obscured. The distortion of dust
emission as well as the diffuse emission downstream of
Her 36 SE, indicates the close physical interaction of Her-
schel 36 and SE. Herschel 36 SE, now in the process of
being blown away, is a showcase for the violent impact of
the dominant O-star in a cluster on another early-type
star nearby.
We thank all the staff and crew of the VLT and Subaru
for their valuable assistance in obtaining the data, and
Thorsten Ratzka, Elena Puga and Wolfgang Brandner
for their indispensable help in reducing data. We appre-
ciate the anonymous referee for many critical comments
that are necessary to improve the paper. M.G. is sup-
ported by Japan Society for the Promotion of Science
fellowship.
Arias, J. I., Barb´a, R. H., Ma´iz Apell´anz, J., Morrell, N. I., &
Rubio, M. 2006, MNRAS, 366, 739
Apai, D., Linz, H., Henning, Th., & Stecklum, B. 2005, A&A, 434,
987
REFERENCES
Infrared Imaging of Herschel 36 SE
7
Allen, D. A. 1986, MNRAS, 219, 35
Bik, A. 2004, PhD thesis, Universiteit van Amsterdam
Charkraborty, A., & Anandarao, B. G. 1997, AJ, 114, 1576
Cohen, M., et al. 1999, AJ, 117, 1864
Crowther, P. A., Massive Star Birth: A Crossroads of Astrophysics,
International Astronomical Union. Symposium no. 227, 389,
on May 2005 in Italy, eds. R. Cesaroni, M. Felli, E.
Churchwell, M. Walmsley, Cambridge: Cambridge University
Press, astro-ph/0506324
Ducati, J. R., Bevilacqua, C. M., Rembold, S. B., & Ribeiro, D.
2001, ApJ, 558, 309
Dyck, H. M. 1977, AJ, 1977, 82, 129
Feldt, M., Stecklum, B., Henning, Th., Launhardt, R., & Hayward,
T. L. 1999, A&A, 346, 243
Feldt, M., Puga, E., Lenzen, R., Henning, Th., Brandner, W.,
Stecklum, B., Lagrange, A.-M., Gendron, E., & Rousset, G. 2003,
ApJ, 599, L91
Gaessler, W., et al. 2002, Proc. SPIE, 4494, 30
Grady, C. A., Woodgate, B., Torres, Carlos A. O., Henning, Th.,
Apai, D., Rodmann, J., Wang, Hongchi, Stecklum, B., Linz,
H., Williger, G. M., Brown, A., Wilkinson, E., Harper, G. M.,
Herczeg, G. J., Danks, A., Vieira, G. L., Malumuth, E., Collins,
N. R., & Hill, R. S. 2004, ApJ, 608, 809
Hartung, M., Lenzen, R., Hofmann, R., Bohm, A., Brandner, W.,
Finger, G., Fusco, T., Lacombe, F. Laun, W., Ganier, P., Storz,
C., & Wagner, K. 2003, Proc. SPIE, 4841, 425
Hayward, T. L, Houch, J. R., & Miles, J. W. 1994, ApJ, 433, 157
Henning, Th., Feldt, M., Stecklum, B., & Klein, R. A&A, 2001,
370, 100
Kobayashi, N., et al. 2000, Proc. SPIE, 4008, 1056
Keto, E. 2003, ApJ, 599, 1196
Leinert, C., Graser, U., Waters, L. B. F. M., Perrin, G. S., Walter,
J., Bruno, L., Przygodda, F., Chesneau, O., Schuller, P. A.,
Grlazenborg-Kluttig, Annelie, W., Laun, W., Ligori, S., Meisner,
J. A., Wagner, K., Bakker, E. J., Cotton, B.; de Jong, J., Mathar,
R., Neumann, U., & Storz, C. 2003, Proc. SPIE 4838, 893
Lenzen, R., Hartung, M., Brandner, W., Finger, G., Hubin, N. N.,
Lacombe, F., Lagrange, A.-M., Lehnert, M. D., Moorwood, A.
F. M., & Mouillet, D. 2003, Proc. SPIE, 4841, 860
Linz, H., Stecklum, B., Henning, Th., Hofner, P., & Brandl, B.
2005, A&A, 429, 903
Lord, S. D. 1992, A New Software Tool for Computing Earth's
Atmosphere Transmissions of Near- and Far-Infrared Radiation,
NASA Technical Memoir 103957 (Moffett Field, CA: NASA
Ames Research Center)
Maiz-Apellaniz, J., Walborn, N. R., Galue H. A., Wei, L. H. 2004,
ApJS, 151, 103
Mathis, J. S. 1990, ARA&A, 28, 37
McCaughrean, M. J., & Stauffer, J. R. 1994, AJ, 108, 1382
McCaughrean, M. J., & Andersen, M. 2002, A&A, 389, 513
Neufeld, D. A., & Hollenbach, D. J. 1996, ApJ, 471, L45
Ossenkopf, V., & Henning, Th. 1994, A&A, 291, 943
Panagia, N. 1973, AJ, 78 929
Pascucci, I., Apai, D., Henning, Th., Stecklum, B., & Brandl, B.
Puga, E., Alvarez, C., Feldt, M., Henning, Th., & Wolf, S. 2004,
2004, A&A, 426, 523
A&A, 425, 543
897
Rieke, G. H., & Lebofsky, M. J. 1985, ApJ, 288, 618
Robberto, M., Beckwith, S. V. W., & Panagia, N. 2002, ApJ, 578,
Rousset, G., Lacombe, F., Puget, P., Gendron, E., Arsenault, R.,
Kern, P. Y., Rabaud, D., Madec, P.-Y., Hubin, N. N., Zins, G.,
Stadler, E., Charton, J., Gigan, P., Feautrier, P. 2000, Proc.
SPIE, 407, 72
Ryter, Ch. E. 1996, Ap&SS, 236, 285
Smith, N., Bally, J., Shuping, R. Y., Morris, M., & Kassis, M. 2005,
AJ, 130, 1778
Stecklum, B., Henning, Th., Eckart, A., Howell, R. R., & Hoare,
M. G. 1995, ApJ, 445, L153
Stecklum, B., Henning, Th., Feldt, M., Hayward, T. L, Hoare, M.
G., Hofner, P., & Richter, S. 1998, AJ, 115, 767
Storey, P. J., & Hummer, D. G. 1995, MNRAS
Takami, H., et al. 2004, PASJ, 56, 225
Tokunaga, A. T., et al. 1998, Proc. SPIE, 3354, 512
van Loon, J. Th., & Oliveira, J. M. 2003, A&A, 405, L33
van den Ancker, M. E., Th´e, P. S., Feinstein, A., V´azquez, R. A.,
de Winter, D., & P´erez, M. R. 1997, A&AS, 123, 63
Woodward, C. E., Pipher, J. L., Helfer, H. L., Sharpless, S., Moneti,
A., Kozikowski, D., Oliveri, M., Willner, S. P., Lacasse, M. G.,
& Herter, T. 1986, AJ, 91, 870
Woodward, C. E., Pipher, J. L, Helfer, H. L, & Forrest, W. J. 1990,
ApJ, 365, 252
Woolf, N. J. 1961, PASP, 73, 206
Woolf, N. J., Stein, W. A., Gillett, F. C., Merrill, K. M., Becklin,
E. E., Neugebauer, G., Pepin, T. J., 1973, ApJ, 179, L111
|
astro-ph/0309501 | 1 | 0309 | 2003-09-17T21:58:48 | Tomographic 3D-Modeling of the Solar Corona with FASR | [
"astro-ph"
] | The "Frequency-Agile Solar Radiotelescope" (FASR) litteraly opens up a new dimension in addition to the 3D Euclidian geometry: the frequency dimension. The 3D geometry is degenerated to 2D in all images from astronomical telescopes, but the additional frequency dimension allows us to retrieve the missing third dimension by means of physical modeling. We call this type of 3D reconstruction "Frequency Tomography". In this study we simulate a realistic 3D model of an active region, composed of 500 coronal loops with the 3D geometry [x(s),y(s),z(s)] constrained by magnetic field extrapolations and the physical parameters of the density n_e(s) and temperature T_e(s) given by hydrostatic solutions. We simulate a series of 20 radio images in a frequency range of f=0.1-10 GHz, anticipating the capabilities of FASR, and investigate what physical information can be retrieved from such a dataset. We discuss also forward-modeling of the chromospheric and Quiet Sun density and temperature structure, another primary goal of future FASR science. | astro-ph | astro-ph |
Tomographic 3D-Modeling of the Solar Corona with
FASR
Markus J. Aschwanden, David Alexander, and Marc DeRosa
Lockheed Martin Advanced Technology Center, Solar & Astrophysics Laboratory,
Dept. L9-41, Bldg.252, 3251 Hanover St., Palo Alto, CA 94304, USA; e-mail:
[email protected]
2002/Oct/14
Abstract. The Frequency-Agile Solar Radiotelescope (FASR) litteraly opens up a
new dimension in addition to the 3D Euclidian geometry: the frequency dimension.
The 3D geometry is degenerated to 2D in all images from astronomical telescopes,
but the additional frequency dimension allows us to retrieve the missing third dimen-
sion by means of physical modeling. We call this type of 3D reconstruction Frequency
Tomography. In this study we simulate a realistic 3D model of an active region,
composed of 500 coronal loops with the 3D geometry [x(s), y(s), z(s)] constrained
by magnetic field extrapolations and the physical parameters of the density ne(s)
and temperature Te(s) given by hydrostatic solutions. We simulate a series of 20
radio images in a frequency range of ν = 0.1 − 10 GHz, anticipating the capabilities
of FASR, and investigate what physical information can be retrieved from such
a dataset. We discuss also forward-modeling of the chromospheric and Quiet Sun
density and temperature structure, another primary goal of future FASR science.
Keywords: Sun : corona -- Sun : chromosphere -- Sun : radio
1.
INTRODUCTION
Three-dimensional (3D) modeling of solar phenomena has always been
a challenge with the available two-dimensional (2D) images, but is an
utmost necessity to test physical models in a quantitative way. Since
solar imaging telescopes never have been launched on multiple space-
craft that separate to a significant parallax angle from the Earth, no
true 3D imaging or solar tomography (Davila 1994; Gary, Davis, &
Moore 1998; Liewer et al. 2001) has been performed so far. The Solar
TErrestrial RElations Observatory (STEREO), now being assembled
and planned for launch in November 2005, will be the first true stereo-
scopic facility, mapping the Sun with an increasing separation angle
of 22◦ per year. Alternative approaches of 3D reconstruction methods
utilize the solar rotation to vary the aspect angle (Altschuler 1979;
Berton & Sakurai 1985; Koutchmy & Molodensky 1992; Aschwanden
& Bastian 1994a,b; Batchelor 1994; Hurlburt et al. 1994; Zidowitz 1999;
Koutchmy, Merzlyakov, & Molodensky 2001), but this method gener-
ally requires static structures over several days. An advanced form of
c(cid:13) 2018 Kluwer Academic Publishers. Printed in the Netherlands.
fasr.tex; 24/11/2018; 9:51; p.1
2
ASCHWANDEN, ALEXANDER, & DEROSA
solar rotation stereoscopy is the so-called dynamic stereoscopy method
(Aschwanden et al. 1999, 2000a), where the 3D geometry of dynamic
plasma structures can be reconstructed as long as the guiding magnetic
field is quasi-stationary. Of course, 3D modeling with 2D constraints
can also be attempted if a-priori assumptions are made for the ge-
ometry, e.g. using the assumption of coplanar and semi-circular loops
(Nitta, VanDriel-Gestelyi, & Harra-Murnion 1999).
A new branch of 3D modeling is the combination of 2D images
I(x, y) with the frequency dimension ν, which we call frequency tomog-
raphy. There have been only very few attempts to apply this method to
solar data, mainly because multi-frequency imaging was not available
or had insufficient spatial resolution. There are essentially only three
published studies that employ the method of frequency tomography:
Aschwanden et al. (1995); Bogod & Grebinskij (1997); Grebinskij et al.
(2000.).
In the first study (Aschwanden et al. 1995), gyroresonance emission
above a sunspot was observed at 7 frequencies in both polarizations in
the frequency range of ν = 10 − 14 GHz with the Owens Valley Radio
Observatory (OVRO) during 4 days. From stereoscopic correlations
the height levels h(ν) of each frequency could be determined above the
sunspot. Correcting for the jump in height when dominant gyroreso-
nance emission switches from the second (s = 2) to the third harmonic
(s = 3), the magnetic field B(ν) = 357(νGHz/s) [G] could then be de-
rived as a function of height, B(h), and was found to fit a classical dipole
field B(h) = B0(1 + h/hD)−3. Moreover, from the measured bright-
ness temperature spectrum TB(ν), using the same stereoscopic height
measurement h(ν), also the temperature profile T (h) as a function of
height above the sunspot could be determined. This study represents
an application of frequency tomography, additionally supported with
solar rotation stereoscopy, and thus is subject to the requirement of
quasi-stationary structures.
In the second study (Bogod & Grebinskij 1997), brightness temper-
ature spectra TB(ν) were measured in 36 frequencies in the wavelength
range of λ = 2 − 32 cm (ν = 0.94 − 15 GHz) with RATAN-600, from
quiet-Sun regions, active region plages, and from coronal holes. A dif-
ferential deconvolution method of Laplace transform inversion was then
used to infer the electron temperature T (τ ) as a function of the opacity
τ . This method does not yield the temperature as a function of an
absolute height h, but if an atmospheric model [T (h), ne(h)] is available
as a function of height, the temperature as a function of the free-free
(bremsstrahlung) opacity T (τ ) can be calculated and compared with
the observations.
fasr.tex; 24/11/2018; 9:51; p.2
FASR CORONAL TOMOGRAPHY
3
B
B
(ν) and T LCP
In the third study (Grebinskij et al. 2000), the brightness temper-
ature in both polarizations is measured as a function of frequency,
i.e. T RCP
(ν). Since the magnetic field has a slightly
different refractive index in the two circular polarizations, the free-
free (bremsstrahlung) opacity is consequently also slightly different, so
that the magnetic field B(ν) can be inferred. Again, a physical model
[T (h), ne(h), B(h)] is needed to predict B(ν) and to compare it with
the observed spectrum TB(ν).
The content of this paper is as follows: In Section 2 we simulate an
active region, with the 3D geometry constrained by an observed mag-
netogram and the physical parameters given by hydrostatic solutions,
which are used to calculate FASR radio images in terms of brightness
temperature maps TB(x, y, ν), and test how the physical parameters
of individual coronal loops can be retrieved with FASR tomography.
In Section 3 we discuss a few examples of chromospheric and Quiet-
Sun corona modeling to illustrate the power and limitations of FASR
tomography. In the final Section 5 we summarize some primary goals
of FASR science that can be pursued with frequency tomography.
2. ACTIVE REGION MODELING
2.1. Simulation of FASR Images
Our aim is to build a realistic 3D model of an active region, in form of
3D distributions of the electron density ne(x, y, z) and electron temper-
ature Te(x, y, z), which can be used to simulate radio brightness temper-
ature maps TB(x, y, ν) at arbitrary frequencies ν that can be obtained
with the planned Frequency-Agile Solar Radiotelescope (FASR).
We start from a magnetogram recorded with the Michelson Doppler
Imager (MDI) instrument onboard the Solar and Heliospheric Obser-
vatory (SoHO) on 1999 May 8, 0-1 UT. We perform a potential field
extrapolation with the magnetogram as lower boundary condition of
the photospheric magnetic field, to obtain the 3D geometry of magnetic
field lines. We apply a threshold for the minimum magnetic field at
the footpoints, which limits the number of extrapolated field lines to
n = 500. The projection of these 3D field lines along the line-of-sight
onto the solar disk is shown in Fig.1. We basically see two groups of
field lines, (1) a compact double arcade with low-lying field lines in an
active region in the north-east quadrant of the Sun, and (2) a set of
large-scale field lines that spread out from the eastern active region to
the west and close in the western hemisphere. From this set of field lines
we have constrained the 3D geometry of 500 coronal loops, defined by
a length coordinate s(x, y, z).
fasr.tex; 24/11/2018; 9:51; p.3
4
ASCHWANDEN, ALEXANDER, & DEROSA
MDI 1999-May-8, 0-1 UT
]
i
i
d
a
r
r
a
o
s
[
x
l
e
c
n
a
i
t
s
d
S
N
1.0
0.5
0.0
-0.5
-0.5
0.0
EW distance x[solar radii]
0.5
Figure 1. Potential field extrapolation of SoHO/MDI magnetogram data from 1999
May 8, 0-1 UT.
In a next step we fill the 500 loops with coronal plasma with density
ne(s) and temperature functions Te(s) that obey hydrostatic solutions.
For accurate analytical approximations of hydrostatic solutions we used
the code given in Aschwanden & Schrijver (2002). Each hydrostatic
solution is defined by three independent parameters: the loop length L,
the loop base heating rate EH0, and the heating scale height sH. The
momentum and energy balance equation (between the heating rate and
radiative and conductive loss rate, i.e. EH(s) + Erad(s) + Econd(s) = 0,
yields a unique solution for each parameter set (L, sH, EH0). For the set
of short loops located in the compact double arcade, which have have
lengths of L ≈ 4 −100 Mm, we choose a heating scale height of sH = 10
Mm and base heating rates that are randomly distributed in the loga-
rithmic interval of EH0 = 10−4, ..., 10−2 erg cm−3 s−1. For the group of
fasr.tex; 24/11/2018; 9:51; p.4
FASR CORONAL TOMOGRAPHY
5
log(L)= 9.9+
log(L)= 9.9+ 0.2
log(L)=10.7+
log(L)=10.7+ 0.3
10-7
10-8
10-9
i
n
b
r
e
p
r
e
b
m
u
N
10-10
N=413,234
1010
1011
length L
log(Te)= 6.1+
log(Te)= 6.1+ 0.2
log(Te)= 5.9+
log(Te)= 5.9+ 0.3
109
i
i
n
n
b
b
r
r
e
e
p
p
r
r
e
e
b
b
m
m
u
u
N
N
10-3
10-3
10-4
10-4
10-5
10-5
10-6
10-6
log(ne)=10.8+
log(ne)=10.8_ 0.7
log(ne)= 9.3+
log(ne)= 9.3_ 0.7
108
107
1010
Max. density ne,max [cm-3]
109
log(ne)= 8.8+
log(ne)= 8.8_ 0.6
log(ne)= 4.3+
log(ne)= 4.3_ 2.6
10-6
10-7
10-8
10-9
i
n
b
r
e
p
r
e
b
m
u
N
1012
106
10-4
10-5
10-6
10-7
i
n
b
r
e
p
r
e
b
m
u
N
105
105
106
106
Max. temperature Te,max [K]
Max. temperature Te,max [K]
107
107
106
108
107
Min. density ne,min [cm-3]
109
1010
1011
1011
Figure 2. Distributions of loop lengths L, loop maximum temperatures Te,max, loop
minimum densities nmin, and maximum densities nmax. The distributions with thick
linestyle correspond to ≈ 400 loops in the compact arcade, while the distributions
with thin linestyle correspond to the group of ≈ 100 large-scale loops.
long loops with lengths of L ≈ 100 − 800 Mm, we choose near-uniform
heating (sH = 800 Mm) and volumetric heating rates randomly dis-
tributed in the logarithmic interval of EH0 = 0.5 × 10−7, ..., 0.5 × 10−5
erg cm−3 s−1. This choice of heating rates produces a distribution of
loop maximum temperatures (at the loop tops) of Te ≈ 1 − 3 MK,
electron densities of ne ≈ 108, ..., 1010 cm−3 at the footpoints, and
ne ≈ 106, ..., 109 cm−3 at the loop tops. We show the distribution of
loop top temperatures, loop base densities, and loop top densities in
Fig.2. These parameters are considered to be realistic in the sense that
they reproduce typical loop densities and temperatures observed with
SoHO and TRACE, as well as correspond to the measured heating scale
heights of sH ≈ 10 − 20 Mm (Aschwanden, Nightingale, & Alexander
2000b), for the set of short loops.
For the simulation of radio images we choose an image size of 512 ×
512 pixels, with a pixel size of 2.25", and 21 frequencies logarithmically
distributed between ν = 100 MHz and 10 GHz. To each magnetic
field line we attribute a loop with a width (or column depth) of w ≈
108, ..., 109 cm. For each voxel, i.e. volume element at x = (xi, yj, zk),
we calculate the free-free absorption coefficient κf f (e.g. Lang 1980,
fasr.tex; 24/11/2018; 9:51; p.5
6
ASCHWANDEN, ALEXANDER, & DEROSA
ν= 100 MHz
TB= 0.06 MK
ν= 125 MHz
TB= 0.07 MK
ν= 158 MHz
TB= 0.08 MK
∆x=200.0 arcsec
ν= 199 MHz
TB= 0.10 MK
∆x=158.9 arcsec
ν= 251 MHz
TB= 0.13 MK
∆x=126.2 arcsec
ν= 316 MHz
TB= 0.16 MK
∆x=100.2 arcsec
ν= 398 MHz
TB= 0.24 MK
∆x= 79.6 arcsec
ν= 501 MHz
TB= 0.35 MK
∆x= 63.2 arcsec
ν= 630 MHz
TB= 0.44 MK
∆x= 50.2 arcsec
ν= 794 MHz
TB= 0.46 MK
∆x= 39.9 arcsec
ν=1000 MHz
TB= 0.51 MK
∆x= 31.7 arcsec
ν=1258 MHz
TB= 0.51 MK
∆x= 25.2 arcsec
∆x= 20.0 arcsec
∆x= 15.9 arcsec
Figure 3. Simulation of radio brightness temperature maps of an active region at 20
frequencies, from ν=100 to 1258 MHz. The maximum brightness temperature (TB)
and the angular resolution ∆x are indicated in each frame.
fasr.tex; 24/11/2018; 9:51; p.6
FASR CORONAL TOMOGRAPHY
7
ν= 0.79 GHz
TB= 0.46 MK
ν= 1.00 GHz
TB= 0.51 MK
ν= 1.26 GHz
TB= 0.51 MK
∆x= 25.2 arcsec
ν= 1.58 GHz
TB= 0.48 MK
∆x= 20.0 arcsec
ν= 2.00 GHz
TB= 0.57 MK
∆x= 15.9 arcsec
ν= 2.51 GHz
TB= 0.47 MK
∆x= 12.6 arcsec
ν= 3.16 GHz
TB= 0.85 MK
∆x= 10.0 arcsec
ν= 3.98 GHz
TB= 1.00 MK
∆x= 8.0 arcsec
ν= 5.01 GHz
TB= 1.06 MK
∆x= 6.3 arcsec
ν= 6.31 GHz
TB= 1.84 MK
∆x= 5.0 arcsec
ν= 7.94 GHz
TB= 1.85 MK
∆x= 4.0 arcsec
ν=10.00 GHz
TB= 1.60 MK
∆x= 3.2 arcsec
∆x= 2.5 arcsec
∆x= 2.0 arcsec
Figure 4. Similar representation as in Fig.3, for frequencies of ν=0.8 to 10 GHz,
with a smaller field-of-view than in Fig.3. The brightness is shown on linear scale in
the first two rows, and on logarithmic scale in the last two rows (with a contrast of
1:100 in the third row and 1:1000 in forth row).
fasr.tex; 24/11/2018; 9:51; p.7
8
p.47),
ASCHWANDEN, ALEXANDER, & DEROSA
f f (xi, yj, zk) = 9.78 × 10−3 n2
κν
e,ijk
ν2T 3/2
e,ijk
[24.2 + ln (Te,ijk) − ln (ν)] ,
(1)
and integrate the opacity τ ν
f f along the line-of-sight z,
τ ν
f f (xi, yj, zk) = Z z
−∞
κν
f f (xi, yj, zk)dz′ ,
(2)
to obtain the radio brightness temperature T ν
transfer equation (in the Rayleigh-Jeans limit),
B(xi, yj) with the radiative
T ν
B(xi, yj) = Z +∞
−∞
Te,ijk exp−τ ν
f f ((xi,yj,zk) κν
f f (xi, yj, zk)dz ,
(3)
The simulated images for the frequency range of ν=100 MHz to 10 GHz
are shown in Figs.3 and 4. The approximate instrumental resolution
is rendered by smoothing the simulated images with a boxcar that
corresponds to the instrumental resolution of FASR,
wres =
20"
νGHz
.
(4)
A caveat needs to be made, that the real reconstructed radio im-
ages may reach this theoretical resolution only if a sufficient number
of Fourier components are available, either from a large number of
baselines (which scale with the square of the number of dishes) or from
aperture synthesis (which increases the number of Fourier components
during Earth rotation proportionally to the accumulation time inter-
val). Also, we did not include here the effects of angular scattering due
to turbulence or other coronal inhomogeneities (Bastian 1994, 1995).
2.2. Peak Brightness Temperature
The intensity of radio maps is usually specified in terms of the observed
brightness temperature TB. We list the peak brightness temperature
in each map in Figs.3 and 4. We see that a maximum brightness is
observed in the second-last map in Fig.4, with TB = 1.85 MK at
a frequency of ν = 7.94 GHz. Let us obtain some understanding of
the relative brightness temperatures TB(ν) as function of frequency
ν, in order to faciliate the interpretation of radio maps. We plot the
peak brightness temperature TB(ν) of the simulated maps as function
of frequency in Fig.5 (cross symbols). There are two counter-acting
effects that reduce the brightness temperature: First, the loops become
fasr.tex; 24/11/2018; 9:51; p.8
FASR CORONAL TOMOGRAPHY
9
2.5
2.0
1.5
1.0
0.5
]
K
M
[
B
T
e
r
u
t
a
r
e
p
m
e
t
s
s
e
n
t
h
g
i
r
B
0.0
0.1
B
H
C
1.0
Frequency ν [GHz]
10.0
Figure 5. The radio peak brightness temperature TB is shown as function of fre-
quency ν: for the background corona (B), for cool (C) fat loops (T=0.5 MK,
ne = 1011 cm−3, w=25 Mm), and hot (H) thin loops (T=2.0 MK, ne = 1011 cm−3,
w=2.5 Mm). The cross symbols indicate the peak brightness temperatures observed
in the simulated maps (Figs.3 and 4), while the medium-thick line represents the
combined model of hot and cool loops. The dashed line indicates the expected
brightness temperature of hot loops if no beam dilution due to the instrumental
angular resolution would occur. The thick grey curve (B) indicates a model of the
background corona.
optically thin at high frequencies due to the ν −2-dependence of the
free-free opacity (Eq.1). Hot loops with a temperature of T=2.0 MK,
a density of ne = 1011 cm−3, and a width of w=2.5 Mm are optically
thick below ν <
∼ 5 GHz, so the brightness temperature would match the
electron temperature TB = Te (dashed line in Fig.5), but falls off at
higher frequencies, i.e. TB(ν > 5 GHz)< Te.
The second effect that reduces the brightness temperature is the
beam dilution, which has a ν2-dependence below the critical frequency
where structures are unresolved. The effectively observed brightness
temperature T ef f
B (ν) due to beam dilution for a structure with width
w is
B (ν) = TB ×( (cid:16) ν
νcrit(cid:17)2
T ef f
1
for ν < νcrit(w)
for ν > νcrit(w)
(5)
fasr.tex; 24/11/2018; 9:51; p.9
10
ASCHWANDEN, ALEXANDER, & DEROSA
where the critical frequency νcrit(w) depends on the width w of the
structure and is for FASR according to Eq.(4),
νcrit(w) =
20"
w"
[GHz] .
(6)
Because the brightness drops drastically below νcrit ≈ 5 GHz in Fig.5,
we conclude that the width of the unresolved structures is about w" =
20"/5 = 4" = 3 Mm. Therefore we can understand the peak brightness
temperatures in the maps, as shown in Fig.5 (crosses) in the range of
ν ≈ 3 − 10 GHz with a combination of these two effects of free-free
opacity and beam dilution.
Below a frequency of ν <
∼ 3 GHz, we see that another group of loops
contributes to the peak brightness of the maps. We find that the peak
brightness below 3 GHz can adequately be understood by a group of
cooler loops with a temperature of T = 0.5 MK, densities of ne = 1011
cm−3, and widths of w = 25 Mm (Fig.5). Thus cool loops dominate
the brightness at low frequencies, and hot loops at higher frequencies.
In the simulations in Figs.3 and 4 we have not included the back-
ground corona. In order to give a comparison of the effect of the
background corona we calculate the opacity for a space-filling corona
with an average temperature of T = 1.0 MK, an average density of
ne = 109 cm−3, and a vertical (isothermal) scale height of w ≈ λT ≈ 50
Mm. The brightness temperature of this background corona is shown
with a thick grey curve (labeled B) in Fig.5. According to this estimate,
the background corona overwhelms the brightest active region loops
at frequencies of ν <
∼ 1 GHz. From this we conclude that it might be
difficult to observe active region loops at decimetric frequencies ν <
∼ 1.0
GHz, unless they are very high and stick out above a density scale
height, i.e. at altitudes of h >
∼ 50 Mm. In conclusion, the contrast of
active region loops in our example seems to be best at frequencies of
ν ≈ 5 GHz, but drops at both sides of this optimum frequency (see
Fig.5).
2.3. Temperature and Density Diagnostic of Loops
FASR will provide simultaneous sets of images I(x, y, ν) at many fre-
quencies ν. In other words, for every image position (xi, yj), a spectrum
T ij
B (ν) can be obtained. A desirable capability is temperature and
density diagnostic of active region loops. Let us parameterize the pro-
jected position of a loop by a length coordinate sk, k = 1, ..., n, e.g.
[xi = x(sk), yj = y(sk)]. If we manage to determine the temperature
Te(xi, yj) and density ne(xi, yi) at every loop position (xi, yj), we have
a diagnostic of the temperature profile T (s) and density profile ne(s) of
fasr.tex; 24/11/2018; 9:51; p.10
FASR CORONAL TOMOGRAPHY
11
2.0
1.5
]
K
M
[
B
T
1.0
0.5
0.0
0.1
2.0
1.5
]
K
M
[
B
T
1.0
0.5
0.0
0.1
2.0
1.5
]
K
M
[
B
T
1.0
0.5
0.0
0.1
Temperature increase
Te=0.1,...,2.0 MK
ne=1.0 * 1010 cm-3
w=10 Mm
Temperature decrease
1.0
10.0
Te=1.0 MK
ne=0.1,...,2.0 * 1010 cm-3
w=10 Mm
Density increase
Density decrease
1.0
10.0
Te=1.0 MK
ne=1.0 * 1010 cm-3
w=1,...,20 Mm
Width increase
Width decrease
1.0
Frequency ν [GHz]
10.0
Figure 6. The variation of the radio brightness temperature spectrum TB(ν) of a
loop by varying the temperature Te (top panel), the electron density ne (middle
panel), and the loop width w (bottom panel). In each of the three panels, one
parameter is varied from 10%, 20%, ..., 90% (dashed curves) to 110%, 120%,... ,
200% (solid lines). The reference curve with parameters Te = 1.0 MK, ne = 1010
cm−3, and w = 10 Mm is indicated with a thick line. The arrows indicate the
spectral shift of the peak.
fasr.tex; 24/11/2018; 9:51; p.11
12
ASCHWANDEN, ALEXANDER, & DEROSA
an active region loop. Thus, the question is whether we can manage to
extract a temperature Te and density ne from a brightness temperature
spectrum TB(ν) at a given pixel position (i, j). In order to illustrate the
feasibility of this task, we show the brightness temperature spectrum
TB(ν) of a typical active region loop in Fig.6, and display its variation
as a function of the physical (Te, ne) and geometric (w) parameters.
We define a typical active region loop by an electron temperature
Te = 1.0 MK, an electron density ne = 1010 cm−3, and a width w = 10
Mm. Such a loop is brightest at frequencies of ν ≈ 1.5 −3.0 GHz (Fig.6;
thick curve). The loop is fainter at higher frequencies because free-free
emission becomes optically thin, while it is optically thick at lower
frequencies. The reason why the loop is also fainter at low frequencies
is because of the beam dilution at frequencies where the instrument
does not resolve the loop diameter. If we increase the temperature,
the brightness temperature increases, and vice versa decreases at lower
electron temperatures (Fig.6 top). If we increase the density, the crit-
ical frequency where the loop becomes optically thin shifts to higher
frequencies, while the peak brightness temperature decreases for lower
densities (Fig.6, middle panel). If we increase the width of the loops, the
brightness temperature spectrum is bright in a much larger frequency
range, because we shift the critical frequency for beam dilution towards
lower frequencies, while the overall brightness temperature decreases
for a smaller loop width (Fig.6 bottom). Based on this little tutorial,
one can essentially understand how the optimization works in spectral
fitting (e.g. with a forward-fitting technique) to an observed brightness
temperature spectrum TB(ν).
To demonstrate how the density and temperature diagnostic works
in practice, we pick a bright loop seen at ν = 5.0 GHz in Fig.4, which
we show as an enlarged detail in Fig.7 (left panel). We pick three lo-
cations (A,B,C) along the loop and extract the brightness temperature
spectra TB(ν) from the simulated datacube TB(x, y, ν) at the locations
(A,B,C), shown in Fig.7 (three middle panels). Each spectrum shows
two peaks, which we interpret as two cospatial loops. For each spectral
peak we can therefore roughly fit a loop model, constrained by three
parameters each, i.e. [Te, ne, w]. We can now fit a brightness temper-
ature spectrum T ef f
B (ν) to the observed (or simulated here) spectrum
T obs
B (ν), physically defined by the same radiation transfer model for
free-free emission as in Eqs.(1-5), but simplified by the approximation of
constant parameters (Te, ne), and thus a constant absorption coefficient
κf f (ν), over the relatively small spatial extent of a loop diameter w,
κf f (ν) = 9.78 × 10−3 n2
ν2T 3/2
e
e
[24.2 + ln (Te) − ln (ν)] ,
(7)
fasr.tex; 24/11/2018; 9:51; p.12
FASR CORONAL TOMOGRAPHY
]
K
M
13
4
3
2
]
K
M
[
B
T
ν= 5.01 MHz
TB= 1.06 MK
B
A
C
]
K
M
[
B
T
A
B
C
1.0
0.8
0.6
0.4
0.2
0.0
0.1
1.0
0.8
0.6
0.4
0.2
0.0
0.1
1.0
0.8
0.6
0.4
0.2
0.0
1.0
10.0
]
3
-
m
c
[
e
n
y
t
i
s
n
e
D
1.0
10.0
[
e
T
p
m
e
T
1
0
0.0
1012
1011
1010
109
108
20
0.0
]
m
M
[
w
h
t
i
d
W
15
10
5
0
0.0
0.5
1.0
1.5
2.0
0.5
1.0
1.5
2.0
0.5
1.5
Loop length s/L
1.0
2.0
0.1
1.0
Frequency ν [MHz]
10.0
Figure 7. Enlarged detail of the active region with a bright loop (left panel). From
the measured brightness temperature spectra TB(ν) (crosses in middle panels) at
the three loop locations (A,B,C) we fit theoretical spectra and determine the tem-
peratures (right top panel), densities (right middle panel), and loop widths (right
bottom panel) at the three loop locations (A,B,C).
τf f (ν) = κf f (ν) w ,
TB(ν) = Te(cid:16)1 − exp−τf f (ν)(cid:17) ,
T ef f
B (ν) = TB ×( (cid:16) ν
νcrit(cid:17)2
1
for ν < νcrit(w)
for ν > νcrit(w)
(8)
(9)
(10)
What can immediately be determined from the observed brightness
temperature spectra T obs
B (ν) are the frequencies of the spectral peaks
(Fig.7, middle panels), which are found around νpeak = 1.2 and 6.0
GHz. Based on the tutorial given in Fig.6 it is clear that these spec-
tral peaks demarcate the critical frequencies where structures become
unresolved. Thus we can immediately determine the diameters of the
two loops with Eq.(6), i.e. w1 = 20"/1.2 = 17" = 12.0 Mm and
w2 = 20"/6.0 = 3.3" = 2.4 Mm. The only thing left to do is to vary the
temperature and density and to fit the model (Eqs.7-10) to the observed
spectrum. For an approximate solution (shown as smooth curves in
fasr.tex; 24/11/2018; 9:51; p.13
14
ASCHWANDEN, ALEXANDER, & DEROSA
the middle panels of Fig.7) we find T1 = 3.0 MK and n1 = 4 × 1010
cm−3 for the first loop (with width w1 = 2.5 Mm and spectral peak
at ν1 = 6.0 GHz), and T2 = 2.9 MK and n2 = 1.9 × 109 cm−3 for the
second loop (with width w2 = 12 Mm and spectral peak at ν1 = 1.2
GHz). The resulting temperature Te(s) and density profiles ne(s) along
the loops are shown in Fig.7 (right panels). This approximate fit is
just an example to illustrate the concept of forward-fitting to FASR
tomographic data. More information can be extracted from the data by
detailed fits with variable loop cross-section along the loop and proper
deconvolution of the projected column depth across the loop diameter
(which is a function of the aspect angle between the line-of-sight and
the loop axis). For a proper determination of the inclination angle of
the loop plane, the principle of dynamic stereoscopy can be applied
(Aschwanden et al. 1999; see Appendix A therein for coordinate trans-
formations between the observers reference frame and the loop plane).
Of course, our example is somehow idealized, in practice there will be
confusion by adjacent or intersecting loops, as well as confusion by other
radiation mechanisms, such as gyroresonance emission that competes
with free-free emission at frequencies of ν >
∼ 5 GHz near sunspots.
2.4. Radio versus EUV and soft X-ray diagnostics
We can ask whether temperature and density diagnostic of coronal
loops is better done in other wavelengths, such as in EUV and soft
X-rays (e.g. Aschwanden et al. 1999), rather than with radio tomog-
raphy. Free-free emission in EUV and soft X-rays is optically thin,
which has the advantage that every loop along a line-of-sight is visible
to some extent, while loops in optically thick plasmas can be hidden
at radio wavelengths. On the other side, the line-of-sight confusion in
optically thin plasmas is larger in EUV and soft X-rays, in particular if
multiple loops along the same line-of-sight have similar temperatures.
Different loops along a line-of-sight can only be discriminated in EUV
and soft X-rays if they have significantly different temperatures, so
that they show different responses in lines with different ionization
temperatures. Two cospatial loops that have similar temperatures but
different widths cannot be distinguished by EUV or soft X-ray detec-
tors. In radio wavelengths, however, even cospatial loops with similar
temperatures, as the two loops in our example in Fig.7 (T1 = 3.0 MK
and T2 = 2.9 MK), can be separated if they have different widths. The
reason is that they have different critical frequencies νcrit(w) where
they become resolved, and thus show up as two different peaks in
the brightness temperature spectrum T ef f
B (ν). Radio tomography has
therefore a number of unique advantages over loop analysis in EUV
fasr.tex; 24/11/2018; 9:51; p.14
FASR CORONAL TOMOGRAPHY
15
and soft X-ray wavelengths: (1) a ground-based instrument is much
less costly than a space-based instrument, (2) a wide spectral radio
wavelength range (decimetric, centimetric) provides straightforwardly
diagnostic over a wide temperature range, while an equivalent tempera-
ture diagnostic in EUV and soft X-rays would require a large number of
spectral lines and instrumental filters, (3) optically thick radio emission
is most sensitive to cool plasma, which is undetectable in EUV and soft
X-rays, except for absorption in the case of very dense cool plasmas,
and (4) radio brightness temperature spectra can discriminate multiple
cospatial structures with identical temperatures based on their spatial
scale, which is not possible with optically thin EUV and soft X-ray
emission.
3. CHROMOSPHERIC AND CORONAL MODELING
The vertical density and temperature structure of the chromosphere,
transition region, and corona has been probed in soft X-rays, EUV,
and in radio wavelengths, but detailed models that are consistent in
all wavelengths are still unavailable. Comprehensive coverage of the
multi-thermal and inhomogeneous solar corona necessarily requires ei-
ther many wavelength filters in soft X-rays and EUV, or many radio
frequencies, for which FASR will be the optimum instrument.
We illustrate the concept of how to explore the vertical structure
of the chromosphere and corona with a few simple examples. We know
that the corona is highly inhomogeneous along any line-of-sight, so a 3D
model has to be composed of a distribution of many magnetic fluxtubes,
each one representing a mini-atmosphere with its own density and tem-
perature structure, being isolated from each other due to the low value
of the plasma-beta, i.e. β = pthermal/pmagn = 2nekBTe/(B2/8π) ≪ 1.
The confusion due to inhomogeneous temperatures and densities is
largest for line-of-sights above the limb (due to the longest column
depths with contributing opacity), and is smallest for line-of-sights near
the solar disk center, where we look down through the atmosphere in
vertical direction.
The simplest model of the atmosphere is given by the hydrostatic
equilibrium in the isothermal approximation, T (h) = const, where the
hydrostatic scale height λT is proportional to the electron temperature
T , i.e.
λT =
(11)
kBT
µmpg⊙
= λ0(cid:16) T
1 MK(cid:17)
with λ0 = 47 Mm for coronal conditions, with µmp the average ion
mass (i.e. µ ≈ 1.3 for H:He=10:1) and g⊙ the solar gravitation. The
fasr.tex; 24/11/2018; 9:51; p.15
16
ASCHWANDEN, ALEXANDER, & DEROSA
height dependence of the electron density is for gravitational pressure
balance,
ne(h) = n0 exp[−
(h − h0)
λ0T
] .
(12)
where n0 = n(h0) is the base electron density. This expression for
the density ne(h) can then be inserted into the free-free absorption
coefficient κ(h, ν), with T (h) = const in the isothermal approximation,
κf f (h, ν) = 9.78 × 10−3 n2
e(h)
ν2T (h)3/2 [24.2 + ln T (h) − ln (ν)] ,
(13)
At disk center, we can set the altitude h equal to the line-of-sight
coordinate z, so that the free-free opacity τf f (h, ν) integrated along
the line-of-sight h = z is,
τf f (h, ν) = Z h
−∞
κf f (h′, ν) dh′ ,
(14)
and the radio brightness temperature TB(ν) is then
TB(ν) = Z 0
−∞
T (h) exp−τf f (h,ν) κf f (h, ν)dh .
(15)
With this simple model we can determine the mean temperature T (h)
by fitting the observed brightness temperature spectra TB(ν) to the
theoretical spectra (Eq.15) by varying the temperature T (h) = const
(in Eqs.13-14). The expected brightness temperature spectra for an
isothermal corona with temperatures of T = 1.0 MK and T = 5.0 MK
and a base density of n0 = 109 cm−3 are shown in Fig.8. We see that
the corona becomes optically thin (TB ≪ Te) at frequencies of ν >
∼ 1 − 2
GHz in this temperature range that is typical for the Quiet Sun.
These hydrostatic models in the lower corona, however, have been
criticized because of the presence of dynamic phenomena, such as spicu-
lae, which may contribute to an extended chromosphere in the statisti-
cal average. The spicular extension of this dynamic chromosphere has
been probed with high-resolution measurements of the Normal Inci-
dence X-Ray Telescope (NIXT) (Daw, DeLuca, & Golub 1995) as well
as with radio submillimeter observations during a total eclipse (Ewell et
al. 1993). Using the radio limb height measurements at various mm and
sub-mm wavelengths in the range of 200-3000 µm (Roellig et al. 1991;
Horne et al. 1981; Wannier et al. 1983; Belkora et al. 1992; Ewell et al.
1993), an empirical Caltech Irreference Chromospheric Model (CICM)
was established, which fits the observed limb heights between 500 km
and 5000 km in a temperature regime of T = 4410 K to T = 7500
K (Ewell et al. 1993), shown in Fig.9. We see that these radio limb
fasr.tex; 24/11/2018; 9:51; p.16
10.000
1.000
0.100
0.010
]
K
M
[
B
T
.
p
m
e
t
s
s
e
n
t
h
g
i
r
B
0.001
0.1
FASR CORONAL TOMOGRAPHY
17
1.0
Frequency ν[GHz]
10.0
100.0
Figure 8. Quiet Sun brightness temperature spectrum for an isothermal corona with
T = 1.0 MK (solid line) or T = 5.0 MK (dashed line) with a base density of n0 = 109
cm−3.
measurements yield electron densities that are 1-2 orders of magnitude
higher in the height range of 500-5000 km than predicted by hydrostatic
models (VAL, FAL, Gabriel 1976), which was interpreted in terms of
the dynamic nature of spiculae (Ewell et al. 1993). This enhanced
density in the extended chromosphere has also been corroborated with
recent RHESSI measurements (Fig.8; Aschwanden, Brown, & Kontar
2002). Hard X-rays mainly probe the total neutral and ionized hydrogen
density that governs the bremsstrahlung and the total bound and free
electron density in collisional energy losses, while the electron density
ne(h) inferred from the radio-based measurements is based on free-
free emission, and shows a remarkably good agreement in the height
range of h ≈ 1000 − 3000 km. The extended chromosphere produces
substantially more opacity at microwave frequencies than hydrostatic
models (e.g. Gabriel 1976).
The atmospheric structure thus needs to be explored with more
general parameterizations of the density ne(h) and temperature Te(h)
structure than hydrostatic models provide. For instance, each of the
atmospheric models shown in Fig.9 provides different functions ne(h)
and Te(h). Observational tests of these models can simply be made
by forward-fitting of the parameterized height-dependent density ne(h)
and temperature profiles Te(h), using the expressions for free-free emis-
sion (Eqs.13-15). In Fig.10 we illustrate this with an example. The
datapoints (shown as diamonds in Fig.10) represent radio observations
of the solar limb at frequencies of ν = 1.4 − 18 GHz during the solar
minimum in 1986-87 by Zirin, Baumert, & Hurford (1991). We show
in Fig.10 an isothermal hydrostatic model for a coronal temperature
of Te = 1.5 MK and a base density of ne = 109 cm−3, as well as
the hydrostatic model of Gabriel (1976), of which the density profile
fasr.tex; 24/11/2018; 9:51; p.17
18
1018
ASCHWANDEN, ALEXANDER, & DEROSA
]
3
-
m
c
[
e
n
y
t
i
s
n
e
d
n
o
r
t
c
e
e
l
d
n
a
0
H
n
n
e
g
o
r
d
y
h
l
a
r
t
u
e
N
Neutral hydrogen density
ME
MM
D
FAL-C
VAL-C
O
FAL-P
Gu
1016
1014
Gu
VAL-C
FAL-C
D
FAL-P
1012
Electron density
MM
ME
O
1010
108
100
50 keV
45 keV
40 keV
35 keV
30 keV
25 keV
20 keV
15 keV
RHESSI
Flare loop
CICM
Radio limb
Gabriel
Corona
10000
1000
Altitude h[km]
Figure 9. A compilation of chromospheric and coronal density models: VAL-C =
Vernazza, Avrett, & Loeser (1981), model C; FAL-C = Fontenla, Avrett, & Loeser
(1990), model C; FAL-P = Fontenla, Avrett, & Loeser (1990), model P; G = Gu,
Jefferies et al. (1997); MM = Maltby et al., (1986), model M; ME = Maltby et
al., (1986), model E; D = Ding & Fang (1989); O = Obridko & Staude (1988);
Gabriel = Gabriel (1976), coronal model; CICM = Caltech Irreference Chromo-
spheric Model, radio sub-millimeter limb observations (Ewell et al. 1993), RHESSI
flare loop (Aschwanden, Brown, & Kontar 2002).
ne(h) is shown in Fig.9. The Gabriel model was calculated based on the
expansion of the magnetic field of coronal flux tubes over the area of a
supergranule (canopy geometry). The geometric expansion factor and
the densities at the lower boundary in the transition region (given by
the chromospheric VAL and FAL models, see Fig.9) then constrains the
coronal density model ne(h), which falls off exponentially with height
in an isothermal fluxtube in hydrostatic equilibrium. We see that the
Gabriel model roughly matches the isothermal hydrostatic model (see
Fig.10), but does not exactly match the observations by Zirin et al.
(1991). However, if we multiply the Gabriel model by a factor of 0.4, to
adjust for solar cycle minimum conditions, and add a temperature of
fasr.tex; 24/11/2018; 9:51; p.18
FASR CORONAL TOMOGRAPHY
19
10.00
1.00
Isothermal
0.10
Gabriel*0.4+Tch
]
K
M
[
B
T
.
p
m
e
t
s
s
e
n
t
h
g
i
r
B
Gabriel
Zirin
0.01
0.1
1.0
Frequency ν[GHz]
10.0
Figure 10. Quiet Sun brightness temperature spectrum for an isothermal corona
with T = 1.5 MK with a base density of n0 = 109 cm−3 (solid thin line), for the
coronal model by Gabriel (1976) (thin solid line), and for a modified Gabriel model
(thick solid line).
Te = 11, 000 K to account for an optically thick chromosphere (similar
to the values determined by Bastian, Dulk, and Leblanc 1996), we find a
reasonably good fit to the observations of Zirin (thick curve in Fig.10).
This example demonstrates that radio spectra in the frequency range
of ν ≈ 1 − 10 GHz are quite sensitive to probe the physical structure
of the chromosphere and transition region.
4. FUTURE FASR SCIENCE
With our study we illustrated some basic applications of frequency
tomography as can be expected from FASR data. We demonstrated
how physical parameters from coronal loops in active regions, from the
Quiet-Sun corona, and from the chromosphere and transition region can
be retrieved. Based on these capabilities we expect that the following
science goals can be efficiently studied with future FASR data:
1. The electron density ne(s) and electron temperature profile Te(s)
of individual active region loops can be retrieved, which constrain
the heating function EH (s) along the loop in the momentum and
energy balance hydrodynamic equations. This enables us to test
whether a loop is in hydrostatic equilibrium or evolves in a dy-
namic manner. Detailed dynamic studies of the time-dependent
heating function EH (s, t) may reveal the time scales of intermittent
plasma heating processes, which can be used to constrain whether
AC or DC heating processes control energy dissipation. Ultimately,
such quantitative studies will lead to the determination and iden-
tification of the so far unknown physical heating mechanisms, a
fasr.tex; 24/11/2018; 9:51; p.19
20
ASCHWANDEN, ALEXANDER, & DEROSA
long-thought goal of the so-called coronal heating problem. Radio
diagnostic is most sensitive to cool dense plasma, but is also sen-
sitive continuously up to the highest temperatures, and this way
nicely complements EUV and soft X-ray diagnostic.
2. Because coronal loops are direct tracers of closed coronal mag-
netic field lines, the reconstruction of the 3D geometry of loops,
as it can be mapped out with multi-frequency data from FASR
in a tomographic manner, this information can be used to test
theoretical models based on magnetic field extrapolations from the
photosphere. The circular polarization of free-free emission contains
additional information on the magnetic field (Grebinskij et al. 2000;
Gelfreikh 2002; Brosius 2002), as well as gyroresonance emission
provides direct measurements of the magnetic field by its propor-
tionality to the gyrofrequency (Lee et al. 1998; White 2002; Ryabov
2002). Ultimately, such studies may constrain the non-potentiality
and the localization of currents in the corona.
3. The density ne(h) and temperature profile Te(h) of the chromo-
sphere, transition region, and corona can be determined in the
Quiet Sun from brightness temperature spectra TB(ν), with least
confusion at disk center. Parameterized models of the density and
temperature structure, additionally constrained by the hydrody-
namic equations and differential emission measure distributions,
can be forward-fitted to the observed radio brightness temperature
spectra TB(ν). This provides a new tool to probe physical condi-
tions in the transition region, deviations from hydrostatic equilib-
ria, and diagnostic of dynamic processes (flows, turbulence, waves,
heating, cooling) in this little understood interface to the corona.
4. Since free-free emission is most sensitive to cool dense plasma,
FASR data will also be very suitable to study the origin, evolution,
destabilization, and eruption of filaments, which seem to play a cru-
cial role in triggering and onset of coronal mass ejections (Vourlidas
2002). Ultimately, the information to forecast CMEs may be chiefly
exploited from the early evolution of filaments.
Previous studies with multi-frequency instruments (VLA, OVRO, Nan-
¸cay, RATAN-600) allowed only crude attempts to pioneer tomographic
3D-modeling of the solar corona, because of the limitations of a rel-
atively small number of Fourier components and a sparse number of
frequencies. FASR will be the optimum instrument to faciliate 3D di-
agnostic of the solar corona on a routine basis, which is likely to lead to
groundbreaking discoveries in long-standing problems of coronal plasma
physics.
fasr.tex; 24/11/2018; 9:51; p.20
FASR CORONAL TOMOGRAPHY
21
References
Altschuler,M.D. 1979, in Image Reconstruction from Projections, (ed. G.T.Herman,
Berlin:Springer, p.105
Aschwanden,M.J. and Bastian,T.S. 1994a, Astrophys.J. 426, 425
Aschwanden,M.J. and Bastian,T.S. 1994b, Astrophys.J. 426, 434
Aschwanden,M.J., Lim,J., Gary,D.E., and Klimchuk,J.A. 1995, Astrophys.J. 454,
512
Aschwanden,M.J., Newmark,J.S., Delaboudiniere,J.P., Neupert,W.M., Klim-
chuk,J.A., Gary,G.A., Portier-Fornazzi,F., and Zucker,A. 1999, Astrophys.J.
515, 842
Aschwanden,M.J., Alexander,D., Hurlburt,N., Newmark,J.S., Neupert,W.M., Klim-
chuk,J.A., and G.A.Gary 2000a, Astrophys.J. 531, 1129
Aschwanden,M.J., Nightingale,R.W., and Alexander,D. 2000b, Astrophys.J. 541,
1059
Aschwanden,M.J. and Schrijver,K.J. 2002, Astrophys.J.Suppl.Ser. 142, 269
Aschwanden,M.J., Brown,J.C., and Kontar,E.P. 2002, Solar Phys. (in press)
Bastian,T.S. 1994, Astrophys.J. 426, 774
Bastian,T.S. 1995, Astrophys.J. 439, 494
Bastian,T.S., Dulk,G.A., and Leblanc,Y. 1996, Astrophys.J. 473, 539
Batchelor,D.A. 1994, Solar Phys. 155, 57
Belkora,L., Hurford,G.J., Gary,D.E. and Woody,D.P. 1992, Astrophys.J. 400, 692
Berton.R. and Sakurai,T. 1985, Solar Phys. 96, 93
Bogod,V.M., and Grebinskij,A.S. 1997, Solar Phys. 176, 67
Brosius,J.W. 2002, (in this volume)
Davila,J.M. 1994, Astrophys.J. 423, 871
Daw,A., DeLuca,E.E., and Golub,L. 1995, Astrophys.J. 453, 929
Ding, M.D. and Fang, C. 1989, Astron.Astrophys. 225, 204.
Ewell, M.W.Jr., Zirin, H., Jensen, J.B., and Bastian, T.S. 1993, Astrophys.J. 403,
426.
Fontenla, J.M., Avrett, E.H., and Loeser, R. 1990, Astrophys.J. 355, 700.
Gabriel, A.H. 1976, Royal Society (London), Philosophical Transactions, Series A,
281, no. 1304, p. 339.
Gary,A., Davis,J.M., and Moore,R. 1998, Solar Phys. 183, 45
Gelfreikh,G.B. 2002, (in this volume)
Grebinskij,A., Bogod,V., Gelfreikh,G., Urpo,S., Pohjolainen,S. and Shibasaki,K.
2000, Astron.Astrophys.Suppl.Ser. 144, 169
Gu, Y., Jefferies, J.T., Lindsey, C. and Avrett, E.H. 1997, Astrophys.J. 484, 960.
Horne, K., Hurford, G.J., Zirin, H., and DeGraauw, Th. 1981, Astrophys.J. 244,
340.
Hurlburt,N.E., Martens,P,C.H., Slater,G.L., and Jaffey,S.M. 1994, in Solar Active
Region Evolution: Comparing Models with Observations, ASP Conf. Ser. 68, 30
Lang,K.R. 1980, Astrophysical Formulae. A Compendium for the Physicist and
Astrophysicist, Berlin: Springer
Lee,J.W., McClymont,A.N., Mikic,Z., White,S.M., and Kundu,M.R. 1998, Astro-
phys.J. 501, 853
Koutchmy,S., and Molodensky,M.M. 1992, Nature 360, 717
Koutchmy,S., Merzlyakov, V. L., and Molodensky, M. M. 2001, Astronomy Reports
45/10, 834
Liewer,P.C., Hall,J.R., DeJong,M., Socker,D.G., Howard,R.A., Crane,P.C.,
Reiser,P., Rich,N., Vourlidas,A. 2001, J.Geophys.Res. 106/A8, 15903
fasr.tex; 24/11/2018; 9:51; p.21
22
ASCHWANDEN, ALEXANDER, & DEROSA
Maltby, P., Avrett, E.H., Carlsson, M., Kjeldseth-Moe, O., Kurucz, R.L., and Loeser,
R. 1986, Astrophys.J. 306, 284.
Nitta,N., VanDriel-Gestelyi,L., and Harra-Murnion,L,K. 1999, Solar Phys. 189, 181
Obridko, V.N. and Staude, J. 1988, Astron.Astrophys. 189, 232.
Roellig, T.L., Becklin, E.E., Jefferies, J.T., Kopp, G.A., Lindsey, C.A., Orral, F.Q.,
and Werner, M.W., 1991, Astrophys.J. 381, 288.
Ryabov,V.B. 2002, (in this volume)
Vernazza, J.E., Avrett, E.H., and Loeser, R. 1981, Astrophys.J.Suppl.Ser. 45, 635.
Vourlidas,A. 2002, (in this volume)
Wannier, P.G., Hurford, G.J., and Seielstad, G.A. 1983, Astrophys.J. 264, 660.
White,S.M. 2002, (in this volume)
Zidowitz,S. 1999, J.Geophys.Res. 104/A5, 9727
Zirin,H., Baumert,B.M., and Hurford,G.J. 1991, Astrophys.J. 370, 779
fasr.tex; 24/11/2018; 9:51; p.22
|
astro-ph/0108354 | 1 | 0108 | 2001-08-22T14:12:25 | The stability of the Circumnuclear Disk clouds in the Galactic Centre | [
"astro-ph"
] | The influence of rotation and magnetic fields on the physical properties of isothermal gas clouds is discussed. The presence of rotation and/or magnetic fields results in an increase of the critical cloud mass with respect to gravitational instability for clouds of a given temperature and external pressure. Rotating clouds have higher densities. Consequently, they are more stable against tidal shear than non-rotating clouds. They can approach the Galactic Centre up to a radius of ~2 pc without being disrupted by the tidal shear due to the gravitational potential. For smaller radii the clouds either collapse or become tidally disrupted. We suggest that this mechanism is responsible for the formation of the inner edge of the Circumnuclear Disk in the Galactic Centre. | astro-ph | astro-ph | Astronomy & Astrophysics manuscript no.
(will be inserted by hand later)
1
0
0
2
g
u
A
2
2
1
v
4
5
3
8
0
1
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
The stability of the Circumnuclear Disk clouds
in the Galactic Centre
B. Vollmer1
,
2 and W.J. Duschl2
,
1
1 Max-Planck-Institut fur Radioastronomie, Auf dem Hugel 69, 53121 Bonn, Germany.
2 Institut fur Theoretische Astrophysik der Universitat Heidelberg, Tiergartenstrasse 15, 69121 Heidelberg,
Germany.
Received / Accepted
Abstract. The influence of rotation and magnetic fields on the physical properties of isothermal gas clouds is
discussed. The presence of rotation and/or magnetic fields results in an increase of the critical cloud mass with
respect to gravitational instability for clouds of a given temperature and external pressure. Rotating clouds
have higher densities. Consequently, they are more stable against tidal shear than non-rotating clouds. They
can approach the Galactic Centre up to a radius of ∼2 pc without being disrupted by the tidal shear due to
the gravitational potential. For smaller radii the clouds either collapse or become tidally disrupted. We suggest
that this mechanism is responsible for the formation of the inner edge of the Circumnuclear Disk in the Galactic
Centre.
Key words. ISM: clouds -- ISM: evolution -- Galaxy: center
1. Introduction
The Galactic Centre1 is surrounded by a large number
of gas and dust clouds. The distribution and the kine-
matics of these clouds are generally interpreted as a thick
disk (Circumnuclear Disk CND). Whether this disk -- like
structure has an outer edge is still a matter of debate.
The CND was studied in molecular lines by Gatley et
al. (1986) (H2), Serabyn et al. (1986) (CO, CS), Gusten
et al. (1987) (HCN), DePoy et al. (1989) (H2), Sutton
et al. (1990) (CO), Jackson et al. (1993) (HCN), and
Marr, Wright, & Backer (1993) (HCN), Coil & Ho (1999,
2000) (NH3), Wright et al. (2001) (HCN). The disk has a
sharply defined inner edge at a radius of ∼1.7 pc. Inside
this radius the gas density drops by more than one or-
der of magnitude. Gusten et al. (1987) suggested that
the density of the clouds is not high enough to resist
tidal shear due to the centre's strong gravitational field.
Therefore, they concluded that the CND is a very short
lived structure (∼ 105 yr). On the other hand, Jackson
et al. (1993) pointed out that the possibility should not
be dismissed that the clouds' density is even high enough
to stabilize individual gas clouds against tidal disruption
(n ∼ 107 cm−3). In a previous paper (Vollmer & Duschl
2001; hereafter VD2001) we constructed an analytic model
for a clumped gas and dust disk and applied it to the
Send offprint requests to: B. Vollmer, e-mail: bvollmer@mpifr-
bonn.mpg.de
1 We assume 8.5 kpc for the distance to the Galactic Centre.
CND. The clouds were treated as isothermal selfgravitat-
ing spheres with a given outer pressure. We succeeded in
reproducing the main characteristics of the observed gas
clouds (mass, central density, density at the outer bound-
ary). A major shortcoming of this model was that these
clouds could not resist tidal shear at distances smaller
than ∼2.5 pc from the Galactic Centre. In this work we
investigate the influence of rotation on the cloud proper-
ties and their stability against tidal shear.
2. Basic Picture
As in VD2001 we assume that during a short accretion
event (∆t ∼ 106 yr) an amount of gas of several 104 M⊙
is driven into the Galactic Centre region to distances less
than 10 pc. The infalling gas has not a uniform density
distribution and might be turbulent. This clumpy medium
is exposed to the ambient UV radiation field due to the
population of young O/B stars in the Galactic Centre.
Low density regions of the infalling gas are evaporated
rapidly while regions of higher density stay molecular and
are heated to an equilibrium temperature during less than
an orbital period. In this way a clumpy circumnuclear disk
is formed. We assume that the clouds are gravitationally
stable. Three kinds of pressure can counterbalance the
selfgravitation of the clouds:
2
B. Vollmer and W.J. Duschl: The stability of the CND clouds in the GC
(i) Turbulent pressure:
if the turbulent pressure is
dominant, we can estimate the turbulent velocity disper-
sion σ with the help of the Virial theorem:
σ2 ∼
2GMcl
rcl
,
(1)
is the cloud mass and rcl
where Mcl
its radius. With
Mcl=30 M⊙ and rcl=0.05 pc, the velocity dispersion is
σ ∼ 2.3 km s−1 (note that the observed FWHM is two
times σ).
(ii) Magnetic pressure: in the case of dominant mag-
netic pressure, we can estimate an equivalent linewidth
c using a local magnetic field strength B=3 mG (Yusef-
Zadeh et al. 1996).
(2)
B2/8π ≃ c2ρcl ,
where ρcl is the cloud density. With a cloud density of
ρcl=3 10−18 g cm−3 we obtain an equivalent linewidth of
c ∼ 3.3 km s−1.
(iii) Thermal pressure: if the thermal pressure dom-
inates, the sound velocity at a temperature of 150 K is
cs ∼ 0.8 km s−1. Thus, the linewidth c of a single selfgrav-
itating cloud is 1 km s−1 ≤ c ≤ 3 km s−1.
The observed linewidths are as large as ∆v =
50 km s−1 (see, e.g., Wright et al. 2001). If the clouds
within the CND are gravitationally stable, their intrin-
sic linewidth is much smaller. A possible explanation of
the large observed linewidths is an enhanced turbulent
velocity dispersion between the clouds due to infalling
gas streamers and/or superposition of clouds with differ-
ent rotational and turbulent velocities (the turbulent ve-
locity dispersion of the model disk of VD2001 is of the
order ∆v ∼ 15 km s−1). The study of a CS(3 -- 2) data
cube observed with the IRAM 30m Telescope (Zylka et
al. 1999) indicates that the features of linewidths up to
50 km s−1 might consist of several clouds with small indi-
vidual linewidths (∼ 2 km s−1) (Zylka, private communi-
cation).
In the following we discuss rotating clouds without a
magnetic field first. As shown above, the inclusion of a
magnetic field increases the equivalent linewidth of the
clouds by a factor ∼3. The implications of the inclusion
of a magnetic field will be discussed in Sect. 6.
3. Cloud Rotation
The equilibrium state of rotating isothermal clouds was
studied both analytically (Hayashi, Narita, & Miyama
1982; Tohline 1985a, b) and numerically (Stahler 1983a,
b; Kiguchi et al. 1987; Narita et al. 1990). These au-
thors showed that the density and the mass of rotating
clouds can exceed that of non-rotating clouds consider-
ably. Kiguchi et al. (1987) carried out numerical simula-
tions of rotating isothermal spheres embedded in a tenuous
intercloud medium, i.e. the outer boundary of the cloud
is determined by the external pressure. Their calculations
(i) M rot
crit/M BE
did not include a magnetic field. They concluded that a
rotating cloud is dynamically stable if
crit < 31, where M rot
crit is the critical mass
for gravitational instability of the rotating cloud and M BE
crit
is the critical mass for gravitational instability of a non-
rotating Bonner -- Ebert sphere with the same sound speed
(temperature) and the same outer pressure;
(ii) the maximum mean rotation velocity is smaller
than 2.7 cs, where cs is the sound velocity;
(iii) ρ/ρext < 6, where ρ is the mean density of the
cloud and ρext is the external density.
Thus, rotating isothermal clouds, which are gravita-
tionally stable are denser and more massive than non-
rotating clouds of the same temperature embedded in a
medium of the same pressure. We conclude that cloud ro-
tation alters the solution for the physical characteristics
of our model clouds (VD2001) in such a way that more
massive and denser clouds result which are more stable
against tidal disruption.
4. Specific angular momentum
4.1. Cloud formation
The spin angular momentum of a cloud in the CND should
reflect the angular momentum of the interstellar medium
from which the cloud has formed (see e.g. Blitz 1993).
Within the central 10 pc the gravitational potential is ap-
proximatly spherical. The rotation curve is determined for
distances smaller than approximatelly 1 pc by the gravita-
tional potential of the central black hole and for distances
greater than this by that of the nuclear star cluster (Genzel
et al. 1996). If we consider that a cloud forms when a
disk-like region becomes gravitationally unstable and col-
lapses, we expect a similar result to that considered ana-
lytically by Mestel (1966). Since the rotation curve in the
Galactic Centre is approximately constant for distances
greater 1 pc and is falling inwards for smaller distances,
the rotation of the cloud can be either prograde or ret-
rograde depending on the details of the cloud formation.
We thus have to estimate the specific angular momen-
tum of the ISM at the place of cloud formation. For an
isolated CND, VD2001 found a very small, mean, radial
drift velocity of the clouds (vdrift < 0.01 km s−1; however,
a single cloud can approach the Galactic Centre faster).
In this case we can assume that the clouds were formed
not far away from the galactic distances where we observe
them today (2 pc ≤ RG ≤ 7 pc). Mestel (1966) has shown
that a cloud can have a specific angular momentum about
its mass-centre up to J/Mcl = 0.5R2Ω, where R is half
the size of the collapsed region and Ω is the local angular
velocity. Let us consider a cloud which forms at a distance
greater than 2 pc from the Galactic Centre. With a typ-
ical density in the disk ρ = 2 10−19 g cm−1 and a typical
cloud mass of 30 M⊙ (VD2001), the size of the collapsing
region is R ≃0.1 pc. The initial value of Ω depends on the
details of the collapse. With Ω(2 pc) ≤ 60 km s−1 pc−1
we obtain a specific spin angular momentum of J/M ≤
B. Vollmer and W.J. Duschl: The stability of the CND clouds in the GC
3
0.3 pc km s−1. If the CND is not isolated, the clouds could
have formed at higher distances from the Galactic Centre
(∼10 pc). Using Ω(10 pc) ∼12 km s−1 pc−1, gives J/M ≤
6 10−2 pc km s−1. In addition, partially inelastic off-center
collisions between the clouds can lead to changes in the
spin angular momentum of the clouds.
4.2. Stability criteria
Limits on the specific angular momentum of the clouds
J/M are given by the onset of bar formation on the one
hand and the onset of core collapse on the other hand.
Kiguchi et al. (1987) suggested that bar formation occurs
for β0 > 1/3, where
β0 =
25
12(cid:0)
4π
3 (cid:1)
1
3
1
3 ρ
extJ 2
GM
10
3
(3)
is the ration of rotational energy to gravitational energy
ratio of the cloud. Here ρext is the boundary density of
the cloud, J is the angular momentum of the cloud, M its
mass, and G the gravitation constant. The criterion for
clouds which are stable against bar formation is thus
1
−
1
6
2
2 ρ
ext M
3 ∼ 7 10−2 pc km s−1 ,
J/M ≤ 0.31G
where we have used typical cloud parameters (M =30 M⊙,
ρext = 3 10−19 g cm−3). Miyama, Hayashi, & Narita
(1984) found that rotating isothermal spheres collapse if
α0β0 < 0.2, where
(4)
α0 =
5
2(cid:0)
4π
3 (cid:1)−
1
3
c2
s
Gρ
1
3
extM
2
3
(5)
is the thermal to gravitational energy of the cloud. The
criterion for stable clouds thus is
J/M ≥ 0.2
GM
cs ∼ 5 10−2 pc km s−1 ,
(6)
using the cloud parameters from above and cs=1 km s−1.
In Sect. 6 it is shown that the presence of magnetic
fields with a field strength of several mG results in an
effective linewidth of c ∼ 3 km s−1. In this case the specific
angular momentum is
GM
c ∼ 2 10−2 pc km s−1 ,
J/M ≥ 0.2
We thus conclude that for a stable rotating cloud 2 10−2 ≤
J/M ≤ 7 10−2 pc km s−1. This is consistent with the upper
limits 6 10−2 ≤ J/M ≤ 0.3 pc km s−1 at the moment of
cloud formation (Sect. 4.1).
(7)
5. A typical rotating cloud
The maximum central density ρc of an isothermal cloud
with external pressure Pext (Bonner -- Ebert sphere) is
ρc/ρ∗ = 14, where ρ∗ = c−2
s Pext. Stable rotating isother-
mal clouds can have central densities up to ρc/ρ∗ ∼ 100.
− 1
2
s P
ext G− 3
Fig. 14 of Kiguchi et al. (1987) shows the region of sta-
ble clouds as a function of the central density and the
cloud mass. In our case M∗ = c4
2 =15 M⊙ and
ρ∗ = 10−18 g cm−3. Since we are interested in rotating
clouds that have higher central densities than non-rotating
clouds, we will only discuss the region for ρc/ρ∗ > 14. In
this case our region of interest is limited by the dashed col-
lapse curve and the dash-dotted curve for bar formation
in Fig. 14 of Kiguchi et al. (1987). A typical cloud in this
∼ 3(cid:0)ρc/ρ∗(cid:1)BE
region has M/M∗ ∼ 2 − 3 or (cid:0)ρc/ρ∗(cid:1)rot
.
Such a cloud has the following physical characteristics
(Kiguchi et al. 1987; Table 3): β0 = 0.16, ρc/ρ∗ =
30, M/M∗ = 2.38, csJ/(GM 2) = 0.164, ρ/ρ∗ = 2.71. In
physical units this gives: M =36 M⊙, ρ = 2.7 10−18 g cm−3,
ρc = 3 10−17 g cm−3. The radius perpendicular to the ro-
tation axis is re = 8 10−2 pc and parallel to the rotation
axis z = 3 10−2 pc. The maximum mean rotation velocity
is vrot ∼ 0.8 km s−1.
This cloud has thus a 3 times higher central density
than a non-rotating isothermal selfgravitating sphere of
the same temperature which is embedded in a tenuous
medium of the same outer pressure.
The clouds are illuminated by the UV radiation field
coming from the central He i star cluster. The mass loss
rate of these clouds due to outflowing ionized gas is ap-
proximately M ∼ 4πr2
clρclci, where rcl is the cloud radius
and ci is the sound velocity in the external ionized gas.
cl < 2 × dBE
Since for the diameter of such a cloud drot
,
rotation lowers the lifetime of a cloud by a factor smaller
than 4. The minimum lifetime is still greater than 107 yr.
cl
6. Magnetic fields
Tomisaka et al. (1989) studied the structure of selfgravita-
tionally stable magnetized clouds. They showed that the
critical mass for stability is
Mmag ∼ 1.18(cid:16)1 − (cid:0)
G
1
0.17
2dMcl/dΦBc(cid:1)2(cid:17)− 3
2
MBE ,
(8)
cl/2(cid:1)−1
where dMcl/dΦBc represents the mass-to-magnetic flux
ratio at the center and MBE is the Bonner -- Ebert criti-
cal mass in the absence of magnetic fields. If we approxi-
mate dMcl/dΦBc ≃ Mcl(cid:0)Br2
and let Mcl =30 M⊙,
rcl=0.05 pc, and B=5 mG, we obtain Mmag ∼ 2.5MBE.
This is consistent with a realistic Bonner -- Ebert mass of
MBE ∼ 10 M⊙ (VD2001). We would thus expect that the
density of a magnetized cloud of size dcl=0.1 pc, mass
Mcl=30 M⊙, and magnetic field strength B of several mG
would also show a density enhancement of a factor 3 with
respect to a non-rotating, non-magnetized isothermal, self-
gravitating cloud.
4
B. Vollmer and W.J. Duschl: The stability of the CND clouds in the GC
c
We have argued in the previous Sections that for a
typical rotating or magnetized cloud ρrot/mag
∼ 3 × ρBE
(Fig. 1 dashed line). The presence of rotation and/or mag-
netic fields alters the properties of the cloud, i.e. they are
more massive and have higher central densities. The crit-
ical Galactic radius thus decreases to Rcrit ∼ 2 pc. In the
framework of our model and its approximations, this is in
agreement with the observed inner edge of the CND at
RG ∼ 1.7 pc. The dependence of the inner edge on the
UV radiation field will be discussed in Sect. 9.
c
8. The influence of the stellar winds
The stellar wind emanating from the central Hei star clus-
ter which is responsible for the UV radiation field exerts
a ram pressure on the illuminated side of the cloud:
Pram ≃ 2.5 10−8 ( M /3 10−3 M⊙ yr−1)(v/600 km s−1)
4 π (R/2 pc)2
erg cm−3 ,(10)
(Yusef-Zadeh & Wardle 1993), where M is the stellar mass
loss rate due to a wind of velocity v. The energy density
due to the selfgravity of the cloud is given by
(Mcl/30 M⊙)G
erg cm−3 ,(11)
(rcl/0.05 pc)
Pgrav ≃ 5 10−8 (ncl/106 cm−3)
where Mcl and rcl are the cloud mass and mean ra-
dius. Ram pressure is thus able to push the ionized and
heated gas (ni ∼ several 103 cm−3) radially away from the
Galactic Centre, and might even shape the neutral con-
densations at the inner edge of the CND. It represents an
additional external force on the illuminated side of clouds
at the inner edge of the CND.
9. The inner edge of a CND
If we assume a spectrum of clouds moving around a galac-
tic centre forming a disk-like equilibrium structure, there
are four effects that determine the physical properties of
the clouds: UV radiation, tidal shear, a radially-directed
wind, and selfgravitation. In the following considerations
we will neglect the effects of stellar winds. Nevertheless,
one has to keep in mind that they can provide an addi-
tional external pressure on the clouds at the inner edge of
the CND. Only those clouds with a sufficiently high cen-
tral density can resist tidal disruption. Thus, the clouds'
mean density must increase with decreasing distance to
the Galactic Centre. If they reach the Jeans mass they be-
come gravitationally unstable and collapse. Taking these
two effects together we obtain an efficient mechanism to
create an inner edge. At this distance, the clouds that
can resist tidal disruption become Jeans unstable, i.e. the
dense cloud structure is lost. The UV radiation plays an
important role in determining the radius of the clouds at
each distance from the galactic centre. With an increas-
ing UV radiation field the cloud radius decreases, because
the ionization front in direction of the central star cluster
is located at smaller cloud radii (Dyson 1968). At these
Fig. 1. Central density of the clouds as a function of
the distance from the Galactic Centre. Solid: critical den-
sity for tidal disruption. Dotted: model central density of
VD2001. Dashed line: central density of a rotating cloud.
Dashed surface: range of densities where clouds are grav-
itationally and tidally stable.
7. Tidal shear
The critical density for a cloud orbiting at a distance R
on a circular orbit around a point mass M below which it
will disintegrate due to tidal shear is given by
ρcrit =
3
2π
M
R3
(9)
(see e.g. Stark et al. 1989). We approximate the
mass distribution in the Galactic Centre by M (R) =
M0 + M1 R1.25, where M0 = 3 106 M⊙ an M1 =
1.6 106 M⊙ pc−1.25 (this is close to the findings of Eckart
& Genzel (1996)). The obtained critical density can be
compared to the model density of VD2001 (Fig. 1). For
these clouds only selfgravitation and thermal pressure of
the neutral and ionized gas are taken into account. Since
our model clouds are on the edge of gravitational insta-
bility, their central density is close to the critical density
for gravitational collapse (Fig. 1 dotted line). As long as
this central density is higher than the critical central den-
sity with respect to tidal shear (Fig. 1 solid line), non-
rotating isothermal clouds can exist. The crossing of the
curves corresponds thus to the minimum distance where
stable clouds can exist. This critical Galactic radius is
Rcrit ∼ 3.5 pc for this estimate. More detailed calcula-
tions show Rcrit ∼ 2.5 pc for the illuminated side and
Rcrit ≤ 3.5 pc for the shadowed side. This difference is
due to the different mechanisms that create the cloud
boundary at the illuminated and shadowed side. At the
illuminated side, the density of the ionized gas is given
by the ionization -- recombination equilibrium. The heated
and ionized gas flows away from the cloud and fills the
space between the clouds. In this way, a low density ion-
ized interclump medium (ne ∼ 103 cm−3) is built up. The
external pressure of this interclump medium is responsible
for the outer edge of the cloud at the shadowed side.
B. Vollmer and W.J. Duschl: The stability of the CND clouds in the GC
5
radii the enhanced pressure of the ionized gas in the ion-
ization front is counterbalanced by the enhanced thermal
pressure due to the increasing density of the neutral gas
in the isothermal cloud. Thus the cloud mass decreases
with increasing UV radiation field and the cloud is less
susceptible to gravitational collapse.
We can estimate this effect quantitatively in the fol-
lowing way. In the case of a cloud with uniform density
the maximum density needed to stabilize a cloud against
tidal forces is
ρtidal
crit =
3
2π
M (R)
R3
,
(12)
where M (R) is the enclosed mass up to the radius R from
the Galactic Centre. The criterion for Jeans instability is
given by
ρJeans
crit =
πRT
Gµr2
cl
,
(13)
where T is the temperature of the cloud, R is the gas
constant, G is the gravitational constant, and µ is the
molecular weight.
The cloud radius due to the balance of ionization and
recombination is given by Dyson (1968)
rcl = ξ2J0n−2
i
,
(14)
where J0 is the number of incident UV photons per cm2
is the number density of the ionized gas in
and s, ni
the ionization front, and ξ = 4.87 106 cm− 3
2 . With
the jump condition across the ionization front ρcle−uc2
s =
2ρic2
i , where ρi and ci are the density and the sound ve-
locity of the external ionized gas and u(x) is a parameter
of the Lane-Emden equation, one can write
2 s
1
rcl = 4ξ2m2
pJ0ρ−2
cl e2u(
ci
cs
)4 .
(15)
Furthermore, one can approximate the function e−u(x) ∼
3x−2 for 2 < x < 8, where x = rcl/cs√4πGρc. This leads
to
Fig. 2. Central density of the heavy clouds versus the
distance to the Galactic Centre. Dashed line: maximum
central density above which gravitational collapse occurs.
Solid line: minimum density in order to resist tidal shear.
Dashed surface: range of densities where clouds are grav-
itationally and tidally stable.
In Fig. 2 the dashed line shows the maximum central den-
sity above which gravitational collapse occurs (ρJeans
crit using
Eq. 16), the solid line shows the minimum density in or-
der to resist tidal shear (ρtidal
crit ). The dashed surface shows
the range of densities where clouds are gravitationally and
tidally stable. The inner edge of the CND arises thus nat-
urally as a selection effect due to external conditions of
the environment on the cloud spectrum.
If there are non-negligible non-thermal pressure com-
ponents, we have to add these components to the sound
speed of the neutral and ionized gas. For a non-thermal
linewidth c > 1 km s−1, the critical Jeans density is always
lower than the critical tidal density, i.e. no stable clouds
can exist. Therefore, we suggest that only rotating clouds,
which have a higher central density can survive.
rcl = 3.645 1015J
− 1
3
0
− 4
3
i
c
8
3
s
c
.
(16)
9.1. A possible scenario for star formation
1
1
8
Thus, the cloud radius depends only on the number of in-
cident UV photons, the sound speed of the ionized gas,
and the sound speed of the neutral gas. If we assume
that the temperature is determined by the radiation field
cs ∝ T
0 , the cloud radius does not change with the
cloud's distance to the Galactic Centre. This means that
the cloud's radius is constant and does not depend on its
central density. Consequently, the critical Jeans density
ρJeans
is proportional to the gas temperature of the neu-
crit
tral gas.
2 ∝ J
Fig. 2 shows the critical densities with respect to tidal
shear and gravitational collapse. These graphs represent
an approximation of the density variations calculated from
the detailed model of VD2001 (Fig. 1). The increase or de-
crease of the neutral gas temperature by a factor 2 results
in a variation of the location of the inner edge of ±1 pc.
In the previous Sections we have investigated an isolated
clumpy gas disk. However, the CND appears to interact
with the surrounding gas. Coil & Ho (1999, 2000) and
Zylka et al. (1999) conclude on the basis of the gas distri-
bution and kinematics in the inner 20 pc of the Galactic
Centre that there are connections between the CND and
the neighbouring GMCs. They claim that there are several
streamers that fall into the Galactic Centre.
From the theoretical point of view Sanders (1998)
pointed out the possibility that the CND can be under-
stood in terms of tidal capture and disruption of gas clouds
falling into the Galactic Centre region. The infalling gas
forms a tidally stretched filament intersecting itself. After
several rotation periods the gas forms a stable ring struc-
ture which can be maintained for more than 106 yr. He
showed that the central star cluster can be created within
6
B. Vollmer and W.J. Duschl: The stability of the CND clouds in the GC
the first few passages of the cloud when the long filament
intersects itself at a large angle.
We will now discuss what happens when a cloud com-
plex falls from a distance greater than 10 pc into the
Galactic Centre.
In VD2001 we have shown that the CND has a life-
time of ∼ 107 yr. It is thus possible that an external cloud
is falling into the Galactic Centre within this period. We
propose a new scenario in which a whole cloud complex
is falling into the Galactic Centre where a clumpy disk
structure already exists. When the cloud hits the CND,
frequent partially inelastic cloud -- cloud collisions will cre-
ate a whole transient spectrum of clouds with different
masses. Those clouds which have masses above the Jeans
limit will collapse and eventually form stars. The massive
stars are thus formed within a very short time during the
collision of the cloud complex and the disk. Partially in-
elastic collisions result in a decrease of the cloud velocities.
Low velocity clouds (v < vrot ∼ 120 km s−1) form stars
and approach the Galactic Centre at the same time. Thus,
if an external gas cloud collides with a pre-existing CND,
we expect that a part of the gas, which loses angular mo-
mentum due to cloud -- cloud collisions spirals inwards with
a velocity of ∼100 km s−1.
If we assume an initial He i star mass of M0 = 30
M⊙ with an initial velocity of V0 = 100 km s−1, falling
into an already existing old stellar cluster of density
ρstellar = 4 106 M⊙pc−3 with stellar masses around 1 M⊙,
collective relaxation dominates over the particle-particle
relaxation (Saslaw 1985). This leads to a relaxation time of
cated in the Circumnuclear Disk in the Galactic Centre.
Rotating selfgravitating isothermal clouds of a given tem-
perature T embedded in a tenuous medium giving rise
to an external pressure Pext are more massive and have
larger densities than non-rotating clouds of same T and
Pext. Stable rotating clouds have an angular momentum
in the range 3 10−2 ≤ J/M ≤ 5 10−2 pc km s−1. This rep-
resents ∼20% -- 100% of the maximum specific angular mo-
mentum that a cloud can acquire during its formation.
These clouds are stable against tidal shear for Galactic
Radii RG ≥ 2 pc. Magnetized selfgravitating isothermal
clouds with magnetic field strengths of several mG have a
critical mass with respect to gravitational collapse of ∼3
times the corresponding Bonner -- Ebert critical mass of a
selfgravitating isothermal sphere.
We suggest a mechanism for the formation of an inner
edge in circumnuclear disks. The external UV radiation
field determines the diameter of the clouds. The density
of the clouds must increase with decreasing galactic radius
because of the tidal shear due to the gravitational poten-
tial in the Galactic Centre. At a critical radius clouds that
are stable against tidal shear become gravitational unsta-
ble and collapse. On the basis of our model, selfgravitating
non-magnetized isothermal clouds should rotate in order
to resist tidal shear at a distance of 2 pc from the Galactic
Centre, whereas magnetized clouds must rotate in order
to resist tidal shear. For the CND in the Galactic Centre
the critical radius corresponds approximately to the ob-
served inner edge if cloud rotation is included. We suggest
that the central Hei star cluster has been formed when an
external cloud collided with a pre-existing CND.
τR,coll = 7 106×(cid:16)
V0
100 km s−1(cid:17)3(cid:16) M0
30 M⊙
(cid:17)−1(cid:16) ρstellar
4 106 M⊙
pc3
(cid:17)−1
yr .(17)
Acknowledgements. We would like to thank P. Ho for helping
us to improve this article significantly.
A single massive star, whose velocity is approximately the
Keplerian velocity of gas moving on circular orbits around
the Galactic Centre, thus loses the information about its
initial position and velocity within ∼7 106 yr.
Gerhard (2000) estimated that a star cluster of ∼
105 M⊙ which is formed at a Galactic radius RG=10 pc
needs several Myr to spiral
into the Galactic Centre.
This timescale is comparable to our collective relaxation
timescale.
The star cluster relaxation time is thus comparable
to the lifetime of the Hei stars. These massive stars are
formed at the same time and at the same distance to the
Galactic Centre forming a star cluster after a few million
years. The observed streaming motions of the Hei stars
in the Galactic Centre (Genzel et al. 1996) shows that
the star cluster is not completely relaxed. Therefore, it is
possible that the central Hei star cluster has been built
during the collision of an infalling cloud complex with an
already existing CND.
10. Conclusion
References
Blitz L., 1993, in: Protostars and Planets III, University of
Arizona Press, p. 125
Coil A.L. & Ho P.T.P., 1999, ApJ, 513, 752
Coil A.L. & Ho P.T.P., 2000, ApJ, 533, 245
DePoy D.L., Gatley I., McLean I.S., 1989, IAU-Symp. 136, 361
Dyson J.E., 1968, ApSS 1, 388
Eckart A., Genzel R., 1996, Nature 383, 415
Gatley I., Jones J.J., Hyland A.R. et al., 1986, MNRAS 222,
299
Genzel R., Thatte N., Krabbe A. et al., 1996, ApJ 472, 153
Gerhard O., 2000, ApJ, in press
Gusten R., Genzel R., Wright M.C.H. et al., 1987, ApJ 318,
124
Hayashi C., Narita S., & Miyama S.M., 1982, Prog. Theor.
Phys., Osaka, 68, 1949
Jackson J.M., Geis N., Genzel R. et al., 1993, ApJ 402, 173
Kiguchi M., Narita S., Miyama S.M., & Hayashi C., 1987, ApJ,
317, 830
Marr J.M., Wright M.C.H., Backer D.C., 1993, ApJ 411, 667
Mestel L., 1966, MNRAS, 131, 307
Narita S., Kiguchi M., Miyama S.M., & Hayashi C., 1990, 244,
We investigated the influence of rotation and magnetic
fields on the physical properties of the gas clouds lo-
349
Sanders R.H., 1998, MNRAS, 1998, 294
B. Vollmer and W.J. Duschl: The stability of the CND clouds in the GC
7
Serabyn E., Gusten R., Walmsley C.M. et al., 1986, A&A 169,
85
Stahler S.W., 1983a, ApJ, 268, 155
Stahler S.W., 1983b, ApJ, 268, 165
Stark A.A., Bally J., Wilson R.W., Pound M.W., 1989, IAU-
Symp. 135, 129
Sutton E.C., Danchi W.C., Jaminet P.A., Masson C.R., 1990,
ApJ 348, 503
Tohline J.E., 1985a, Icarus, 61, 10
Tohline J.E., 1985b, ApJ, 292, 181
Vollmer B., Duschl W.J., 2001, A&A, 367, 72 (=VD2001)
Wright M.C.H., Coil A.L., McGary R.S., Ho P.T.P., & Harris
A.I., 2001, ApJ, 551, 254
Yusef-Zadeh F., Wardle M., 1993, ApJ, 405, 584
Yusef-Zadeh F., Roberts D.A., Goss W.M., Frail D.A., & Green
A.J., 1996, ApJ, 466, L25
Zylka R., Gusten R., Philipp S. et al., 1999, in: The Central
Parsecs of the Galaxy, ed. H. Falke, A. Cotera, W.J. Duschl,
F. Melia, & M.J. Rieke, ASP Conference Series, Vol. 186,
p.415
|
astro-ph/0103212 | 1 | 0103 | 2001-03-14T15:51:04 | Kinematics of young stars. II. Galactic spiral structure | [
"astro-ph"
] | The young star velocity field is analysed by means of a galactic model which takes into account solar motion, differential galactic rotation and spiral arm kinematics. We use two samples of Hipparcos data, one containing O- and B-type stars and another one composed of Cepheid variable stars. The robustness of our method is tested through careful kinematic simulations. Our results show a galactic rotation curve with a classical value of $A$ Oort constant for the O and B star sample ($A^{\mathrm{OB}} =$ 13.7-13.8 km s$^{-1}$ kpc $^{-1}$) and a higher value for Cepheids ($A^{\mathrm{Cep}} =$ 14.9-16.9 km s$^{-1}$ kpc $^{-1}$, depending on the cosmic distance scale chosen). The second-order term is found to be small, compatible with a zero value. The study of the residuals shows the need for a $K$-term up to a heliocentric distance of 4 kpc, obtaining a value $K = -$(1-3) km s$^{-1}$ kpc$^{-1}$. The results obtained for the spiral structure from O and B stars and Cepheids show good agreement. The Sun is located relatively near the minimum of the spiral perturbation potential ($\psi_\odot =$ 284-20$\degr$) and very near the corotation circle. The angular rotation velocity of the spiral pattern was found to be $\Omega_{\mathrm{p}} \approx 30$ km s$^{-1}$ kpc $^{-1}$. | astro-ph | astro-ph | Astronomy & Astrophysics manuscript no.
(will be inserted by hand later)
1
0
0
2
r
a
M
4
1
1
v
2
1
2
3
0
1
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
Kinematics of young stars ⋆
II. Galactic spiral structure
D. Fern´andez1, F. Figueras1, and J. Torra1
,
2
1 Departament d'Astronomia i Meteorologia, Universitat de Barcelona, Av. Diagonal 647, E-08028 Barcelona,
Spain
2 IEEC (Institut d'Estudis Espacials de Catalunya), Edif. Nexus-104, Gran Capit`a 2-4, E-08034 Barcelona, Spain
Received ¡date¿ / Accepted ¡date¿
Abstract. The young star velocity field is analysed by means of a galactic model which takes into account
solar motion, differential galactic rotation and spiral arm kinematics. We use two samples of Hipparcos data,
one containing O- and B-type stars and another one composed of Cepheid variable stars. The robustness of our
method is tested through careful kinematic simulations. Our results show a galactic rotation curve with a classical
value of A Oort constant for the O and B star sample (AOB = 13.7-13.8 km s−1 kpc −1) and a higher value for
Cepheids (ACep = 14.9-16.9 km s−1 kpc −1, depending on the cosmic distance scale chosen). The second-order
term is found to be small, compatible with a zero value. The study of the residuals shows the need for a K-term
up to a heliocentric distance of 4 kpc, obtaining a value K = −(1-3) km s−1 kpc−1. The results obtained for the
spiral structure from O and B stars and Cepheids show good agreement. The Sun is located relatively near the
minimum of the spiral perturbation potential (ψ⊙ = 284-20◦) and very near the corotation circle. The angular
rotation velocity of the spiral pattern was found to be Ωp ≈ 30 km s−1 kpc −1.
Key words. Galaxy: kinematics and dynamics -- Galaxy: solar neighbourhood -- Galaxy: structure -- Stars: early-
type -- Stars: kinematics -- Stars: variables: Cepheids
1. Introduction
Young stars have been traditionally used as probes of the
galactic structure in the solar neighbourhood. Their lu-
minosity makes them visible at large distances from the
Sun, and their age is not great compared to the dynamical
evolution timescales of our galaxy.
These stars show kinematic characteristics that can-
not be explained by solar motion and differential galactic
rotation alone. Small perturbations in the galactic grav-
itational potential induce the formation of density waves
that can explain part of the special kinematic features of
young stars in the solar neighbourhood and also the ex-
istence of spiral arms in our galaxy. The first complete
mathematical formulation of this theory was done by Lin
and his associates (Lin & Shu 1964; Lin et al. 1969).
Lin's theory has several free parameters that can be de-
rived from observations. Two of the main parameters are
the number of spiral arms (m) and their pitch angle (i).
Although the original theory proposed a 2-armed spiral
Send offprint requests to: D. Fern´andez,
e-mail: [email protected]
⋆ Based on data from the Hipparcos astrometry satellite
(European Space Agency)
structure with i = −6◦, as early as the mid-70s Georgelin
& Georgelin (1976) found 4 spiral arms with i = −12◦
from a study of the spatial distribution of HII regions.
This controversy is still not resolved: in a review Vall´ee
(1995) concluded that the most suitable value is m = 4,
whereas Drimmel (2000) found that emission profiles of
the galactic plane in the K band -- which traces stellar
emission -- are consistent with a 2-armed pattern, whereas
the 240 µm emission from dust is compatible with a 4-
armed structure. In a recent paper, L´epine et al. (2001)
described the spiral structure of our Galaxy in terms of a
superposition of 2- and 4-armed wave harmonics, studying
the kinematics of a sample of Cepheids stars and the l-v
diagrams of HII regions.
The angular rotation velocity of the spiral pattern Ωp
is another parameter of the galactic spiral structure. It
determines the rotation velocity of the spiral structure as
a rigid body. The classical value proposed by Lin et al.
(1969) is Ωp ≈ 13.5 km s−1 kpc−1. The angular rota-
tion velocity in the solar neighbourhood due to differen-
tial galactic rotation is Ω⊙ ≈ 26 km s−1 kpc−1 (Kerr &
Lynden-Bell 1986). Thus, the value of Ωp implies that the
so-called corotation circle (the galactocentric radius where
Ωp = Ω) is in the outer region of our galaxy (cor ≈ 15-20
2
D. Fern´andez et al.: Kinematics of young stars
kpc, depending on the galactic rotation curve assumed).
Nevertheless, several authors found higher values of Ωp,
about 17-29 km s−1 kpc−1 (Marochnik et al. 1972; Cr´ez´e
& Mennessier 1973; Byl & Ovenden 1978; Avedisova 1989;
Amaral & L´epine 1997; Mishurov et al. 1997; Mishurov
& Zenina 1999; L´epine et al. 2001). These values place
the Sun near the corotation circle, in a region where the
difference between the galactic rotation velocity and the
rotation of the spiral arms is small. This fact has very
important consequences for the star formation rate in the
solar neighbourhood, since the compression of the inter-
stellar medium due to shock fronts induced by density
waves could be chiefly responsible for this process (Roberts
1970).
Other parameters of Lin's theory are the amplitudes
of induced perturbation in the velocity (in the antigalac-
tocentric and the galactic rotation directions) of the stars
and gas, and the phase of the spiral structure at the Sun's
position. The interarm distance and the phase of the spiral
structure can be determined from optical and radio obser-
vations (Burton 1971; Bok & Bok 1974; Schmidt-Kaler
1975; Elmegreen 1985). The interarm distance gives us a
relation between the number of arms and the pitch angle.
In this paper we obtain the galactic kinematic param-
eters from two samples of Hipparcos stars described in
Sect. 2: one that contains O- and B-type stars, and an-
other one composed of Cepheid variable stars. In Sect. 3
we propose a model of our galaxy, a generalization of that
previously used by Comer´on & Torra (1991). The authors
only applied their model to radial velocities. The accu-
rate astrometry of the Hipparcos satellite offers a good
opportunity to also use the proper motion data. The res-
olution of the condition equations, based on a weighted
least squares fit, is explained in Sect. 4. An extensive set
of simulations is performed in Sect. 5 in order to assess the
capabilities of the method, that is, to analyze the influence
of the observational errors and biases in the kinematic pa-
rameters. We finally present our results and discussion in
Sect. 6 and 7, respectively.
2. The working samples
2.1. Sample of O and B stars
Our initial sample contains 6922 Hipparcos O- and B-type
stars. The astrometric data for these stars come from the
Hipparcos Catalogue (ESA 1997), radial velocities from
the compilation of Grenier (1997) and Stromgren pho-
tometry from Hauck & Mermilliod's (1998) catalogue.
The procedure followed to elaborate the sample consid-
ered here is fully described in Torra et al. (2000; hereafter
referred to as Paper I) along with a study of the possible
biases in the trigonometric distances and the availabil-
ity of radial velocities. Our sample contains 3915 stars
with known distance and proper motions (2272 stars with
known radial velocity). In Paper I we characterized the
structure and kinematics of the Gould Belt system using
this sample of O and B stars, establishing its boundary
−2.0
−1.0
)
c
p
k
(
X
0.0
1.0
2.0
−2.0
−1.0
0.0
Y (kpc)
1.0
2.0
Fig. 1. Star distribution in the X-Y galactic plane for the
sample of O and B stars with 0.6 < R < 2 kpc.
to a distance of about 0.6 kpc from the Sun. Taking into
account the kinematic peculiarities of the Gould Belt, in
this paper we do not consider those stars with R < 0.6
kpc. In Fig. 1 we show the position of those stars with
0.6 < R < 2 kpc projected on the galactic plane (X pos-
itive towards the galactic center and Y positive towards
the galactic rotation direction), which is our working sam-
ple (448 stars; 307 of them with distance, radial velocity
and proper motions and 141 with only distance and proper
motions).
2.2. Sample of Cepheid stars
The initial sample contains all the Hipparcos classi-
cal Cepheids. Astrometric data were taken from the
Hipparcos Catalogue (ESA 1997), whereas radial veloc-
ities come from Pont et al. (1994, 1997).
Individual distances were computed following two
period-luminosity (PL) relations. In both, periods come
from the Hipparcos Catalogue (ESA 1997) and individual
reddenings from Fernie et al.'s (1995) compilation (contin-
uous updating). A classification between fundamental and
overtone Cepheids from light curves and Fourier analysis
was adopted (Beaulieu 1999), using only the former in the
kinematic analysis.
The first PL relation (Luri 2000) adopts a slope from
EROS (Beaulieu 1999) and corresponds to the short cos-
mic distance scale:
M Short
v
= −1.08 − 2.72 log P
(1)
On the other hand, the second PL relation (Feast &
Catchpole 1997) corresponds to the large distance scale:
M Large
v
= −1.41 − 3.46 log P
(2)
D. Fern´andez et al.: Kinematics of young stars
3
−4.0
−2.0
)
c
p
k
(
X
0.0
2.0
4.0
−4.0
−2.0
0.0
Y (kpc)
2.0
4.0
Fig. 2. Star distribution in the X-Y galactic plane for the
sample of Cepheid stars with 0.6 < R < 4 kpc. Distances
computed from a short cosmic distance scale (Luri 2000).
The sample contains 186 stars with known distance and
proper motions (165 stars with radial velocity). Their dis-
tribution in the galactic plane (0.6 < R < 4 kpc) is shown
in Fig. 2 (164 stars; 145 of them with distance, radial ve-
locity and proper motions and 19 with only distance and
proper motions).
3. A galactic kinematic model for the solar
neighbourhood
where we show the relationship between ar and the A Oort
constant. B Oort constant can be derived from:
B = A −
Θ(⊙)
⊙
(5)
We considered as free parameters the galactocentric dis-
tance of the Sun (⊙) and the circular velocity at the
Sun's position (Θ(⊙)). Finally, spiral arm kinematics
was modelled within the framework of Lin's theory. We
considered the number of the spiral arms (m) and their
pitch angle (i) as free parameters, and we derived the
perturbation velocity amplitudes in the antigalactocentric
and tangential directions (Πb and Θb, respectively; they
were considered as constant magnitudes, assuming that
their variation with the galactocentric distance is smooth),
the phase of the spiral structure at the Sun's position (ψ⊙)
and the parameter f⊙, which takes into account the dif-
ference in the velocity dispersion between the solar-type
stars and the considered stars. This parameter is defined
by the relationship between the spiral perturbation veloc-
ity amplitudes for the sample stars (Πb, Θb) and for the
Sun (Πb⊙, Θb⊙):
Πb⊙ =
Θb⊙ =
1 − ν2 + xstars
1 − ν2 + x⊙
1 − ν2 + xstars
1 − ν2 + x⊙
Πb ≡ f⊙Πb
Θb ≡ f⊙Θb
(6)
where ν is the dimensionless rotation frequency of the
spiral structure and x is the stability Toomre's number
(Toomre 1969), which depends on the velocity dispersion
of the considered stars (see details in Appendix A). The
inclusion of f⊙ is new with regard to the model proposed
by Comer´on & Torra (1991).
Eqs. (3) can be expressed as:
3.1. Systematic velocity field in the galactic model
10
We propose an extension of the bidimensional model pre-
viously applied for radial velocities by Comer´on & Torra
(1991). In our model we considered the systematic velocity
components:
vr(R, l, b) =
vl(R, l, b) =
vr = vr1 + vr2 + vr3
vl = vl1 + vl2 + vl3
vb = vb1 + vb2 + vb3
vb(R, l, b) =
(3)
10
Xj=1
Xj=1
Xj=1
10
ajf r
j (R, l, b)
ajf l
j(R, l, b)
ajf b
j (R, l, b)
(7)
where subindex 1 refers to the solar motion contribution,
subindex 2 to differential galactic rotation and subindex 3
to spiral arm kinematics, respectively (see Appendix A).
Solar motion is expressed through the three components
of the Sun's velocity in galactic coordinates (U⊙, V⊙, W⊙).
Galactic rotation curve was developed up to second-order
approximation, ar and br being the first- and second-order
terms, respectively:
where the constants aj contain combinations of the kine-
matic parameters we wish to determine (U⊙, V⊙, W⊙, ar,
br, ψ⊙, Πb, Θb and f⊙) and f i
j (R, l, b) are functions of
the heliocentric distance and the galactic longitude and
latitude (see Eqs. (A.21) and (A.22)).
3.2. Free parameters of our galactic model
=
∂(cid:19)⊙
ar = (cid:18) ∂Θ
2(cid:18) ∂2Θ
∂2(cid:19)⊙
br =
1
Θ(⊙)
⊙
− 2A
A 2-armed Galaxy was the first proposed view for our stel-
lar system, mainly derived from HI and HII observations,
but also from the spatial distribution of supergiant stars
and other bright objects. These classical studies show the
existence of at least two arms inside the solar circle (the
(4)
4
D. Fern´andez et al.: Kinematics of young stars
Sagittarius-Carina or −I arm and the Norma-Scutum or
−II arm), one local arm (Orion-Cygnus or 0 arm) and one
external arm (Perseus or +I arm). The Orion-Cygnus arm
seems to be a local spur (Bok 1958). Lin et al. (1969) pro-
posed a galactic system with 2 main spiral arms, where
the Norma-Scutum and the Perseus arms are two seg-
ments of the same arm. By taking into account the in-
terarm distance between the Sagittarius-Carina and the
Perseus arms, these authors deduced a pitch angle of −6◦.
But, as early as the mid-70s, Georgelin & Georgelin (1976)
proposed a 4-armed galactic system with a pitch angle of
−12◦ from a study of the spatial distribution of HII re-
gions. However, Bash (1981) examined this 4-armed model
and found that a 2-arm pattern predicts HII regions in the
same direction and with the same radial velocities as those
used by Georgelin & Georgelin (1976), provided that dis-
persion velocities were considered.
In some recent papers several authors have also called
this classical view in question. Vall´ee (1995) reviewed the
subject of the determination of the pitch angle and the
number of spiral arms and concluded that the Galaxy has
a pitch angle of i = −12±1◦ and that, taking into account
the observed interarm distance, it would be a system of 4
spiral arms. This is also the opinion expressed by Amaral
& L´epine (1997), who, fitting the galactic rotation curve
to a mass model of the Galaxy, found an autoconsistent
solution with a system of 2 + 4 spiral arms (2 arms for
2.8 < < 13 kpc and 4 arms for 6 < < 11 kpc, with
the Sun placed at ⊙ = 7.9 kpc) and a pitch angle of
i = −14◦. Englmaier & Gerhard (1999) used the COBE
NIR luminosity distribution and connected it with the
kinematic observations of HI and molecular gas in l-v di-
agrams. They found a 4-armed spiral pattern between the
corotation of the galactic bar and the solar circle. Drimmel
(2000) found that the galactic plane emission in the K
band is consistent with a 2-armed structure, whereas the
240 µm emission from dust is compatible with a 4-armed
pattern. In a recent work, L´epine et al. (2001) analyzed the
kinematics of a sample of Cepheid stars and found the best
fit for a model with a superposition of 2+4 spiral arms.
Contrary to the model by Amaral & L´epine (1997), L´epine
et al. allowed the phase of both spiral patterns to be inde-
pendent, deriving pitch angles of approximately −6◦ and
−12◦ for m = 2 and m = 4, respectively. They argued that
this spiral pattern is in good agreement with the l-v dia-
grams obtained from observational HII data, though they
admit that pure 2-armed model produces similar results.
In the visible spiral structure of the Galaxy derived by
L´epine et al. (see their Fig. 3) the Orion-Cygnus or local
arm is seen as a major structure with a small pitch angle.
In contrast, Olano (2001) proposed that the local arm is
an elongated structure of only 4 kpc of length and a pitch
angle of about −40◦ (in better agreement with the obser-
vational determinations of the inclination of the local arm
found in the literature) formed from a supercloud about
100 Myr ago, when it entered into a major spiral arm.
In the case of the galactocentric distance of the Sun
and the circular velocity at the Sun's position there are
also some inconsistencies among the different values found
in the literature. In 1986, the IAU adopted the values
⊙ = 8.5 kpc and Θ(⊙) = 220 km s−1 (Kerr &
Lynden-Bell 1986). Recently, several authors have found
values of nearly 7.5 kpc for ⊙ (Racine & Harris 1989;
Maciel 1993). A complete review was done by Reid (1993),
who concluded that the most suitable value seems to be
⊙ = 8.0 ± 0.5 kpc. In kinematic studies there are seri-
ous discrepancies between different authors. Metzger et al.
(1998) found ⊙ = 7.7±0.3 kpc and Θ(⊙) = 237±12 km
s−1, whereas Feast et al. (1998) found ⊙ = 8.5 ± 0.3 kpc
(both from radial velocities of Cepheid stars). Glushkova
et al. (1998) found ⊙ = 7.3 ± 0.3 kpc from a com-
bined sample including open clusters, red supergiants and
Cepheids. Olling & Merrifield (1998) calculated mass mod-
els for the Galaxy and concluded that a consistent picture
only emerges when considering ⊙ = 7.1 ± 0.4 kpc and
Θ(⊙) = 184 ± 8 km s−1. This value for the galactocen-
tric distance of the Sun is in very good agreement with
the only direct distance determination (⊙ = 7.2 ± 0.7
kpc), which was made employing proper motions of H2O
masers (Reid 1993).
In the view of all that, we decided to derive the kine-
matic parameters of our model from different combina-
tions of the free parameters involved. On the one hand,
concerning spiral structure, two models of the Galaxy were
considered: a first model with m = 2 and i = −6◦, and a
second one with m = 4 and i = −14◦. Both models are
consistent with an interarm distance of about 2.5-3 kpc,
depending on the adopted value for the distance from the
Sun to the galactic center. On the other hand, concerning
the galactocentric distance of the Sun and the circular ve-
locity at the Sun's position, two different cases were also
taken into account: a first one with ⊙ = 8.5 kpc and
Θ(⊙) = 220 km s−1, and a second one with ⊙ = 7.1
kpc and Θ(⊙) = 184 km s−1. In both cases, the angular
rotation velocity at the Sun's position is Ω⊙ = 25.9 km
s−1 kpc−1.
4. Resolution of the condition equations
We determined the kinematic parameters of the galactic
model via least squares fit from the equations:
10
vr =
vl = k R µl cos b =
vb = k R µb =
10
Xj=1
Xj=1
Xj=1
10
ajf r
j (R, l, b)
ajf l
j(R, l, b)
ajf b
j (R, l, b)
(8)
where vr is the radial velocity of the star in km s−1, k
= 4.741 km yr (s pc ′′)−1, R is the heliocentric distance
of the star in pc, µl and µb are the proper motion in
galactic longitude and latitude of the star in ′′ yr−1 and
b the galactic latitude. To derive the parameters aj, an
D. Fern´andez et al.: Kinematics of young stars
5
iterative scheme extensively explained in Fern´andez (1998)
was followed.
The weight system was chosen as:
pk =
1
σ2
k,obs + σ2
k,cos
(9)
where σobs are the individual observational errors in each
velocity component of the star, calculated by taking into
account the correlations between the different variables
provided by the Hipparcos Catalogue, and σcos is the pro-
jection of the cosmic velocity dispersion ellipsoid in the
direction of the velocity component considered (see Paper
I).
To check the quality of the least squares fits we con-
sidered the χ2 statistics for N − M degrees of freeedom,
defined as:
χ2 =
N
Xk=1
[yk − y(xk; a1, ..., a10)]2
σ2
k,obs + σ2
k,obs
(10)
where xk are the independent data (sky coordinates and
distances), yk the dependent data (radial and tangential
velocity components), N the number of equations and M
number of parameters to be fitted. According to Press
et al. (1992), if the uncertainties (cosmic dispersion and
observational errors) are well-estimated, the value of χ2
for a moderately good fit would be χ2 ≈ N − M , with an
uncertainty of p2 (N − M ).
As we did in Paper I, to eliminate the possible
outliers present in the sample due to both the exis-
tence of high residual velocity stars (Royer 1997) or
stars with unknown large observational errors, we re-
jected those equations with a residual larger than 3 times
the root mean square residual of the fit (computed as
p[yk − y(xk; a1, ..., a10)]2/N ) and recomputed a new set
of parameters.
5. Test of robustness
The spiral arm potential is expected to contribute between
5-10% to the whole galactic gravitational field in the so-
lar neighbourhood. This small contribution, and the large
observational errors and constraints present on the spatial
and kinematic parameters of distant stars, make it very
difficult to quantify the kinematic perturbations induced
by the spiral potential. Then, the results found in the liter-
ature have been characterised by significant uncertainties
and discrepancies. A good example of this is the contradic-
tory results obtained for the phase of the spiral structure
at the Sun's position (see Sect. 7.2): between arms or near
an arm? Interesting questions to answer after the release
of the Hipparcos data are:
-- How does the unprecedented astrometric precision pro-
vided by Hipparcos help to diminish such uncertain-
ties? Are they small enough to measure the spiral arm
kinematic parameters to any useful degree of accuracy?
-- Can we quantify the biases induced by our observa-
tional errors and constraints?
Table 1. Number of stars with distance and proper mo-
tions (in brackets those stars with also radial velocity)
in several distance intervals. Distances for Cepheids com-
puted from Luri's (2000) PL relation.
sample of O and B stars
0.1 < R < 2 kpc
0.6 < R < 2 kpc
3418 (1903)
448 (307)
sample of Cepheid stars
0.1 < R < 4 kpc
0.6 < R < 2 kpc
0.6 < R < 4 kpc
119 (111)
103 (95)
164 (145)
-- How can the correlations among variables present in
the condition equations affect the kinematic parame-
ters?
With regard to our stellar data, as seen in Sect. 2, both
O and B star and Cepheid samples suffer from different
observational limitations. Although the O and B star sam-
ple is large in number, it is limited in distance to no more
than 1.5-2 kpc from the Sun (see Fig. 1 and Table 1). In
contrast, the Cepheid sample reaches distances of up to
about 4 kpc, but the number of stars with reliable data
still remains very small (Fig. 2 and Table 1).
Numerical simulations were performed to answer the
above questions, that is, to evaluate and quantify all the
uncertainties and biases involved in our resolution pro-
cess. In Appendix B we present the detailed procedure
followed to build simulated samples as similar as possible
to the real data for both O and B stars and Cepheids.
This Appendix also contains an exhaustive analysis of the
full set of cases which were simulated. Next we present the
main conclusions arising from this work, which substan-
tially contribute to the analysis of the real data:
-- As a general trend, owing to the correlations between
variables, the biases and uncertainties in some kine-
matic parameters substantially vary when changing
the real values of ψ⊙ and Ωp.
-- The first- and second-order terms of the galactic rota-
tion curve are accurately derived, with an uncertainty
of less than 1.3 km s−1 kpc−1 (km s−1 kpc−2) in all
cases. Whereas a negative bias for both parameters is
detected for O and B stars (up to −0.7 km s−1 kpc−1
for ar and −1 km s−1 kpc−2 for br), it is negligible for
Cepheids.
-- For the spiral structure parameters, we obtain better
results from O and B stars than from Cepheids. This
is due to the larger number of O and B stars. For both
samples we found a clear dependence of some param-
eters with the assumed values for ψ⊙ and Ωp. We can
expect a bias in ψ⊙ up to ±20◦, and uncertainties of
10-20◦ for O and B stars and 30-60◦ for Cepheids. For
Πb and Θb we found biases up to ±2 km s−1 and uncer-
tainties of 1-2 km s−1. We did not find an important
bias for Ωp neither for O and B stars nor Cepheids,
though its standard deviation is large (4-10 km s−1
kpc−1 for O and B stars and 10-20 km s−1 kpc−1 for
6
D. Fern´andez et al.: Kinematics of young stars
Cepheids). These ranges of values have allowed us to
check how compatible are the results independently
obtained for O and B stars and Cepheids.
-- If we choose an incorrect set of free parameters (m,
i, ⊙, Θ(⊙)), results change, but not to a large ex-
tent. This change is similar to the dispersion obtained
when solving the condition equations from the differ-
ent simulated samples (see Appendix B for details).
Therefore, it will be difficult to decide between 2- and
4-armed models.
From these results, in Appendix B we conclude that,
with the present available observational data, we are able
to determine the kinematic parameters of the galactic
model proposed in this paper, though we are not able to
decide between a 2- or 4-armed Galaxy.
6. Results
In this section we present the results obtained from our
working samples. To avoid those stars belonging to the
Gould Belt, which can produce important deviations in
our results (see Paper I), we always considered stars fur-
ther than 0.6 kpc from the Sun. Following the results ob-
tained in Appendix B, the working distance intervals are
0.6 < R < 2 kpc for O and B stars and 0.6 < R < 4 kpc
for Cepheids.
A first set of results is presented in Table 2, consider-
ing a classical view of our galaxy (Lin et al. 1969; Kerr &
Lynden-Bell 1986): we supposed a differential galactic ro-
tation (ar, br) and a system with 2 spiral arms and a pitch
angle of −6◦, the Sun placed at a galactocentric distance
of 8.5 kpc a the circular velocity at the Sun's position of
220 km s−1.
Although we do not only present solutions from radial
velocity or proper motion data, we would like to remark
that the parameters obtained when solving these cases
(and comparing with the combined solutions presented in
Table 2) are compatible between themselves within the
errors bars. This is true for the sample of O and B stars
and also in the case of the sample of Cepheids (for both
short and large cosmic scales). Nevertheless, as we deter-
mined in Paper I, for this kind of kinematic analysis it is
advisable to solve a combined resolution, in order to mini-
mize the influence of the correlations present between the
different parameters. This is especially important in the
present case, since some correlations reach a large value
(up to 0.8, though in the majority of cases they do not
exceed 0.3). In Appendix B we check that these correla-
tions do not impede our obtaining reliable results from
our stellar samples.
If we compare the results obtained from O and B stars
and Cepheids, we observe several discrepancies: differences
up to 3.5 km s−1 kpc−1 in the A Oort constant (derived
from ar) and up to 125◦ for the phase of the spiral struc-
ture at the Sun's position. We must keep in mind that,
whereas this parameter (like Ωp and cor) is independent
of the sample used, Πb, Θb and f⊙ depend on the cosmic
Table 2. Resolution for the samples of O and B stars
and Cepheids (SCS: short cosmic scale; LCS: large cos-
mic scale). Units: U⊙, V⊙, W⊙, Πb, Θb and σ in km
s−1; ar in km s−1 kpc−1; br in km s−1 kpc−2; ψ⊙ in
degrees. χ2/N is the value of χ2 divided by the num-
ber of equations minus the degrees of freedom. A cos-
mic dispersion of (σU , σV , σW ) = (8, 8, 5) km s−1 and
(σU , σV , σW ) = (13, 13, 6) km s−1 was used for O and
B stars and Cepheids, respectively.
O and B stars
0.6 < R < 2 kpc
Cepheid stars
0.6 < R < 4 kpc
8.8 ± 0.7
12.4 ± 1.0
8.4 ± 0.5
−1.3 ± 1.0
−0.8 ± 1.5
45. ± 52.
1.6 ± 1.1
2.1 ± 1.5
0.42 ± 0.10
12.11
2.07
SCS distances
6.5 ± 1.2
10.4 ± 1.9
5.7 ± 0.7
−8.1 ± 1.2
1.8 ± 0.8
282. ± 20.
0.1 ± 1.6
4.9 ± 1.7
0.96 ± 0.01
11.57
0.84
LCS distances
8.3 ± 1.2
9.3 ± 2.1
7.0 ± 0.8
−4.5 ± 1.1
−0.1 ± 0.8
306. ± 24.
0.7 ± 1.4
4.8 ± 1.7
0.96 ± 0.01
11.77
0.84
U⊙
V⊙
W⊙
ar
br
ψ⊙
Πb
Θb
f⊙
σ
χ2/N
dispersion of the sample (thus, they do not have the same
value for O and B stars and Cepheids).
As in Paper I, we found values of χ2/N ≈ 2 for O and
B stars. We think that the difference from the expected
value (χ2/N ≈ 1) could be due to an underestimation
of the errors in the photometric distances and/or radial
velocities for far stars. In the case of Cepheids we found
values which agree with the expected χ2/N ≈ 1.
When studying the residuals of the equations we re-
alized that radial velocity equations have a non-null av-
erage residual of about −(3-4) km s−1 for both O and
B stars and Cepheids. In the case of O and B stars (see
Fig. 3) we can observe the peculiar motion of some OB
associations. The kinematics of those associations located
near the Sagittarius arm (e.g. Sgr OB1, Ser OB1 and Sct
OB2, placed at 0◦ <∼ l <∼ 30◦) were studied by Mel'nik
et al. (1998), who found that their motions are in gen-
eral agreement with what would be expected according to
Lin's theory if these stars are located within the corota-
tion radius. But we found these stars have a large residual,
even taking into account in our model the galactic spiral
structure, possibly due to the fact that they are still re-
flecting a motion peculiar to their birthplaces owing to
their youth. Another region with a high residual is that
located in the direction of the open clusters h and χ Per
(NGC 869 and 884), at l ≈ 135◦, where we observe a
group of stars (distributed in an area of 10◦ in the sky)
with a high residual directed towards the Sun's position.
Thus, our results indicate that these stars do not fit the
systematic velocity field defined by the whole sample. In
this region of the sky the associations Per OB1 and Cas
OB6 are to be found. The former is generally considered
to include h and χ Per and surrounding O and B stars.
D. Fern´andez et al.: Kinematics of young stars
7
However, the stars in Fig. 3 have distances 1.1 < R < 1.4
kpc, whereas the open clusters are located further on, at
R ≈ 2-2.5 kpc (Schild 1967). In our initial sample (see
Sect. 2), we have in this region a group of stars with dis-
tances 2.0 < R < 2.6 kpc, most of them belonging to Per
OB1 according to the membership list done by Garmany &
Stencel (1992). But we did not use these stars in our kine-
matic analysis, due to their large distance errors. Garmany
& Stencel (1992) questioned if h and χ Per really belong
to Per OB1 due to the fact that B main sequence stars
do not define a ZAMS at the distance of the open clus-
ters. These authors derived a distance modulus of 11.8
(R = 2.3 kpc) for Per OB1 and 11.9 (R = 2.4 kpc) for
Cas OB6. More recently, Mel'nik & Efremov (1995) stud-
ied the spatial distribution of O and B stars within 3 kpc
from the Sun and derived a new partition into OB associ-
ations using Battinelli's (1991) modification of the cluster
analysis method. They found that Per OB1 split into four
groups, two at R ≈ 1.6 kpc and another two at R ≈ 1.9
kpc. Two of them are reliable at a 90% confidence level
(one at 1.6 kpc and another at 1.9 kpc), with a mean he-
liocentric radial velocity of −21.6 km s−1 and −41.4 km
s−1, respectively. The first association found by Mel'nik &
Efremov might correspond to the group of stars we have
at 1.1 < R < 1.4 kpc. In view of all that, the physical
explanation of the large negative radial residuals present
in this region, even considering the contribution from the
spiral arm kinematics, is still unresolved.
If we exclude those stars belonging to both regions
from our calculations, the average residual in radial ve-
locity decreases to −(1-2) km s−1. For Cepheids, as men-
tioned above, the residuals have the same trend and reach
up to −(3-4) km s−1, though in this case we cannot clearly
identify groups of stars with a common motion. Therefore,
the mean negative residual motion observed does not seem
to be explained as the peculiar motion of some groups
of stars. Moreover, we found that, for both samples, this
residual motion apparently does not depend on the galatic
longitude of the stars.
Taking into account that Cepheids, with a large mean
distance, show larger residuals than O and B stars, we
can suppose that this systematic trend could be explained
through the addition of a K-term in Eqs. (3), as developed
in Paper I. Then, the following terms can be added to Eqs.
(8):
11 = R cos2 b
f r
f l
11 = 0
f b
11 = −R sin b cos b
(11)
which allow us to derive the parameter a11 = K. Table
3 shows the results obtained in this way, and considering
different sets of free parameters (as proposed in Sect. 3.2).
The correlation coefficient between K and the other pa-
rameters is always low, being smaller than 0.4 for O and
B stars and 0.2 for Cepheids. As stated in Appendix B, a
change in the value of the free parameters (m, i, ⊙ and
Θ(⊙)) does not significantly alter the kinematic param-
Fig. 3. Residual heliocentric space velocity vectors pro-
jected on the galactic plane for O and B stars. GR stands
for galactic rotation direction and GC for galactic center
direction.
eters derived. In Table 3 we also show the results when
rejecting those stars belonging to the problematic regions
mentioned above in the case of O and B stars. We observe
that there is not a great difference in the results between
these cases, which are always compatible within the er-
rors. So, we conclude that the observed peculiar motion
of some OB associations present in our sample does not
alter the main parameters of our model. In view of that,
in the next section we discuss the results obtained when
considering all the stars, without rejecting associations.
7. Discussion
7.1. Galactic rotation curve
A difference in the A Oort constant of approximately 3
km s−1 kpc−1 between solutions for O and B stars and
Cepheids with short cosmic scale distances is obtained,
whereas large cosmic scale provides an intermediate value:
AOB ≈ 13.7-13.8 km s−1 kpc−1
ACep
Short ≈ 16.6-16.9 km s−1 kpc−1
ACep
Large ≈ 14.9-15.1 km s−1 kpc−1
(12)
As can be seen in Appendix B, these differences cannot be
explained by the observational constraints present in both
samples. A similar discrepancy was found by Frink et al.
(1996), who derived A = 14.0 ± 1.2 km s−1 kpc−1 from a
sample of O and B stars, and A = 15.8±1.6 km s−1 kpc−1
from a sample of Cepheids (in both cases the authors only
considered those stars with a heliocentric distance of less
than 1 kpc). Nonetheless, our values for Cepheids do not
reach the higher values obtained by Glushkova et al. (1998;
8
D. Fern´andez et al.: Kinematics of young stars
Table 3. Resolution by least squares fit considering differ-
ent values of the imposed galactic parameters m, i, ⊙ and
Θ(⊙). Case A: m = 2, i = −6◦, ⊙ = 8.5 kpc, Θ(⊙) =
220 km s−1; Case D: m = 4, i = −14◦, ⊙ = 7.1 kpc,
Θ(⊙) = 184 km s−1. Units: U⊙, V⊙, W⊙, Πb, Θb and
σ in km s−1; ar, K, A, B and Ωp in km s−1 kpc−1; br in
km s−1 kpc−2; ψ⊙ in degrees; ∆cor = ⊙ − cor in kpc.
χ2/N is the value of χ2 divided by the number of equa-
tions minus the degrees of freedom. G1 and G2 stand for
those groups of stars located in 0 < l < 50◦, 1 < R < 2
kpc (29 stars) and 130 < l < 140◦, 1 < R < 2 kpc (14
stars), respectively.
O and B stars with 0.6 < R < 2 kpc
Case A
Case D
Case D
2.3 ± 1.5
0.5 ± 1.3
9.2 ± 0.7
12.7 ± 1.1
8.3 ± 0.5
9.2 ± 0.8
13.2 ± 1.0
8.3 ± 0.5
10.0 ± 0.7
12.7 ± 1.0
8.3 ± 0.5
Case D
excluding G1 excluding G2
9.6 ± 0.8
U⊙
12.3 ± 1.1
V⊙
8.4 ± 0.5
W⊙
−1.7 ± 1.0 −1.5 ± 1.0 −2.3 ± 1.2 −0.9 ± 1.0
ar
−0.4 ± 1.5
0.0 ± 1.4
br
−3.2 ± 0.7 −2.8 ± 0.7 −1.4 ± 0.7 −2.4 ± 0.7
K
8. ± 52.
ψ⊙
2.8 ± 1.1
Πb
1.8 ± 1.3
Θb
0.19 ± 0.77
f⊙
A
13.4 ± 0.5
B −12.7 ± 0.8 −13.2 ± 0.7 −14.1 ± 0.8 −13.0 ± 0.7
Ωp
36. ± 8.
2.0 ± 1.2
11.85
1.89
329. ± 47.
2.6 ± 1.0
2.1 ± 1.3
0.23 ± 0.57 0.35 ± 0.25
13.7 ± 0.5
315. ± 32.
3.2 ± 1.1
3.2 ± 1.5
0.40 ± 0.10
14.1 ± 0.6
33. ± 6.
1.5 ± 0.9
11.88
1.90
32. ± 3.
1.1 ± 0.6
11.74
1.90
45. ± 14.
3.6 ± 1.6
11.85
1.89
20. ± 53.
3.1 ± 1.1
2.0 ± 1.3
13.8 ± 0.5
∆cor
σ
χ2/N
Cepheid stars with 0.6 < R < 4 kpc
SCS distances
LCS distances
1.6 ± 0.8
0.5 ± 0.8
Case A
8.4 ± 1.2
9.1 ± 2.1
7.0 ± 0.8
Case A
6.5 ± 1.2
10.5 ± 1.9
5.7 ± 0.7
Case D
6.5 ± 1.2
11.0 ± 2.1
5.7 ± 0.8
Case D
8.2 ± 1.2
U⊙
10.1 ± 2.3
V⊙
7.0 ± 0.8
W⊙
−7.9 ± 1.2 −7.4 ± 1.2 −4.4 ± 1.1 −3.9 ± 1.1
ar
0.1 ± 0.8 −0.9 ± 0.8
br
−0.8 ± 0.5 −1.0 ± 0.5 −1.2 ± 0.5 −1.2 ± 0.5
K
321. ± 43.
286. ± 21.
ψ⊙
0.4 ± 1.2
Πb
2.7 ± 1.7
Θb
0.96 ± 0.01
f⊙
A
14.9 ± 0.6
B −13.7 ± 0.4 −13.2 ± 0.4 −13.0 ± 0.4 −12.5 ± 0.4
Ωp
27. ± 2.
0.2 ± 0.6
11.90
0.86
284. ± 39.
0.0 ± 1.6 −1.4 ± 1.6
2.4 ± 1.7
4.8 ± 1.7
0.96 ± 0.01 0.95 ± 0.01
16.6 ± 0.6
23. ± 4.
0.0 ± 1.0 −0.7 ± 1.2
11.66
0.85
310. ± 21.
0.9 ± 1.3
5.6 ± 1.7
0.96 ± 0.01
15.1 ± 0.6
28. ± 3.
0.5 ± 0.8
11.70
0.83
11.56
0.84
16.9 ± 0.6
26. ± 3.
σ
∆cor
χ2/N
A = 19.5 ± 0.5 km s−1 kpc−1), Mishurov et al. (1997;
A = 20.9 ± 1.2 km s−1 kpc−1), Mishurov & Zenina (1999;
A = 18.8±1.3 km s−1 kpc−1) and L´epine et al. (2001; A =
17.5 ± 0.8 km s−1 kpc−1). Pont et al. (1994) and Metzger
et al. (1998), from radial velocities of Cepheid stars, found
values of A = 15.9 ± 0.3 km s−1 kpc−2 and A = 15.5 ±
0.4 km s−1 kpc−2, respectively. More recently, and using
proper motions and distance calibration from Hipparcos
data on Cepheid stars, Feast & Whitelock (1997) found
a value of A = 14.8 ± 0.8 km s−1 kpc−1. Using a similar
sample, Feast et al. (1998) found A = 15.1 ± 0.3 km s−1
kpc−1 from radial velocities. As we have mentioned in
Sect. 6, we found good coherence for radial velocity, proper
motion and combined solutions for both cosmic distance
scales.
An attempt to explain these discrepancies was made
by Olling & Merrifield (1998), who studied the variation
of the A and B Oort functions and found that they sig-
nificantly differ from the general ∼ Θ()/ dependence
expected for a nearly flat rotation curve. Inside the so-
lar circle, the value of A rises to 18 km s−1 kpc−1 for
∆ = − ⊙ ≈ −0.5 kpc, disminishes to 16 km s−1
kpc−1 for ∆ ≈ −1.2 kpc, and rises continuously for
∆ <∼ −1.5 kpc, to 19 km s−1 kpc−1 for ∆ ≈ −2 kpc
(see Fig. 3 in Olling & Merrifield 1998). Contrary to that,
beyond the solar circle A decreases to 10-12 km s−1 kpc−1,
maintaining this value in the interval 0 <∼ ∆ <∼ 2.5 kpc.
Although A is a local parameter, describing the local shape
of the rotation curve, Olling & Merrifield already pointed
out that the discrepancies in the results published in the
literature may be produced by their dependence on the
galactocentric distance. Our O and B stars are distributed
along all the galactic longitudes, whereas the Cepheids are
predominantly concentrated inside the solar circle, with
a peak in the spatial distribution for ∆ ≈ −0.6 kpc
corresponding to the Sagittarius-Carina arm (see Figs. 1
and 2). For O and B stars we found a classical value of
A ≈ 14 km s−1 kpc−1 (a value between 10-12 and 16-18
km s−1 kpc−1), whereas for Cepheids a value A ≈ 15-17
km s−1 kpc−1 (depending on the PL relation considered)
was derived, in agreement with Olling & Merrifield's as-
sumptions.
From the results obtained in Appendix B, we would ex-
pect uncertainties in the second-order term of the rotation
curve of ≈ 1.0 km s−1 kpc−2 for O and B stars and ≈ 0.5
km s−1 kpc−2 in the case of Cepheids. Taking this and the
results in Table 3 into account, we can state that br does
not differ from a null value more than 2 km s−1 kpc−2.
Pont et al. (1994) found a value br = −1.7 ± 0.2 km s−1
kpc−2, whereas Feast et al. (1998) found br = −1.6 ± 0.2
km s−1 kpc−2, both using radial velocities of Cepheid stars
(the latter with a Hipparcos distance calibration). A large
positive value was found by L´epine et al. (2001), who de-
rived br = 5.0 ± 1.0 km s−1 kpc−2 from their sample of
Cepheid stars. As we will see in Sect. 7.2, these authors
also found a large value for the amplitude of the velocity
component in the galactic rotation direction due to the
spiral potential (Θb). Without specific simulations, con-
sidering both their observational data and resolution pro-
cedure, it is difficult to guess how the correlations between
both parameters can affect their determination.
−4.0
−2.0
)
c
p
k
(
X
0.0
2.0
4.0
−4.0
−2.0
0.0
Y (kpc)
2.0
4.0
Fig. 4. Star distribution in the X-Y galactic plane for O
and B stars (filled circles) with 0.6 < R < 2 kpc and
Cepheids (empty circles) with 0.6 < R < 4 kpc (short
cosmic scale distances). Spiral arms were drawn consider-
ing ψ⊙ = 330◦. Dotted lines show the center of the spiral
arms (ψ = 0◦) and dashed lines draw their approximate
edges (ψ = ±90◦). Looking at this figure we must have in
mind the possible drift between the position of the optical
tracers in the spiral arms (i.e., our young stars) and the
minimum of the spiral potential (ψ = 0◦).
7.2. Spiral structure
Appendix B demonstrates that the available observational
data allow the characterization of the galactic spiral struc-
ture, though the biases and uncertainties on the parame-
ters have to be taken into account in the interpretation of
the results. Furthermore, we realized that it is very diffi-
cult to establish the correct number of spiral arms of the
Galaxy.
A first remarkable result is the fairly good coherence
obtained for the phase of the spiral structure at the Sun's
position ψ⊙ when using different free parameters (cases
A, D) or different samples compared to the great discrep-
ancies found in the literature (see below). We take into
consideration that Πb, Θb and f⊙ depend on the cosmic
dispersion and so they do not have the same value for O
and B stars and Cepheids.
From Fig. 2, and assuming that Cepheids fairly trace
the center of the Sagittarius-Carina arm, we can infer a
value of ψ⊙ ≈ 250◦ (the center of the inner visible spiral
arm -- the Sagittarius-Carina arm -- at about 1 kpc from
the Sun), depending on the exact value of the interarm
distance. Nevertheless, the density wave theory predicts
that the center of the visible arm (traced by its young
stars) does not coincide with the position of the spiral
potential minimum (Roberts 1969, 1970). We must bear in
D. Fern´andez et al.: Kinematics of young stars
9
mind that the kinematically derived value for ψ⊙ informs
us about the position of the Sun with respect this potential
minimum, not with respect the visible arm. In Table 3 we
found values inside the interval:
ψ⊙ ≈ 284-20◦
(13)
We take into consideration that in the minimum of the spi-
ral potential (near the center of an arm) ψ = 0◦, whereas
in the inner edge of the arm ψ ≈ 90◦ and in the outer
edge ψ ≈ −90◦ = 270◦. In the case of the phase of the
spiral structure at the Sun's position we expect a bias
of ∆ψ⊙ = ψobtained
⊙ ≈ 20◦ for O and B stars
and ∆ψ⊙ ≈ 0◦ for Cepheids, with uncertainties of ≈ 25◦
and ≈ 75◦, respectively (see Appendix B). These high
uncertainties can explain the range of values obtained.
According to the value found for ψ⊙, the Sun is located
between the center and the outer edge of an arm, nearer
to the former (see Fig. 4).
− ψreal
⊙
Our result stands in apparent contradiction to the clas-
sical mapping of the spiral structure tracers (Schmidt-
Kaler 1975; Elmegreen 1985), which locates the Sun in a
middle position between the inner arm (Sagittarius-Carina
arm) and the outer one (Perseus arm), that is, ψ⊙ ≈ 180◦.
The local arm (Orion-Cygnus arm) is normally considered
as a local spur rather than a real arm. In our model, as
we supposed an interarm distance of approximately 3 kpc,
the local arm is also considered a local spur (see Sect. 3.2).
But the value we obtained for ψ⊙ differs significantly from
180◦. If we consider a difference of 30-100◦ between the op-
tical tracers and the potential minimum of the spiral arm,
we notice that our ψ⊙ value is in good agreement with
the picture of the spiral structure that emerges from the
spatial distribution of stars in Fig. 4, with the inner spi-
ral arm at approximately 1 kpc from the Sun. A similar
result was obtained by Mel'nik et al. (1998), who found
that nearly 70% of the stars in the OB associations of the
Sagittarius-Carina arm have a residual motion (after cor-
recting their heliocentric velocities for solar motion and
galactic rotation) in the direction opposite to the galac-
tic rotation, as one would expect for those stars between
the inner edge and the center of the arm. Then, they also
found a shift between the optical position of the visible
arm (traced by young stars) and the minimum of the spi-
ral potential. Therefore, from our results we conclude that
the Sun is placed relatively near the potential minimum
of the Sagittarius-Carina arm, and the Perseus arm is lo-
cated far away, at about 2.5 kpc from the Sun.
Our range of values for ψ⊙ includes that obtained
by Cr´ez´e & Mennessier (1973) from a sample of O-B3
stars, ψ⊙ = 352 ± 30◦. Later, Mennessier & Cr´ez´e (1975)
found from a sample of O-B3 stars a value of ψ⊙ ≈ 90◦,
which locates the Sun near the inner edge of an arm.
Other authors obtained contradictory results. G´omez &
Mennessier (1977) found the Sun's position near the edge
of an arm from several samples of stars from FK4 and
FK4 Supplement Catalogues. Byl & Ovenden (1978) and
Comer´on & Torra (1991) obtained values of ψ⊙ = 165±1◦
and ψ⊙ = 135 ± 18◦ respectively, from samples of O and
10
D. Fern´andez et al.: Kinematics of young stars
B stars. Mishurov et al. (1997) found ψ⊙ = 290 ± 16◦
from a sample of Cepheid stars with radial velocities.
Mishurov & Zenina (1999) found ψ⊙ = 322 ± 9◦ (sup-
posing m = 2 and ⊙ = 7.5 kpc), from the same sample
of Cepheid stars, but also including the Hipparcos proper
motions. When these authors supposed m = 4 they found
ψ⊙ = 340 ± 9◦. More recently, Rastorguev et al. (2001)
found ψ⊙ = 274 ± 22◦ from a sample of 55 open clus-
ters younger than 40 Myr and 67 Cepheids with periods
smaller than 9 days, all of them at R < 4 kpc. We can see
that there is poor agreement between the results published
in the literature, though the last ones (the only ones with
Hipparcos data) are all included in our range of possible
values for ψ⊙.
The velocity amplitudes due to the spiral perturbation
obtained are less than 4 km s−1 for O and B stars (Πb ≈ 3
km s−1, Θb ≈ 2-3 km s−1) and 6 km s−1 for Cepheid
stars (Πb ≈ −1-1 km s−1, Θb ≈ 2-6 km s−1). The dif-
ference between results for O and B stars and Cepheids
is explained by Lin's theory as a result of their slight de-
pendence on the galactocentric distance and the differ-
ent cosmic dispersion for both samples. Mishurov et al.
(1997) found Πb = 6.3 ± 2.4 km s−1 and Θb = 4.4 ± 2.4
km s−1 from their radial velocity data of Cepheid stars
within 4 kpc from the Sun, supposing a 2-armed spiral
pattern. Mel'nik et al. (1999) found Πb = 6.4 ± 1.2 km
s−1 and Θb = 2.4 ± 1.2 km s−1 from Cepheids within
3 kpc from the Sun. Mishurov & Zenina (1999) found
Πb = 3.3 ± 1.6 km s−1 and Θb = 7.9 ± 2.0 km s−1 for
m = 2 and Πb = 3.5 ± 1.7 km s−1 and Θb = 7.5 ± 1.8 km
s−1 for m = 4. From the same sample, L´epine et al. (2001)
found Πm=2
= 14.0 ± 3.0 km
s−1 and Πm=4
= 10.9 ± 2.9 km
s−1. Their 2+4-armed model yields large values for Θb,
implying a large value for the ratio between the spiral po-
tential and the axisymmetric galactic field, much greater
than that 5-10% normally accepted. According to the re-
sults obtained in Appendix B, the present observational
uncertainties, biases and the correlations involved in the
resolution procedure cannot completely explain the dis-
crepancies among the values found in the literature. They
can be attributed to our poor knowledge of the real wave
harmonic structure of the galactic spiral pattern or to the
approximation performed in the linear density-wave the-
ory.
= 0.4 ± 3.0 km s−1, Θm=2
= 0.8 ± 3.3 km s−1, Θm=4
b
b
b
b
The derived angular rotation velocity of the spiral pat-
tern is:
Ωp ≈ 30 km s−1 kpc−1
(14)
though there is a great dispersion around this value in our
results from different samples and cases (dispersion of 2-7
km s−1 kpc−1, but up to 15 km s−1 kpc−1 in one extreme
case). This is an expected dispersion, taking into account
the results obtained in Appendix B (standard deviation of
2-5 km s−1 kpc−1 for O and B stars, but up to 15 km s−1
kpc−1 for ψ⊙ ≈ 270◦; for Cepheids dispersions are higher,
of about 10-20 km s−1 kpc−1). Such a value places the
Sun very near the corotation circle and is not in agreement
with the classical Ωp ≈ 13.5 km s−1 kpc−1 proposed by
Lin et al. (1969). Nevertheless, other studies show results
similar to ours. Avedisova (1989) obtained Ωp = 26.8 ±
2 km s−1 kpc−1 from the spatial distribution of young
objects of different ages in the Sagittarius-Carina arm.
More recently, Amaral & L´epine (1997), working with a
selection of young members of the open cluster catalogue
by Mermilliod (1986), obtained a value of Ωp ≈ 20-22 km
s−1 kpc−1. Mishurov et al. (1997) and Mishurov & Zenina
(1999) used their sample of Cepheid stars and found Ωp =
28.1 ± 2.0 km s−1 kpc−1 and Ωp ≈ 27.7 km s−1 kpc−1,
respectively. L´epine et al. (2001) found Ωm=2
≈
26.5 km s−1 kpc−1 from their sample of Cepheid stars.
Rastorguev et al. (2001) thought that the evaluation of
Ωp from kinematical data alone cannot be resolved. But
our simulations (see Appendix B) seem to indicate that, at
least, the tendency to find high values of Ωp is confirmed,
though there is still an uncertainty of about 5-10 km s−1
kpc−1 in its value. These large values for Ωp can explain in
a natural way the presence of a gap in the galactic gaseous
disk (see Kerr 1969 and Burton 1976 for observational
evidences and L´epine et al. 2001 for simulated results),
since they placed the Sun near the corotation circle, where
the gas is pumped out under the influence of the spiral
potential.
≈ Ωm=4
p
p
7.3. The K-term
In relation to the K-term, we found good agreement be-
tween both O and B stars and Cepheids. A value of
K ≈ −(1-3) km s−1 kpc−1 is found in all cases, confirm-
ing an apparent compression of the solar neighbourhood
of up to 3-4 kpc. In our opinion, what most clearly proves
the existence of a non-null value of K is that it is inde-
pendently found from the samples of O and B stars and
Cepheids, which have a very different spatial distribution,
an independent distance estimation and a different way of
deriving their radial velocities. As we can see comparing
Tables 2 and 3, the inclusion of the K-term does not sub-
stantially modify the other model parameters derived by
least squares fit.
It is very difficult to find in the literature other esti-
mations of K at these large heliocentric distances, since
in the majority of cases authors consider an axisymmet-
ric rotation curve. But some authors have pointed out the
persistence of a residual in the radial velocity equations.
Comer´on & Torra (1994) found K = −1.9 ± 0.5 km s−1
kpc−1 from O-B5.5 stars and K = −1.3 ± 0.9 km s−1
kpc−1 from B6-A0 stars with R < 1.5 kpc. The radial ve-
locity residual for Cepheids was first recognised by Stibbs
(1956). Pont et al. (1994) studied three possible origins for
it: a statistical effect, an intrinsic effect in the measured
radial velocities for Cepheids (γ-velocities) and a real dy-
namical effect. They finally suggested that it could be due
to a non-axisymmetric motion produced by a central bar
of ≈ 5 kpc in extent. Metzger et al. (1998) found a residual
of ≈ −3 km s−1 for a sample of Cepheids when consid-
D. Fern´andez et al.: Kinematics of young stars
11
ering an axisymmetric galactic rotation. They concluded
that it might be due to the influence of spiral structure,
not included in their model. However, in the present work
we found a non-null K value even taking into account
spiral arm kinematics. An understanding of the physical
explanation of the K-term requires further study.
Appendix A: systematic velocity components in
the proposed galactic model
In this appendix we show the expressions of the velocity
components in the three systematic contributions consid-
ered in our galactic model: solar motion, differential galac-
tic rotation and spiral arm kinematics.
8. Conclusions
A.1. Solar motion
Hipparcos astrometric data were used to derive the spiral
structure of the Galaxy in the solar neighbourhood. We
considered two different samples of stars as tracers of this
structure. In the sample of O and B stars the astromet-
ric data were complemented with a careful compilation
of radial velocities and Stromgren photometry, providing
reliable distances and spatial velocities for these stars. In
the sample of Cepheid stars we used two period-luminosity
relations, one considering a short cosmic scale (Luri 2000)
and another one standing for a large cosmic scale (Feast
& Catchpole 1997).
A kinematic model of our galaxy that takes into ac-
count solar motion, differential galactic rotation, spiral
arm kinematics and a K-term was adopted. The model
parameters were derived via a classical weighted least
squares fit. The robustness of our results was checked by
means of careful insimulations.
A galactic rotation curve with a value of A Oort con-
stant of 13.7-13.8 km s−1 kpc−1 was found for the sam-
ple of O and B stars. In the case of Cepheid stars, we
found A = 14.9-16.9 km s−1 kpc−1, depending on the
case and the cosmic scale chosen. We confirmed the dis-
crepancies appearing in A when samples with a different
distance horizon were used (Olling & Merrifield 1998). For
both cosmic scales we found an acceptable coherence be-
tween radial velocity, proper motion and combined solu-
tions. Concerning the second-order term of the rotation
curve, we always found a low value, compatible with zero.
The study of the residuals for radial velocity data made
it evident that a K-term was needed, which was found to
be K = −(1.4-3.2) km s−1 kpc−1 for O and B stars and
K = −(0.8-1.2) km s−1 kpc−1 for Cepheids. Although
small, an apparent compression motion seems to exist in
the galactic local neighbourhood for heliocentric distances
up to 4 kpc, though its physical mechanism is still un-
known.
The phase of the spiral structure at the Sun's posi-
tion obtained (ψ⊙ = 284-20◦) places it between the center
and the outer edge of an arm. This is in good agreement
with the spatial distribution of Cepheid stars. The angu-
lar rotation velocity of the spiral structure was found to
be Ωp ≈ 30 km s−1 kpc−1, which places the Sun near the
corotation circle.
Acknowledgements. This work has been supported by the
CICYT under contracts ESP 97-1803 and AYA 2000-0937. DF
acknowledges the FRD grant from the Universitat de Barcelona
(Spain).
A star with galactic longitude l and galactic latitude b has
the following radial and tangential velocity components
owing to solar proper motion:
vr1 = −U⊙ cos l cos b − V⊙ sin l cos b − W⊙ sin b
vl1 = U⊙ sin l − V⊙ cos l
vb1 = U⊙ cos l sin b + V⊙ sin l sin b − W⊙ cos b
(A.1)
where U⊙, V⊙ and W⊙ are the components of the solar
motion in galactic coordinates.
A.2. Galactic rotation
We consider axisymmetric differential rotation of our
galaxy, with a rotation curve that can be developed in
the solar neighbourhood as:
Θ() ≈ Θ(⊙) +(cid:18) ∂Θ
∂(cid:19)⊙
∆ +
≡ Θ(⊙) + ar∆ + br∆2
1
2(cid:18) ∂2Θ
∂2(cid:19)⊙
∆2
(A.2)
where ∆ = −⊙ (⊙ is the galactocentric distance of
the Sun) and Θ(⊙) is the circular velocity at the Sun's
position. We note that ar allows us to calculate the A Oort
constant in the Sun's vicinity:
A =
1
2" Θ(⊙)
⊙
−(cid:18) ∂Θ
∂(cid:19)⊙# =
1
2(cid:20) Θ(⊙)
⊙
− ar(cid:21)
(A.3)
The radial and tangential velocity components of a star
in the galactic plane due to differential galactic rotation
are:
vr2 = Θ(⊙) [sin(l + θ) − sin l] cos b
+ar∆ sin(l + θ) cos b
+br∆2 sin(l + θ) cos b
vl2 = Θ(⊙) [cos(l + θ) − cos l]
+ar∆ cos(l + θ)
+br∆2 cos(l + θ)
vb2 = −Θ(⊙) [sin(l + θ) − sin l] sin b
−ar∆ sin(l + θ) sin b
−br∆2 sin(l + θ) sin b
(A.4)
where θ is the galactocentric longitude of the star.
12
D. Fern´andez et al.: Kinematics of young stars
A.3. Spiral structure kinematics
Lin's theory (Lin & Shu 1964; Lin et al. 1969; see also
Rohlfs 1977) assumes a spiral potential of the form:
Vb = A cos ψ
(A.5)
(A is the amplitude -- A < 0 -- and ψ is the phase of
the density wave) which disturbs the axially symmetric
gravitational potential. The shape of the spiral arms is
well represented by a logarithmic spiral:
q(, θ, t) = q eiψ(,θ,t)
(A.6)
The amplitude q is a slowly varying function of and the
phase of the spiral structure can be related to the phase
at the Sun's position by the following expression:
ψ(, θ, t) = ψ⊙(t) + m(Ωpt − θ) +
⇒ ψ(, θ, t = 0) = ψ⊙ + mθ +
m ln
⊙(t)
tan i
⇒
m ln
⊙
tan i
where Ωp is the angular rotation velocity of the spiral pat-
tern, m the number of spiral arms and i the pitch angle (for
trailing spiral arms, i < 0). The phase of the spiral struc-
ture at the Sun's position and the pitch angle can be de-
termined from optical and radio indicators. Nevertheless,
we point out that the maximum in the distribution of spi-
ral arm tracers (position of the observed spiral arms) may
be shifted in relation to the minimum in the perturbation
potential (defined as ψ = 0◦; see Roberts 1969).
The mean peculiar velocities due to the spiral arm per-
turbations on the velocity field are, in the gas approxima-
tion, the following:
and x is the stability Toomre's number (Toomre 1969)
defined as:
x =
k2a2
o
κ2
(A.12)
where ao is the velocity dispersion of the gas particles.
Since the velocity amplitudes Πb and Θb depend on this
velocity dispersion, we introduce a dimensionless param-
eter (f⊙) that relates the velocity amplitudes of the Sun
to those of the sample stars:
Πb⊙ =
Θb⊙ =
1 − ν2 + xstars
1 − ν2 + x⊙
1 − ν2 + xstars
1 − ν2 + x⊙
Πb ≡ f⊙Πb
Θb ≡ f⊙Θb
(A.13)
The so-called Lindblad resonances are defined as:
Ωp = Ω ±
κ
m
(A.7)
(A.14)
where the − sign corresponds to the inner resonance and
the + sign to the outer one. In the region between both
resonances (ν < 1), Θb is always positive and Πb has a
sign that depends on the sign of ν.
The amplitude of the spiral potential can be expressed
as:
A =
κΠb
1 − ν2 + x
k
ν
(A.15)
The maximum value of the radial force owing to this spiral
potential is:
F max
r1 = kA
(A.16)
whereas the radial force due to the axisymmetric field is:
Π1 =
Θ1 = −
kA
κ
1
2
ν
1 − ν2 + x
kA
1
Θ
1 − ν2 + x
cos ψ ≡ Πb cos ψ
sin ψ ≡ −Θb sin ψ
(A.8)
Fr0 =
dV0
d
≈ Ω2
⊙⊙
Π1 is positive towards the galactic anti-center and Θ1 is
positive towards the galactic rotation. In tightly wound
spirals (i.e., those with tan i ≪ 1) the amplitudes Πb and
Θb are slowly varying quantities with the galactocentric
distance. k is the radial wave number (for trailing spiral
arms, k < 0):
(A.17)
(A.18)
Therefore, the ratio between both quantities is:
fr =
F max
r1
Fr0
≈
κΠb
Ω2
⊙⊙
1 − ν2 + x
ν
The contributions in the radial and tangential velocity
components of a star due to the spiral arm perturbation
velocities Π1 and Θ1 are the following:
k =
d
d m ln
tan i ! =
⊙
m
tan i
(A.9)
vr3 = −Π1 cos(l + θ) cos b + Π1⊙ cos l cos b
ν is the dimensionless rotation frequency of the spiral
structure, expressed in terms of the epicyclic frequency
(κ):
ν =
m
κ (cid:18)Ωp −
Θ
(cid:19)
(A.10)
(notice that ν < 0 in the region with Ωp < Ω = Θ/, i.e.
inner to the corotation circle). Furthermore:
+Θ1 sin(l + θ) cos b − Θ1⊙ sin l cos b
= −Πb cos ψ⊙
ln
⊙
× cos"m θ −
−Πb sin ψ⊙ sin"m θ −
−Θb sin ψ⊙
tan i !# cos(l + θ) − f⊙ cos l! cos b
ln
⊙
tan i !# cos(l + θ) cos b
κ2 =
2Θ2
2 (cid:18)1 +
Θ
dΘ
d(cid:19)
(A.11)
× cos"m θ −
ln
⊙
tan i !# sin(l + θ) − f⊙ sin l! cos b
D. Fern´andez et al.: Kinematics of young stars
13
+Θb cos ψ⊙ sin"m θ −
ln
⊙
tan i !# sin(l + θ) cos b
vl3 = Π1 sin(l + θ) − Π1⊙ sin l +
Θ1 cos(l + θ) − Θ1⊙ cos l
= Πb cos ψ⊙
tan i !# sin(l + θ) − f⊙ sin l!
a3 = W⊙
a4 = Θ(⊙)
a5 = ar
a6 = br
a7 = Πb cos ψ⊙
a8 = Πb sin ψ⊙
a9 = Θb sin ψ⊙
a10 = Θb cos ψ⊙
(A.21)
ln
⊙
tan i !# sin(l + θ)
j (R, l, b) are functions of the heliocentric distance
and f i
and the galactic longitude and latitude:
ln
⊙
× cos"m θ −
+Πb sin ψ⊙ sin"m θ −
−Θb sin ψ⊙
ln
⊙
× cos"m θ −
+Θb cos ψ⊙ sin"m θ −
tan i !# cos(l + θ) − f⊙ cos l!
ln
⊙
tan i !# cos(l + θ)
f r
1 = − cos l cos b
f r
2 = − sin l cos b
f r
3 = − sin b
f r
4 = [sin(l + θ) − sin l] cos b
f r
5 = ∆ sin(l + θ) cos b
f r
6 = ∆2 sin(l + θ) cos b
vb3 = Π1 cos(l + θ) sin b − Π1⊙ cos l sin b
−Θ1 sin(l + θ) sin b + Θ1⊙ sin l sin b
= Πb cos ψ⊙
ln
⊙
× cos"m θ −
+Πb sin ψ⊙ sin"m θ −
+Θb sin ψ⊙
ln
⊙
× cos"m θ −
−Θb cos ψ⊙ sin"m θ −
tan i !# cos(l + θ − f⊙ cos l! sin b
ln
⊙
tan i !# cos(l + θ) sin b
tan i !# sin(l + θ) − f⊙ sin l! sin b
f r
ln
⊙
tan i !# sin(l + θ) sin b
(A.19)
A.4. Systematic velocity field in the proposed galactic
model
The systematic radial and tangential velocity components
of a star in our galactic model are given by:
10
ajf r
j (R, l, b)
ajf l
j(R, l, b)
vr = vr1 + vr2 + vr3 =
vl = vl1 + vl2 + vl3 =
vb = vb1 + vb2 + vb3 =
10
Xj=1
Xj=1
Xj=1
10
ajf b
j (R, l, b)
(A.20)
where the constants aj contain combinations of the kine-
matic parameters that we wish to determine:
f r
f r
7 = − cos"m θ −
8 = − sin"m θ −
9 = − cos"m θ −
10 = sin"m θ −
f r
ln
⊙
ln
⊙
tan i !# cos(l + θ) − f⊙ cos l! cos b
tan i !# cos(l + θ) cos b
tan i !# sin(l + θ) − f⊙ sin l! cos b
ln
⊙
ln
⊙
tan i !# sin(l + θ) cos b
f l
1 = sin l
f l
2 = − cos l
f l
3 = 0
f l
4 = cos(l + θ) − cos l
f l
5 = ∆ cos(l + θ)
6 = ∆2 cos(l + θ)
f l
f l
f l
ln
⊙
ln
⊙
7 = cos"m θ −
8 = sin"m θ −
9 = − cos"m θ −
10 = sin"m θ −
tan i !# sin(l + θ) − f⊙ sin l
tan i !# sin(l + θ)
tan i !# cos(l + θ) − f⊙ cos l
tan i !# cos(l + θ)
ln
⊙
ln
⊙
f l
f l
a1 = U⊙
a2 = V⊙
f b
1 = cos l sin b
f b
2 = sin l sin b
14
D. Fern´andez et al.: Kinematics of young stars
f b
3 = − cos b
f b
4 = − [sin(l + θ) − sin l] sin b
f b
5 = −∆ sin(l + θ) sin b
f b
6 = −∆2 sin(l + θ) sin b
f b
f b
f b
ln
⊙
ln
⊙
7 = cos"m θ −
8 = sin"m θ −
9 = cos"m θ −
10 = − sin"m θ −
tan i !# cos(l + θ) − f⊙ cos l! sin b
tan i !# cos(l + θ) sin b
tan i !# sin(l + θ) − f⊙ sin l! sin b
tan i !# sin(l + θ) sin b
ln
⊙
ln
⊙
f b
(A.22)
In our resolution procedure the parameters aj are com-
puted following an iterative scheme (extensively explained
in Fern´andez 1998) and, from these, the kinematic param-
eters U⊙, V⊙, W⊙, ar, br, ψ⊙, Πb, Θb and f⊙ are derived.
With regard to the free parameters of our model, dif-
ferent values for m, i, ⊙, Θ(⊙) were considered (see
Sect. 7).
Appendix B: simulations to check the kinematic
analysis
In Paper I we did numerical simulations in order to evalu-
ate the biases in the kinematic model parameters (in that
case, the Oort constants and the solar motion compo-
nents) induced by our observational constraints and er-
rors. In the present work, we have also generated sim-
ulated samples in the same way, though the significant
correlations detected between some parameters make it
advisable, in this case, to carry out a more detailed study.
In this section we present the process used to generate
the simulated samples (the same as in Paper I, except for
the change in the systematic contributions considered),
the results we obtained and, finally, the quantification of
the biases present in our real samples.
B.1. Process used to generate the simulated samples
To take into account the irregular spatial distribution of
our stars and their observational errors, parameters de-
scribing the position of each simulated pseudo-star were
generated as follow:
-- From each real star we generated a pseudo-star that
has the same nominal position (R0, l, b) -- not affected
by errors -- as the real one.
-- We assumed that the angular coordinates (l, b) have
negligible observational errors.
-- The distance error of the pseudo-star has a distribution
law:
ε(R) = e− 1
σR (cid:1)2
2(cid:0) R−R0
(B.1)
where σR is the individual error in the photometric
distance of the real star (R0).
− 1
− 1
σV (cid:1)2
2(cid:0) V −V0
σU (cid:1)2
2(cid:0) U −U0
To generate the kinematic parameters we randomly as-
signed to each pseudo-star a velocity (U, V, W ) by assum-
ing a cosmic dispersion (σU , σV , σW ) and a Schwarzschild
distribution:
v(U, V, W ) = e− 1
ϕ′
where (U0, V0, W0) are the reflex of solar motion. These
components were transformed into radial velocities and
proper motions in galactic coordinates using the nominal
position of the pseudo-star (R0, l, b). The systematic mo-
tion due to galactic rotation and spiral arm kinematics
was added following Eqs. (A.4) and (A.19), obtaining the
components (vr0 , µl0, µb0) for each pseudo-star. Finally, in-
dividual observational errors were introduced by using the
error function:
σW (cid:1)2
2(cid:0) W −W0
(B.2)
ε(vr, µl, µb) = e
− 1
σvr (cid:1)2
2(cid:0) vr−vr0
− 1
σµl (cid:17)2
2(cid:16) µl −µl0
− 1
σµb (cid:17)2
2(cid:16) µb−µb0
(B.3)
where σvr , σµl and σµb are the observational errors of the
real star.
At the end of this process we had the following data
for each pseudo-star: galactic coordinates (R, l, b), veloc-
ity parameters (vr, µl, µb), errors in the velocity parame-
ters (σvr , σµl , σµb ) and error in the photometric distance
(σR). The simulated radial component of those pseudo-
stars generated from a real star without radial velocity
was not used, thus we imposed on the simulated sample
the same deficiency in radial velocity data that is present
in our real sample (see Sect. 2.2 and Appendix B in Paper
I for more details).
Following this scheme, several sets of 50 simulated
samples for both O and B stars and Cepheids were built.
A classical solar motion of (U, V, W ) = (9, 12, 7) km s−1
was considered, taking the dispersion velocity compo-
nents (σU , σV , σW ) = (8, 8, 5) km s−1 for O and B stars
(see Paper I) and (σU , σV , σW ) = (13, 13, 6) km s−1 for
Cepheids (Luri 2000). For the galactic rotation parame-
ters, we chose the values ar = −2.1 km s−1 kpc−1 and
br = 0.0 km s−1 kpc−2, which correspond to a linear rota-
tion curve with an A Oort constant of 14.0 km s−1 kpc−1.
On the other hand, for the spiral structure parameters
several sets of values were used for ψ⊙ (from ψ⊙ = 0◦ to
ψ⊙ = 315◦, in steps of 45◦) and Ωp (from Ωp = 10 km
s−1 kpc−1 to Ωp = 40 km s−1 kpc−1, in steps of 5 km s−1
kpc−1), whereas a fixed value of fr = 0.05 was considered
(Yuan 1969). From (σU , σV , σW ), ar, Ωp and fr, the values
of Πb, Θb and f⊙ were inferred for each set of samples.
56 sets of 50 samples for both O and B stars and
Cepheids were generated. Concerning the free parameters
of our model, in a first stage we adopted classical val-
ues (m = 2, i = −6◦, Lin et al. 1969; ⊙ = 8.5 kpc,
D. Fern´andez et al.: Kinematics of young stars
15
Table B.1. Parameters of the simulated samples.
U⊙
V⊙
W⊙
(σU , σV , σW )
ar
br
fr
ψ⊙
Ωp
Case A
Case B
Case C
Case D
9 km s−1
12 km s−1
7 km s−1
(8, 8, 5) km s−1 (O and B stars)
(13, 13, 6) km s−1 (Cepheid stars)
−2.1 km s−1 kpc−1
0.0 km s−1 kpc−2
0.05
from 0◦ to 360◦,
in steps of 45◦
from 10 km s−1 kpc−1 to 40 km s−1 kpc−1,
in steps of 5 km s−1 kpc−1
m = 2, i = −6◦,
⊙ = 8.5 kpc, Θ(⊙) = 220 km s−1
m = 2, i = −6◦,
⊙ = 7.1 kpc, Θ(⊙) = 184 km s−1
m = 4, i = −12◦,
⊙ = 8.5 kpc, Θ(⊙) = 220 km s−1
m = 4, i = −12◦,
⊙ = 7.1 kpc, Θ(⊙) = 184 km s−1
Θ(⊙) = 220 km s−1, Kerr & Lyndell-Bell 1986), though
we also tested cases with m = 4, i = −12◦ (Amaral &
L´epine 1997) and ⊙ = 7.1 kpc, Θ(⊙) = 184 km s−1
(Olling & Merrifield 1998). In Table B.1 we summarize all
the adopted kinematic parameters.
B.2. Results and discussion
B.2.1. Results for a 2-armed model of the Galaxy
A complete solution simultaneously taking into account
radial velocity and proper motion data was computed. Our
test showed that the number of Cepheids within 2 kpc
from the Sun prevents the obtainment of reliable results.
In Figs. B.1 and B.2 we show the results obtained for the
simulated samples of O and B stars (0.6 < R < 2 kpc) and
Cepheids (0.6 < R < 4 kpc) in case A (see Table B.1).
As a first conclusion, and confirming our suspicions, a
systematic trend with ψ⊙ and/or Ωp is observed in most
cases. This behaviour is produced by the correlations be-
tween some terms in the least squares fit, which depends
on the spatial distribution of each sample.
For solar motion a bias between −1.5 and 1.5 km s−1
(depending on ψ⊙ and Ωp) was found for U⊙ and V⊙, and
of only −0.3 km s−1 for W⊙. For both O and B stars and
Cepheids we found the bias on V⊙ and W⊙ to be indepen-
dent of Ωp, with a slight dependence on ψ⊙. For ψ⊙ = 270◦
and Ωp = 10 km s−1 kpc−1 we found a large negative bias
on V⊙, but with a great standard deviation. This occurred
in several samples (inside this set) with serious conver-
gence problems in the iteration procedure we use to solve
the least squares fit. Similar problems in other cases with
ψ⊙ = 270◦ will be found later. For O and B stars, the
standard deviations in the solar motion components are
≈ 0.6 km s−1 for U⊙ and V⊙ (except for ψ⊙ = 270◦), and
≈ 0.3 km s−1 for W⊙. On the other hand, in the case of
2.0
1.0
0.0
−1.0
−2.0
1.0
0.0
−1.0
−2.0
1.0
0.0
−1.0
−2.0
1.0
0.5
0.0
−0.5
)
1
−
s
m
k
(
Θ
U
∆
)
1
−
s
m
k
(
Θ
V
∆
)
1
−
s
m
k
(
Θ
W
∆
)
1
−
c
p
k
1
−
s
m
k
(
r
a
∆
−1.0
)
2
−
c
p
k
1
−
s
m
k
(
0.5
0.0
−0.5
r
b
∆
−1.0
40
20
0
−20
−40
1.0
0.0
−1.0
−2.0
1.0
0.0
−1.0
−2.0
0.2
0.0
−0.2
−0.4
10
0
−10
)
s
e
e
r
g
e
d
(
Θ
ψ
∆
)
1
−
s
m
k
(
b
Π
∆
)
1
−
s
m
k
(
b
Θ
∆
Θ
f
∆
)
1
−
c
p
k
1
−
s
m
k
(
p
Ω
∆
−20
)
1
−
s
m
k
(
Θ
U
σ
)
1
−
s
m
k
(
Θ
V
σ
)
1
−
s
m
k
(
Θ
W
σ
0
90
180
ψΘ (degrees)
270
360
)
1
−
c
p
k
1
−
s
m
k
(
r
a
σ
)
2
−
c
p
k
1
−
s
m
k
(
4.0
3.0
2.0
1.0
0.0
3.0
2.0
1.0
0.0
3.0
2.0
1.0
0.0
2.0
1.5
1.0
0.5
0.0
1.5
1.0
0.5
0
90
180
ψΘ (degrees)
270
360
0
90
180
ψΘ (degrees)
270
360
r
b
σ
0.0
0
90
180
ψΘ (degrees)
270
360
80
60
40
20
0
3.0
2.0
1.0
0.0
3.0
2.0
1.0
0.0
1.0
0.5
0.0
30
20
10
0
0
90
)
s
e
e
r
g
e
d
(
Θ
ψ
σ
)
1
−
s
m
k
(
b
Π
σ
)
1
−
s
m
k
(
b
Θ
σ
Θ
f
σ
)
1
−
c
p
k
1
−
s
m
k
(
p
Ω
σ
360
180
270
ψΘ (degrees)
360
0
90
180
270
ψΘ (degrees)
Fig. B.1. Bias (obtained value−simulated value; left) and
standard deviation (right) for each set of 50 samples in the
solar motion (top), galactic rotation (middle) and spiral
arm kinematic (bottom) parameters for O and B pseudo-
stars (Case A). Values of Ωp: 10 km s−1 kpc−1 (black
double solid line), 15 (black solid line), 20 (grey solid line),
25 (dotted line), 30 (dashed line), 35 (long dashed line)
and 40 (dot-dashed line).
16
)
1
−
s
m
k
(
Θ
U
∆
)
1
−
s
m
k
(
Θ
V
∆
)
1
−
s
m
k
(
Θ
W
∆
2.0
1.0
0.0
−1.0
−2.0
1.0
0.0
−1.0
−2.0
1.0
0.0
−1.0
−2.0
1.0
0.5
0.0
−0.5
)
1
−
c
p
k
1
−
s
m
k
(
r
a
∆
−1.0
)
2
−
c
p
k
1
−
s
m
k
(
0.5
0.0
−0.5
r
b
∆
−1.0
40
20
0
−20
−40
1.0
0.0
−1.0
−2.0
1.0
0.0
−1.0
−2.0
0.2
0.0
−0.2
−0.4
10
0
−10
)
s
e
e
r
g
e
d
(
Θ
ψ
∆
)
1
−
s
m
k
(
b
Π
∆
)
1
−
s
m
k
(
b
Θ
∆
Θ
f
∆
)
1
−
c
p
k
1
−
s
m
k
(
p
Ω
∆
−20
)
1
−
s
m
k
(
Θ
U
σ
)
1
−
s
m
k
(
Θ
V
σ
)
1
−
s
m
k
(
Θ
W
0
90
180
ψΘ (degrees)
270
360
σ
)
1
−
c
p
k
1
−
s
m
k
(
r
a
σ
)
2
−
c
p
k
1
−
s
m
k
(
4.0
3.0
2.0
1.0
0.0
3.0
2.0
1.0
0.0
3.0
2.0
1.0
0.0
2.0
1.5
1.0
0.5
0.0
1.5
1.0
0.5
0
90
180
ψΘ (degrees)
270
360
0
90
180
ψΘ (degrees)
270
360
r
b
σ
0.0
0
90
180
ψΘ (degrees)
270
360
80
60
40
20
0
3.0
2.0
1.0
0.0
3.0
2.0
1.0
0.0
1.0
0.5
0.0
30
20
10
0
0
90
)
s
e
e
r
g
e
d
(
Θ
ψ
σ
)
1
−
s
m
k
(
b
Π
σ
)
1
−
s
m
k
(
b
Θ
σ
Θ
f
σ
)
1
−
c
p
k
1
−
s
m
k
(
p
Ω
σ
360
180
270
ψΘ (degrees)
360
0
90
180
270
ψΘ (degrees)
D. Fern´andez et al.: Kinematics of young stars
Cepheids these values increased to ≈ 1.4-2.0 km s−1 and
≈ 0.8 km s−1, respectively.
The biases found in the first- and second-order terms
of the galactic rotation curve are negligible for Cepheids,
with a level fluctuation of ±0.3 km s−1 kpc−1 (or km s−1
kpc−2). In this case there is a standard deviation of 1.3
for ar and 0.5 for br. For O and B stars the biases clearly
depend on ψ⊙, varying from −0.7 to −0.5 km s−1 kpc−1
for ar, and from −1.0 to 0.1 km s−1 kpc−2 for br. The
standard deviations are 0.8 km s−1 kpc−1 and 1.2 km s−1
kpc−2, respectively.
Let us study the biases that have an effect on the de-
termination of the spiral structure parameters. As a gen-
eral conclusion, Figs. B.1 and B.2 show that our sample
of O and B stars supplies better results than the Cepheid
sample.
In the case of O and B stars, we found a clear depen-
dence with ψ⊙ and Ωp in ψ⊙, Πb and Θb determinations,
whereas f⊙ and Ωp only show peculiar behaviour around
ψ⊙ = 270◦. Concerning ψ⊙, the bias oscillates from −20◦
to 30◦. The standard deviation of the mean for the 50
samples of each set is about 10-20◦. On the other hand,
for the amplitudes Πb and Θb the biases are of ±2 km
s−1, with a standard deviation of about 1 km s−1. Neither
f⊙ nor Ωp have a considerable bias, except for ψ⊙ = 270◦,
where both biases and standard deviations go up.
For Cepheid stars similar results were obtained, but
with larger standard deviations in all cases. The bias in
ψ⊙ changes from −20◦ to 10◦, with standard deviations
of about 30-60◦. In the case of Πb, we found a clear de-
pendence on both ψ⊙ and Ωp, with a bias of ±1.5 km s−1
and a standard deviation of about 2 km s−1. On the other
hand, for Θb the bias is smaller, from 0 to 0.8 km s−1,
and the standard deviation is 1-1.5 km s−1. As for O and
B stars, for f⊙ and Ωp small biases are found, though the
standard deviations are larger in this case.
B.2.2. Results considering possible errors in the choice
of the free parameters
An interesting point to analyse is the study of the biases
produced by a bad choice of the free parameters in our
model (m, i, ⊙, Θ(⊙)). In the same way as in the
previous section, we simulated 50 samples for each one of
the cases considered in the real resolution, i.e. cases A, B,
C and D (see Table B.2). The simulated parameters were
the same as in Table B.1 for solar motion and galactic
rotation. For spiral arm kinematics, we considered ψ⊙ =
315◦ and Ωp = 30 km s−1 kpc−1 (similar values to those
obtained from real samples; see Sect. 6).
In Table B.2 we show the biases and standard devia-
tions when solving the model equations in crossed solu-
tions (e.g. we generated 50 simulated samples considering
the free parameters in case A, and then we solved equa-
tions using the free parameters adopted for cases A, B, C
and D, and so on for the other cases). As a first conclusion,
we can observe that a bad choice in the free parameters
Fig. B.2. Bias (obtained value−simulated value; left) and
standard deviation (right) for each set of 50 samples in the
solar motion (top), galactic rotation (middle) and spiral
arm kinematic (bottom) parameters for Cepheid pseudo-
stars (Case A). Values of Ωp: 10 km s−1 kpc−1 (black
double solid line), 15 (black solid line), 20 (grey solid line),
25 (dotted line), 30 (dashed line), 35 (long dashed line)
and 40 (dot-dashed line).
D. Fern´andez et al.: Kinematics of young stars
17
Table B.2. Bias and standard deviation in ψ⊙ and Ωp for crossed solutions for the simulated samples of O and B
stars and Cepheids. Units: ψ⊙ in degrees; Ωp in km s−1 kpc−1.
Cepheid stars with 0.6 < R < 4 kpc
O and B stars with 0.6 < R < 2 kpc
Case A Case B Case C Case D Case A Case B Case C Case D
∆ψ⊙
σψ⊙
∆Ωp
σΩp
∆ψ⊙
σψ⊙
∆Ωp
σΩp
∆ψ⊙
σψ⊙
∆Ωp
σΩp
∆ψ⊙
σψ⊙
∆Ωp
σΩp
23.
28.
2.2
6.2
32.
34.
0.7
7.1
18.
24.
6.7
6.5
29.
31.
4.3
7.1
17.
27.
3.1
7.3
24.
32.
1.9
8.0
14.
23.
8.2
7.3
23.
29.
6.3
8.2
24.
25.
−1.1
2.9
32.
31.
−1.6
3.8
17.
21.
1.2
2.9
25.
27.
0.2
3.7
25.
30.
−1.3
3.9
19.
25.
−0.8
3.5
Case A simulated
−11.
65.
−4.3
18.9
Case B simulated
0.
83.
−4.3
16.2
Case C simulated
4.
62.
−2.4
25.1
Case D simulated
9.
84.
−3.9
17.1
14.
21.
2.0
3.3
21.
26.
1.1
3.8
0.
61.
−3.5
16.6
4.
75.
−14.3
52.8
2.
58.
−1.1
21.5
10.
78.
−4.9
19.5
−2.
71.
−2.5
8.8
−11.
82.
−4.3
7.0
−2.
60.
−1.0
10.9
−9.
81.
−3.2
7.8
3.
71.
−4.0
7.3
1.
78.
−4.1
8.1
10.
63.
−3.0
9.2
1.
73.
−2.8
8.6
does not substantially alter the derived kinematic param-
eters, particularly ψ⊙. In other words, for each set of sim-
ulated samples we obtained nearly the same values for the
parameters whether we solved the Eqs. (8) with the cor-
rect set of free parameters or with a wrong combination
of them. Differences in ψ⊙ do not exceed 10◦ for O and
B stars and 20◦ for Cepheids. In the case of Ωp we found
large differences in some cases, but always when the stan-
dard deviation was also large. This is especially true for
Cepheids. A remarkable point is that the minimum bias
was not always produced when we properly chose the free
parameters.
B.2.3. Conclusions
In the light of these results, we conclude that we are able
to determine the kinematic parameters of the proposed
model of the Galaxy from the real star samples described
in Sect. 2, supposing that the velocity field of the stars is
well described by this model. We studied case A (m = 2,
i = −6◦, ⊙ = 8.5 kpc, Θ(⊙) = 220 km s−1 in detail in
these simulations, but we also looked at the other combi-
nations of the free parameters (cases B, C and D), with
similar conclusions. Nevertheless, the study of crossed so-
lutions has shown that it will be very difficult to decide
between the several set of free parameters discussed in
Sect. 6 (see also Table B.2), owing to the small differences
obtained when changing the free parameters in the condi-
tion equations.
References
Amaral, L.H., & L´epine, J.R.D. 1997, MNRAS, 286, 885
Avedisova, V.S. 1989, Astrophysics (Tr. Astrofizika), 30, 83
Bash, F.N. 1981, ApJ, 250, 551
Battinelli, P. 1991, A&A, 244, 69
Beaulieu, J.-P. 1999, private communication
Bok, B.J. 1958, Observatory, 79, 58
Bok, B.J., & Bok, P.F. 1974, The Milky Way, Cambridge,
Harvard University Press
Burton, W.B. 1971, A&A, 10, 76
Burton, W.B. 1976, ARA&A, 14, 275
Byl, J., & Ovenden, M.W. 1978, ApJ, 225, 496
Comer´on, F., & Torra, J. 1991, A&A, 241, 57
Comer´on, F., Torra, J., & G´omez, A.E. 1994, A&A, 286, 789
Cr´ez´e, M., & Mennessier, M.O. 1973, A&A, 27, 281
Drimmel, R. 2000, A&A, 358, L13
Elmegreen, D.M. 1985, The Milky Way galaxy, eds. H. van
Worden, R.J. Allen, W.B. Burton, IAU Symposium 106,
255
Englmaier, P., & Gerhard, O. 1999, MNRAS, 304, 512
ESA 1997, The Hipparcos Catalogue, ESA SP-1200
Feast, M.W., & Catchpole, R.M. 1997, MNRAS, 286, L1
Feast, M.W., & Whitelock, P.A. 1997, MNRAS, 291, 683
Feast, M.W., Pont, F., & Whitelock, P.A. 1998, MNRAS, 298,
L43
Fern´andez, D. 1998, Degree of Physics (Master Thesis),
Universitat de Barcelona, Spain (available in Spanish lan-
guage from http://www.am.ub.es/∼dfernand)
Fernie, J.D., Beattie, B., Evans, N.R., & Seager, S. 1995, IBVS,
4148
Frink, S., Fuchs, B., Roser, S., & Wielen, R. 1996, A&A, 314,
430
Garmany, C.D., & Stencel, R.E. 1992, A&AS, 94, 211
Georgelin, Y.M., & Georgelin, Y.P. 1976, A&A, 49, 57
G´omez, A., & Mennessier, M.O. 1977, A&A, 54, 113
18
D. Fern´andez et al.: Kinematics of young stars
Glushkova, E.V., Dambis, A.K., Mel'nik, A.M., & Rastorguev,
A.S. 1998, A&A, 329, 514
Grenier, S. 1997, private communication
Hauck, B., & Mermilliod, J.C. 1998, A&AS, 129, 431
Kerr, F.J. 1969 ARA&A, 7, 39
Kerr, F.J., & Lynden-Bell, D. 1986, MNRAS, 221, 1023
L´epine, J.R.D., Mishurov, Yu.N., & Dedikov, S.Yu. 2001, AJ,
546, 234
Lin, C.C., & Shu, F.H. 1964, ApJ, 140, 646
Lin, C.C., Yuan, C., & Shu, F.H. 1969, ApJ, 155, 721
Luri, X., 2000, private communication
Maciel, W.J. 1993 Ap&SS, 206, 285
Marochnik, L.S., Mishurov, Yu.N., & Suchkov, A.A. 1972,
Ap&SS, 19, 285
Mel'nik, A.M., & Efremov, Yu.N. 1995, Astro. Lett., 21, 10
Mel'nik, A.M., Sitnik, T.G., Dambis, A.K., Efremov, Yu.N., &
Rastorguev, A.S. 1998, Astro. Lett., 24, 594
Mel'nik, A.M., Dambis, A.K., & Rastorguev, A.S. 1999, Astro.
Lett., 25, 518
Mennessier, M.O., & Cr´ez´e, M. 1975, in "La dynamique des
galaxies spirales", colloque n◦241, Centre National de la
Recherche Scientifique, Paris
Mermilliod, J.C. 1986, A&AS, 63, 293
Metzger, M.R., Caldwell, J.A.R., & Schechter, P.L. 1998, AJ,
115, 635
Mishurov, Yu.N., Zenina, I.A., Dambis, A.K., Mel'nik, A.M.,
& Rastorguev, A.S. 1997, A&A, 323, 775
Mishurov, Yu.N., & Zenina, I.A. 1999, A&A, 341, 81
Olano, C.A. 2001, AJ, 121, 295
Olling, R.P., & Merrifield, M.R. 1998, MNRAS, 297, 943
Pont, F., Mayor, M., & Burki, G. 1994, A&A, 285, 415
Pont, F., Queloz, D., Bratschi, P., & Mayor, M. 1997, A&A,
318, 416
Press, W.H., Teukolsky, S.A., Vetterling, W.T., & Flannery,
B.P. 1992, Numerical Recipes, Cambridge University Press,
Cambridge
Racine, R., & Harris, W.E. 1989, AJ, 98, 1609
Rastorguev, A.S., Glushkova, E.V., Zabolotskikh, M.V., &
Baumgardt, H. 2001, Astronomical and Astrophysical
Transactions, in press
Reid, M.J. 1993, ARA&A, 31, 345
Roberts Jr., W.W. 1969, ApJ, 158, 123
Roberts Jr., W.W. 1970, The spiral structure of our galaxy,
eds. W. Becker, G. Contopoulos, IAU Symposium 38, 415
Rohlfs, K. 1977, Lectures in density waves, Springer-Verlag,
Berlin
Royer, F. 1999, PhD Thesis, Observatoire de Paris-Meudon,
France
Schild R. 1967, ApJ, 148, 449
Schmidt, M. 1965, Galactic structure, eds. A. Blaauw and M.
Schmidt, University Chicago Press, Chicago
Schmidt-Kaler, T. 1975, Vistas Astron., 19, 69
Stibbs, D.W.N. 1956, MNRAS, 116, 453
Toomre, A. 1969, ApJ, 158, 899
Torra, J., Fern´andez, D., & Figueras, F. 2000, A&A, 359, 82
(Paper I)
Vall´ee, J.P. 1995, ApJ, 454, 119
Yuan, C. 1969, ApJ, 158, 889
|
astro-ph/9909162 | 1 | 9909 | 1999-09-09T19:22:57 | Can Light Echoes Account for the Slow Decay of Type IIn Supernovae? | [
"astro-ph"
] | The spectra of type IIn supernovae indicate the presence of apre-existing slow, dense circumstellar wind (CSW). If the CSW extends sufficiently far from the progenitor star, then dust formation should occur in the wind. The light from the supernova explosion will scatter off this dust and produce a light echo. Continuum emission seen after the peak will have contributions from both this echo as well as from the shock of the ejecta colliding with the CSW, with a fundamental question of which source dominates the continuum. We calculate the brightness of the light echo as a function of time for a range of dust shell geometries, and use our calculations to fit to the light curves of SN 1988Z and SN 1997ab, the two slowest declining IIn supernovae on record. We find that the light curves of both objects can be reproduced by the echo model. However, their rate of decay from peak, color at peak and their observed peak absolute magnitudes when considered together are inconsistent with the echo model. Furthermore, when the observed values of M$_{B}$ are corrected for the effects of dust scattering, the values obtained imply that these supernovae have unrealistically high luminosities. We conclude that light echoes cannot properly account for the slow decline seen in some IIn's, and that the shock interaction is likely to dominate the continuum emission. | astro-ph | astro-ph | Can Light Echoes Account for the Slow Decay of Type IIn
Supernovae?
Bruce Roscherr and Bradley E. Schaefer
Physics Department, Yale University, PO Box 208120, New Haven, CT, 06520-8121;
[email protected], [email protected]
Received
;
accepted
9
9
9
1
p
e
S
9
1
v
2
6
1
9
0
9
9
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
-- 2 --
ABSTRACT
The spectra of type IIn supernovae indicate the presence of apre-existing
slow, dense circumstellar wind (CSW). If the CSW extends sufficiently far from
the progenitor star, then dust formation shouldoccur in the wind. The light
from the supernova explosion will scatter off this dust and produce a light
echo. Continuum emission seen after the peak will have contributions from
both this echo as well as from the shock of the ejecta colliding with the CSW,
with a fundamental question of which source dominates the continuum. We
calculate the brightness of the light echo as a function of time for a range of
dust shell geometries, and use our calculations to fit to the light curves of SN
1988Z and SN 1997ab, the two slowest declining IIn supernovae on record. We
find that the light curves of both objects can be reproduced by the echo model.
However, their rate of decay from peak, color at peak and their observed peak
absolute magnitudes when considered together are inconsistent with the echo
model. Furthermore, when the observed values of MB are corrected for the
effects of dust scattering, the values obtained imply that these supernovae have
unrealistically high luminosities. We conclude that light echoes cannot properly
account for the slow decline seen in some IIn's, and that the shock interaction is
likely to dominate the continuum emission.
-- 3 --
1.
Introduction
Supernovae are designated to be of type IIn based on their spectra (Schelegel 1990).
The group is quite heterogeneous, but a typical IIn would have the following properties: an
Hα profile composed of a narrow peak (∼ 100 km/s) sitting on a broad base (∼ 104 km/s),
no P Cygni absorption feature, a strong blue continuum, and slow spectral evolution. There
is occasionally an intermediate width component (∼ 2000 km/s) to the Hα line. Absorption
lines tend to be weak or absent. The narrow Hα component is seen to vary in intensity
and has thus been interpreted to indicate the presence of a pre-existing slow-moving stellar
wind. These winds can be very dense, with M as high as 10−2 M⊙/yr as is the case with
SN1997ab (Salamanca et.al. 1998). The overall spectral properties of these supernovae
have successfully been interpreted as being due to a shock interaction between the rapidly
expanding supernova ejecta and this dense (106 − 108cm−3) stellar wind (Chevalier &
Fransson 1985, Chugai 1991, Terlevich 1994). IIn's comprise about 5% of observed type II
supernovae.
A number of IIn's show an abnormally slow decline in their light curves. SN1997ab
and SN 1988Z are the extreme examples of this, both declining by only about 5 magnitudes
in 1000 days. Simple models of the shock interaction suggest that the shock can produce
sufficient optical continuum emission to account for this slow decline (Terlevich 1994, Plewa
1995). It was suggested by Chugai(1992) that light echoes might also provide a significant
contribution to the late time light curves of IIn's, i.e. the observed continuum should be a
combination of shock and echo contributions. When the supernova explodes, it vaporizes
any dust in the CSW out to a radius of ∼ 1016 cm (Pearce & Mayes 1986). Any dust that
lies outside this evaporation radius will scatter the supernova light. Scattered light travels
a longer path before reaching us and so is delayed relative to any unscattered light, i.e. a
light echo is produced. As little as 10−6M⊙ of dust is required to make the circumstellar
-- 4 --
shell of matter optically thick to dust scattering.
We here investigate whether light echoes could help account for the slow decline in the
light curve seen in some IIn's. We have performed a detailed calculation of the light echo
produced by a range of dust shell geometries. Our assumptions and method of calculation
are described in the next section. Our results are presented in Section 3, and in Section 4
we fit to the observed light curves of SN 1988Z and SN 1997ab. Section 5 is devoted to the
discussion of our results.
2. Calculating the Light Echo
The light echo in the U, B, V , and R bands was computed by a Monte Carlo
simulation. The calculation requires knowledge of the dust size distribution, the dust spatial
distribution, the optical properties of the dust, and the underlying Type II supernova light
curve.
The dust was assumed to be composed of a mixture of silicon and graphite grains, in
the ratio 3:1 by number. Both grain types were assumed to have a grain size distribution
described by the MRN grain model (e.g. Evans 1994), i.e.
the number density of
grains varies as n(a) ∝ a−3.5 where a is the grain diameter. Our grain size ranged from
amin = 0.005µm up to amax = 1.0µm.
We assumed that the dust lies in a spherical shell of inner radius Rmin and outer radius
Rmax about the star. Rmin is the radius out to which the supernova evapourates all the
dust in the CSW. Rmin can be estimated from first principles and for a type II supernova
has a value of a few times 1016 cm (Pearce & Mayes 1986). We used this value as a guide,
and took both Rmin and Rmax to be free parameters. We assumed that the dust density in
the shell varies as ρ(r) ∝ r−2. The amount of dust in the shell is characterized by τ , the R
-- 5 --
band optical depth for a purely radial photon trajectory. The total dust mass is related to
τ by the following expression:
Mdust =
8π
τ
max
− a0.5
min(cid:17) RmaxRmin (f ρgr + (1 − f )ρSi) ×
3 (cid:16)a0.5
R amax
amin da a−1.5Pi=abs,scat (f Qi,gr + (1 − f )Qi,Si)
= 7.0 × 10−7 τ Rmin
0.01pc! Rmax
1.0pc! M⊙
(1)
where f = 0.25, ρgr = 2.5g cm−3 and ρSi = 2.35g cm−3 are the densities of graphite and
silicon respectively, and Qabs and Qscat are the absorption and scattering efficiencies.
The photons are either absorbed by the dust, or they scatter elastically. The cross
sections for the dust-photon interactions are functions of both the photon wavelength and
the dust grain size. We used the scattering and absorption cross sections, σabs(= 4πa2Qabs)
and σscat(= 4πa2Qscat), calculated by Draine(1987). The angular distribution for the
scattering is given by the Henyey-Greenstein function (Henyey & Greenstein 1941),
F (θ) =
1 − g2
(1 + g2 − 2g cos θ)3/2
(2)
where g is the degree of forward scattering and is a function of wavelength and grain size.
Values for g were also taken from Draine(1987).
We used the composite II-L and II-P light curves calculated by Doggett and
Branch(1985) as input. They provide averaged light curves for the B and V bands. We
took the U and B bands to be initially identical, and likewise the V and R. The input
supernova was taken to have B − V = 0 at peak.
EDITOR: PLACE FIGURE 1 HERE.
Our Monte Carlo code follows the photon trajectories in three dimensions. It keeps
track of the time delay for each photon, as well as the photon weight. The weight was
-- 6 --
decreased at each interaction by the absorption probability. The photon was also split
at each interaction into a part which has no further interactions and thus escapes, and a
second part which is forced to interact at least once more. This substantially improves the
output statistics. Our code was tested against a number of analytical test cases, and also
against the calculations of Chevalier(1986), who calculated light echoes in the limit of low
optical depth.
Our code produces the R band light curve as well as the V − R, B − V , and U − B
colors as output. The output is dependent on the choice of Rmin, Rmax, and τ . 106 input
photons were used to calculate each band for each choice of the model parameters. Our
output does not include any contribution from line emission.
3. Results
Our echo model has three parameters: Rmin, Rmax and τ . Figures 2,3, and 4 show the
effect of varying each of these in turn.
EDITOR: PLACE FIGURE 2 HERE.
EDITOR: PLACE FIGURE 3 HERE.
EDITOR: PLACE FIGURE 4 HERE.
We see from the figures that the echo is fairly insensitive to both Rmin and Rmax as long
as the shell is thicker than a few hundredths of a parsec. As the shell becomes thinner,
the light curves fall off faster, and more closely trace the intrinsic light curve. We can also
-- 7 --
see that the presence of a small amount of dust has a pronounced effect on the light curve
at late times. The optical depths in the U, B, V and R bands are always in the ratio
1.56:1.35:1.20:1.00. The V − R, B − V and U − B colors remain nearly constant after about
200 days for all the non-thin shell cases. For the the thin shells the colors tend to redden at
late times.
While there are variations amongst models, the following relations serve as useful
approximations:
(U − B)peak ∼ 0.17τ , (B − V )peak ∼ (V − R)peak ∼ 0.15τ
∆mU ∼ 1.4τ , ∆mB ∼ 1.3τ , ∆mV ∼ 1.1τ, ∆mR ∼ 0.95τ
β100,U ∼ 4.3 − 1.5τ, β100,B ∼ 4.3 − 1.4τ , β100,V ∼ 4.0 − 1.2τ, β100,R ∼ 4.0 − 1.1τ
∆m is the number of magnitudes by which the peak of the light curve has faded due to
dust scattering compared to the peak in the composite type II light curve, and β100 is the
number of magnitudes by which the light curve drops from peak over the first 100 days
after peak.
4. Fits to SN 1988Z and SN 1997ab
SN 1988Z and SN 1997ab are the two slowest declining IIn's on record. (SN 1995N has
a similar decay rate, but data on this object is not yet available). The slow decline suggests
that these objects are embedded in a dense CSW. This is bourne out by analysis of the
spectra of these objects. Salamanca et.al.(1998) calculate that the electron density in the
CSW of SN 1997ab has a value close to 108 cm−3. Chugai(1992) found a similar result for
SN 1988Z. We have used our echo model to fit to the light curves of these two supernovae.
-- 8 --
4.1. SN 1988Z
SN 1988Z was discovered on 1988 December 12 in the galaxy MGC+03-28-022
(Cappelloro & Turatto, 1988). The supernova was past peak at this time, but its high
luminosity (mB = 16.75, MB = -18.3 for H0 = 65 kms−1Mpc−1) argues that the maximum
occured not long before discovery. The supernova faded monotonically at an unusually slow
rate. Over the first 137 days it faded by only 1.5 magnitudes in both the B and V bands,
whereas the composite II-L light curve drops by some 5 magnitudes over the same period.
The color evolution of SN 1988Z is very unusual (Turatto et.al. 1993). Over the first 100
days B − V decreases from 0.4 to 0.1 mag. The supernova then becomes redder, until the
color reaches a constant value of about 0.65 mag after 700 days.
Figure 5 shows our fit to the B and V band light curves of SN 1988Z. The B band fit
corresponds to τ = 2.0 ±0.2, Rmin = 0.10 ±0.02 pc, Rmax = 0.35 ±0.05 pc. This corresponds
to a dust mass in the shell of (eq. (1)) (4.9 ± 1.1) × 10−6M⊙. The narrow component of
the Hα line is unresolved, so if we adopt a value of 100 kms−1 for the wind velocity, and
assume a constant mass loss rate, we infer that the progenitor star emitted material for
about 3400 years before going supernova at a rate of (1.3 ± 0.3) × 10−6 (cid:16) fdust
where fdust is the fraction of the wind mass which has formed into dust. The fit to the B
10−3(cid:17)−1
M⊙yr−1,
band determines the fit to the V band, i.e. the V band is not fit separately. In both cases,
the fit is reasonably good up to about 800 days, after which the observed light curves are
seen to flatten significantly with respect to the model predictions.
EDITOR: PLACE FIGURE 5 HERE.
-- 9 --
4.2. SN 1997ab
SN 1997ab was discovered on 1996 April 11 in the galaxy HS 0948+2018 (Hagen &
Reimers 1997). It had a B magnitude of 14.7 which corresponds to an absolute magnitude,
MB = −19.1 (H0 = 65kms−1Mpc−1). The date of peak is uncertain, but the high luminosity
at discovery argues that the peak had occured recently. The light curve of this object is
unfortunately not well sampled, there being only three other B band observations and two
V band observation since discovery (Hagen, Engels & Reimers 1997, Schaefer & Roscherr
1998, 1999).
Our model fits the B band light curve (figure 6) for τ = 3.0 ± 0.3, Rmin = 0.10 ± 0.02
pc, Rmax = 0.30 ± 0.05 pc. This implies a dust mass in the shell of (6.3 ± 1.4) × 10−6M⊙.
The wind velocity is measured to be 90 kms−1, and so we infer that the progenitor
emitted material for some 3000 years before going supernova at an average rate of
(2.9 ± 0.7) × 10−6 (cid:16) fdust
10−3(cid:17)−1
M⊙yr−1. Salamanca et.al. (1998) find a mass loss rate of
approximately 10−2M⊙yr−1 from the analysis of the spectrum of this object. They find that
the emitted material lies in a shell that extends no more than 0.05pc from the progenitor.
This implies that the progenitor had a brief episode of strong mass loss shortly before going
supernova. The material emited during this episode lies too close to the star to allow any
dust formed to survive the supernova explosion. The dust which is responsible for the echo
must have been emitted during an earlier, more sustained period of lower mass loss.
EDITOR: PLACE FIGURE 6 HERE.
-- 10 --
5. Echo Model versus Shock Model
If the progenitors of Type II supernovae emit material for more than a few hundered
years prior to going supernova, then they should be surrounded by a dusty CSW. All the
dust that lies beyond the evapouration radius of the supernova will scatter the supernova
light and produce a light echo. There is likely to be an echo component in the light curves
of many Type II supernovae, the only quesiton is how significant a component it is.
We have seen in the previous section that the echo model can provide reasonably good
fits to the light curves of SN 1988Z and SN 1997ab. Chugai(1992), however, produced
a similar quality fit to the light curve of SN 1988Z with a simple model of the shock
continuum emission. Fits to the light curves alone are thus not sufficient to discriminate
between the models.
In figure 7 we have plotted(B − V )peaK against β100,B. The region between the two
solid lines is the region of parameter space allowed by the echo model. As the amount
of circumstellar dust increases, we expect the supernova light to suffer more scatterings,
producing an echo that becomes progressively redder and longer lived. In shock models the
optical light is primarily produced by the reprocessing of X-rays. The color of the optical
light produced in this way is not very sensitive to the amount of circumstellar material
present, so on the plot of (B − V )peak versus β100,B, we do not expect to see a correlation.
SN 1988Z and SN 1997ab both lie within the region allowed by the echo model. The other
IIn's plotted on the figure, however, do not. Their distribution shows no definite correlation,
and is thus consistent with the shock model prediction. The IIn's plotted represent the
complete sample for which (B − V )peak, β100,B, and MB,peak are available in the literature
(Patat et.al. 1994). The data for SN 1998S was taken from the CfA Supernova Group
website.
In figure 8 we have plotted the observed MB,peak versus β100,B. In the echo model, for a
-- 11 --
given underlying supernova luminosity, as the dust mass increases, the decay rate decreases
and the observed MB,peak increases, i.e. the supernova appears fainter as more of the
emission at peak is scattered by the dust. The opposite trend is expected for shock models,
i.e. we expect the peak luminosity to increase as the amount of circumstellar material
increases. In an MB,peak versus β100,B plot, shock models thus predict a downward sloping
line. The observed IIn points are consistent with the shock prediction, and inconsistent
with the prediction of the echo model. In particular, the points for both SN 1997ab and SN
1988Z lie outside the region allowed by the echo model.
In figure 9 we have plotted the observed MB,peak versus (B − V )peak. The echo model
predicts that the emission should redden as the amount of dust, and thus MB,peak, increases.
The observed data points are inconsistent with this prediction.
Another argument against the echo model is the implied values for MB,peak for the
underlying supernovae. SN 1988Z had an observed MB,peak of -18.3 and SN 1997ab, -19.1.
If we use the values of τ , Rmin and Rmax obtained from the fits to the light curves to find
∆mB, we find that the underlying supernovae must have had MB,peak of -20.9 in the case of
SN 1988Z, and -22.8 for SN 1997ab. This would make these supernovae by far the brightest
Type II supernovae yet observed.
We conclude that light echoes cannot properly account for the slow decline seen in some
IIn's. The predictions of shock models are consistent with the data, and shocks are likely
to dominate the continuum emission. The absence of a strong echo component in the light
curves of IIn's argues that the progenitors of these supernovae undergo the bulk of their
mass loss just prior to going supernova, and thus very little dust survives the explosion.
-- 12 --
REFERENCES
Cappellaro, E. and Turatto, M. 1988,IAU Circ. No. 4691
Chevalier, R.A., & Fransson, C. 1985, In: Supernovae as Distance Indicators, ed. Bartel,
N., Springer, Berlin
Chevalier, R.A. 1986,Ap.J.,308,225
Chugai, N.N. 1991,MNRAS,250,513
Chugai, N.N. 1992,Sov.Astron.,36,63
Doggett, J.B., & Branch, D. 1985,A.J.,90,2303
Draine, B.T. 1987,Princeton Observatory Preprints POP213
Evans, A. 1994,The Dusty Universe, John Wiley & Sons, New York
Hagen, H.J., Engels, D., & Reimers, D. 1997,A&A,324, L29
Hagen, H.J., & Reimers, D. 1997,IAU Circ. No. 6589
Henyey, L.G., & Greenstein, J.L. 1941,Ap.J.,93, 70
Patat, F., Barbon, R., Cappellaro, E., & Turatto, M. 1994,A&A,282, 731
Pearce, G., & Mayes, A.J. 1986,A&A,155,291
Plewa, T. 1995,MNRAS,275,143
Salamanca, I. et.al. 1998,MNRAS,300,L17
Schaefer, B.E., & Roscherr, B. 1998,IAU Circ. No. 7058
Schaefer, B.E., & Roscherr, B. 1999,IAU Circ. No. 7141
-- 13 --
Schlegel, E.M. 1990,MNRAS,244,269
Terlevich, R.J. 1994, In:Circumstellar Media in the Late Stages of Stellar Evolution, ed.
Clegg, R.E.S., Stevens, I.R., & Meikle, W.P.S., Cambridge University Press,153
Turatto, M. 1993,et.al.,MNRAS,262,128
This manuscript was prepared with the AAS LATEX macros v4.0.
-- 14 --
Fig. 1. -- (a) and (b) respectively show the composite II-L and the composite II-P light
curves used as input for the echo calculations. (c) and (d) show the values of B − V for each
set of composite curves Doggert & Branch(1985).
Fig. 2. -- The effect on the light curves due to changes in τ . Rmin and Rmax are held fixed
at 0.1 and 1.0 pc respectively. (a) and (b) show the R band light curve for input II-L and
II-P respectively. The remaining plots are for V − R, B − V and U − B for each of the two
sets of input. The curves have been shifted downward by ∆ mags for clarity.
Fig. 3. -- The effect of changes in Rmin. The curves are for Rmin = 0.01, 0.5, 0.95 and 0.99
pc. τ is held fixed at 1.0 and Rmax at 1.0 pc. (a) and (b) show the R band light curve for
input II-L and II-P respectively. The remaining plots are for V − R, B − V and U − B for
each of the two sets of input. The curves have been shifted downward by ∆ mags for clarity.
Fig. 4. -- The effect of changes in Rmax. The curves are for Rmax = 0.11, 0.15, 0.5, 1.0 and
2.0 pc. τ is held fixed at 1.0 and Rmin at 0.1 pc. (a) and (b) show the R band light curve
for input II-L and II-P respectively. The remaining plots are for V − R, B − V and U − B
for each of the two sets of input. The curves have been shifted downward by ∆ mags for
clarity.
Fig. 5. -- Fit to the B (crosses) and V (circles) band light curves of SN 1988Z (Turatto et.al.
1993).
Fig. 6. -- Fit to the B (crosses) and V (circles) band light curves of SN 1997ab.
Fig. 7. -- Plot of (B − V )peak versus β100,B. The area between the solid lines are the values
allowed by the echo model. The underlying supernova was taken to have B − V = 0 at peak.
Fig. 8. -- Plot of MB,peak versus β100,B. MB,peak for the underlying supernova was allowed to
range between -16 and -19.
-- 15 --
Fig. 9. -- Plot of MB,peak versus (B − V )peak. For the underlying supernova, MB,peak was
allowed to range between -16 and -19, and we assumed (B − V )peak = 0.
0.6
0.4
0.2
0
-0.2
-0.4
SN 1997ab
SN 1987B
SN 1988Z
echo model
SN 1987F
SN 1998S
SN 1984E
SN 1989C
SN 1983K
0
2
4
6
-20
-18
-16
-14
-12
0
SN 1983K
SN 1997ab
SN 1987F
SN 1988Z
echo model
SN 1998S
SN 1989C
SN 1987B
SN 1984E
2
4
6
-20
-18
-16
-14
-12
SN 1983K
SN 1997ab
SN 1987F
SN 1998S
SN 1988Z
SN 1989C
SN 1987B
SN 1984E
echo model
-0.4
-0.2
0
0.2
0.4
0.6
|
astro-ph/0202005 | 1 | 0202 | 2002-01-31T23:27:04 | A Cosmic Perspective from Lapland in 2001 | [
"astro-ph"
] | A convergence of ideas, observations and technology have led to the greatest period of cosmological discovery yet. Over the past three years we have determined the basic features of our Universe. We are now challenged to make sense of what we have found. The outcome of planned experiments and observations as well as new ideas will be required. If we succeed, ours truly will be a Golden Age of Cosmology. | astro-ph | astro-ph |
A Cosmic Perspective from Lapland in 2001
MICHAEL S. TURNER1
Center for Cosmological Physics
Departments of Astronomy & Astrophysics and of Physics
Enrico Fermi Institute, The University of Chicago
Chicago, IL 60637-1433, USA
NASA/Fermilab Astrophysics Center
Fermi National Accelerator Laboratory
Batavia, IL 60510-0500, USA
E-mail: [email protected]
A convergence of ideas, observations and technology have led to the great-
est period of cosmological discovery yet. Over the past three years we have
determined the basic features of our Universe. We are now challenged to make
sense of what we have found. The outcome of planned experiments and ob-
servations as well as new ideas will be required. If we succeed, ours truly will
be a Golden Age of Cosmology.
PRESENTED AT
COSMO-01
Rovaniemi, Finland,
August 29 -- September 4, 2001
1Work supported by the NASA at Fermilab and the DOE at Chicago and Fermilab.
1 How Did We Get Here?
Cosmology is currently in its most exciting period ever. However, the past hundred
years haven't been too bad either. In 1916, Einstein introduced general relativity, the
first theoretical framework up to describing the evolution of the Universe. In the next
decade, Hubble established the extragalactic nature of the nebulae and put forth the first
systematic evidence for the expansion of the Universe. This burst of activity was powered
by ideas -- general relativity -- and technology -- big telescopes on high mountains -- and
marked the beginning of modern cosmology.
Cosmology then stalled for some twenty years. During that time, the 100-inch Mt.
Wilson telescope and its counterpart at the Lick Observatory charted only a few hundred
galaxies, with redshifts less than z = 0.1 [1]. Most of the Universe lay beyond their reach.
While some important new theoretical ideas were introduced, most notably the idea of a
hot big bang by Gamow and his collaborators, theory did little to push things forward
either.
First light at the 200-inch Hale telescope on Palomar and the steady state vs. hot big
bang controversy signaled the beginning of a new period of discovery around 1950. Once
again, it was the combination of theory and observation that propelled the field forward.
The discovery of quasars and radio-source counts murdered cosmology's most beautiful
theory -- the steady state -- in the early 1960s. I note the important role played here by a
strong theory -- that is, one that makes sharp predictions and is therefore easily falsified.
Because the steady state predicts a nonevolving Universe, any evidence of evolution can
falsify it (here, the distribution of radio sources vs. their strength and the large abundance
of quasars seen at high redshift).
The discovery of the cosmic background radiation in 1964 by Penzias and Wilson
broke things wide open, and cosmology became a legitimate branch of physics. The
details of Gamow's nucleosynthesis scheme were worked out in detail and became big-
bang nucleosynthesis, and the large mass fraction of 4He predicted by the hot big-bang
model became its first success.
In Chapter 15 of his influential text, Gravitation and
Cosmology, Steven Weinberg laid out "The Standard Cosmology" and coined the term [2].
The Standard Cosmology meant the hot big-bang model, from shortly before the epoch of
big-bang nucleosynthesis to the present. A convergence of ideas (a better understanding
of general relativity and the careful application of physics to the early Universe) and
observations (the discovery of the cosmic microwave background) again had powered a
significant advance our understanding of the Universe.
According to Weinberg's same text, the period before BBN (t <∼ 10−4 sec) was "cosmos
incognito" (cf, Section 15.11). He recognized quite correctly that the key to further
progress was a better understanding of the elementary particles. A subatomic world of
strongly interacting particles with exponentially rising numbers, which was the world
view of subatomic physics then, leads to a maximum temperature and a breakdown of
any simple statistical mechanical treatment.
1
The breakthrough came in the mid 1970s with the emergence of the standard model
of particle physics and its point-like quark/lepton constituents with asymptotically weak
interactions. The grander ideas about unification of the strong and electroweak interac-
tions (GUTs) that came somewhat later were even more influential. Asymptotically free
gauge theories, GUTs and the sturdy framework of the hot big-bang fueled a decade of
very fertile speculation about the early Universe in the 1980s ("the go-go junk bond days
of early-Universe cosmology").
There was optimism that the marriage of ideas about unification with early-Universe
cosmology could solve some of the most pressing questions in cosmology. The puzzles at
the time included [3]: What is the origin of the baryon asymmetry? Why is the Universe
so smooth, flat, and old? How did the primeval lumpiness that seeded all structure
arise? What is the dark matter? And the marriage brought a new problem: the glut of
superheavy magnetic monopoles expected from the GUT phase transition (about one per
baryon!).
Grand ideas, clever scenarios, and interesting schemes were suggested -- topological
defects, baryogenesis, particle dark matter, phase transitions, inflation, superheavy relics,
superconducting cosmic strings, unstable relics, colored relics, relics produced by the
decay of other relics, hot dark matter, cold dark matter, warm dark matter, shadow
worlds, mirror worlds, parallel universes, and on and on.
It was a fantastic time for
theorists.
A handful of these ideas stood the test of time. In the go-go 80s, that meant surviving
more than a month without being falsified, being replaced by a better idea, or going out of
style. The survivors -- particle dark matter, inflation, topological defects, and baryogenesis
-- began to form the basis of a new cosmological framework. However, only a few years
later, the very beautiful idea that topological defects seeded large-scale structure was
killed by cosmic microwave background measurements that began to show a series of well
formed acoustic peaks. (Only adiabatic density perturbations lead to such a structure in
the CMB power spectrum.)
Inflation and cold dark matter survived the cut. Moreover, they were expansive and
eminently testable. Theorists like to think (and with some good reason) that the tremen-
dous growth in observational cosmology over the past decade has something to do with
how powerful their ideas are. Technology -- the advent of large format CCDs, 6-meter,
8-meter, and 10-meter ground based telescopes, the Hubble Space Telescope, HEMTs,
bolometers and large-scale computing -- of course played a role too. An avalanche of data,
which has made cosmological phenomenology possible and the term "precision cosmology"
a reality, started in the mid 1990s. And there is no end in sight.
We are now in the midst of a great period of cosmic discovery. Once again, it came
about through a convergence of ideas and technology. Thus far, we have determined the
basic features of our Universe, which are pointing to a new standard cosmological model.
While we learned much about the Universe, we still have much more to understand.
2
2 The New Cosmology
Over the past three years a New Cosmology has been emerging. It incorporates the highly
successful standard hot big-bang cosmology [2, 4] and may extend our understanding of
the Universe to times as early as 10−32 sec, when the largest structures in the Universe
were still subatomic quantum fluctuations.
This New Cosmology is characterized by
• Flat, critical density accelerating Universe
• Early period of rapid expansion (inflation)
• Density inhomogeneities produced from quantum fluctuations during inflation
• Composition: 2/3 dark energy; 1/3 dark matter; 1/200 bright stars
• Matter content: (29 ± 4)% cold dark matter; (4 ± 1)% baryons; >∼ 0.3% neutrinos
• T0 = 2.725 ± 0.001 K
• t0 = 14 ± 1 Gyr
• H0 = 72 ± 7 km s−1 Mpc−1
The New Cosmology is not as well established as the standard hot big-bang cosmology.
However, the evidence is growing.
2.1 Mounting Evidence: Recent Results
The position of the first acoustic peak in the multipole power spectrum of the anisotropy
of the cosmic microwave background (CMB) radiation provides a powerful means of de-
termining the global curvature of the Universe. With the recent DASI observations of
CMB anisotropy on scales of one degree and smaller, the evidence that the Universe is at
most very slightly curved is quite firm [5]. The curvature radius of the Universe (≡ Rcurv)
and the total energy density parameter Ω0 = ρTOT/ρcrit, are related:
Rcurv = H −1
0 /Ω0 − 11/2
The spatial flatness is expressed as Ω0 = 1.0 ± 0.04, or said in words, the curvature radius
is at least 50 times greater than the Hubble radius.
I will discuss the evidence for accelerated expansion and dark energy later.
The series of acoustic peaks in the CMB multipole power spectrum and their heights
indicate a nearly scale-invariant spectrum of adiabatic density perturbations with n = 1±
0.07. Nearly scale-invariant density perturbations and a flat Universe are two of the three
3
hallmarks of inflation. Thus, we are beginning to see the first significant experimental
evidence for inflation, the driving idea in cosmology for the past two decades.
The striking agreement of the BBN determination of the baryon density from measure-
ments of the primeval deuterium abundance [6, 7], ΩBh2 = 0.020 ± 0.001, with those from
from recent CMB anisotropy measurements [5], ΩBh2 = 0.022 ± 0.004, make a strong
case for a small baryon density, as well as the consistency of the standard cosmology
(h = H0/100 km sec−1 Mpc−1). There can now be little doubt that baryons account for
but a few percent of the critical density.
Our knowledge of the total matter density is improving, and becoming less linked to
the distribution of light. This makes determinations of the matter less sensitive to the
uncertain relationship between the clustering of mass and of light (what astronomers call
the bias factor b) [10]. Both the CMB and clusters of galaxies allow a determination of
the ratio of the total matter density (anything that clusters -- baryons, neutrinos, cold
dark matter) to that in baryons alone: ΩM /ΩB = 7.2 ± 2.1 (CMB) [8], 9 ± 1.5 (clusters)
[9]. Not only are these numbers consistent, they make a very strong case for something
beyond quark-based matter. When combined with our knowledge of the baryon density,
one infers a total matter density of ΩM = 0.33 ± 0.04 [10].
The many successes of the cold dark matter (CDM) scenario -- from the sequence of
structure formation (galaxies first, clusters of galaxies and larger objects later) and the
structure of the intergalactic medium, to its ability to reproduce the power spectrum of
inhomogeneity measured today -- makes it clear that CDM holds much, if not all, of the
truth in describing the formation of structure in the Universe.
The two largest redshift surveys, the Sloan Digital Sky Survey (SDSS) and the 2-degree
Field project (2dF), have each recently measured the power spectrum using samples of
more than 100,000 galaxies and found that it is consistent with that predicted in a flat
accelerating Universe comprised of cold dark matter [11]. The SDSS will eventually use a
sample of almost one million galaxies to probe the power spectrum. [Interestingly enough,
according to the 2dF Collaboration, bias appears to be a small effect, b = 1.0 ± 0.09 [12].]
All of this implies that whatever the dark matter particle is, it moves slowly (i.e.,
the bulk of the matter cannot be in the form of hot dark matter such as neutrinos) and
interacts only weakly (e.g., with strength much less than electromagnetic) with ordinary
matter.
The evidence from SuperKamiokande [13] for neutrino oscillations makes a strong
case that neutrinos have mass (Pi mν >∼ 0.1 eV) and therefore contribute to the mass
budget of the Universe at a level comparable to, or greater than, that of bright stars.
Particle dark matter has moved from the realm of a hypothesis to a quantitative question
-- how much of each type of particle dark matter is there in the Universe? Structure
formation in the Universe (especially the existence of small scale structure) suggests that
neutrinos contribute at most 5% or 10% of the critical density, corresponding to Pi mnu =
Pi mν/90h2 eV <∼ 5 eV [14].
Even the age of the Universe and the pesky Hubble constant have been reined in. The
4
uncertainties in the ages of the oldest globular clusters have been better identified and
quantified, leading to a more precise age, t0 = 13.5 ± 1.5 Gyr [15]. The CMB can be
used to constrain the expansion age, independent of direct measurements of H0 or the
composition of the Universe, texp = 14 ± 0.5 Gyr [16].
A host of different techniques are consistent with the Hubble constant determined
by the HST key project, H0 = 72 ± 7 km s−1 Mpc−1. Further, the error budget is now
well understood and well quantified [17]. [The bulk of the ±7 uncertainty is systematic,
dominated by the uncertainty in the distance to the LMC and the Cepheid period --
luminosity relation.] Moreover, the expansion age derived from this consensus Hubble
constant, which depends upon the composition of the Universe, is consistent with the
previous two age determinations.
The poster child for precision cosmology continues to be the present temperature of the
CMB. It was determined by the FIRAS instrument on COBE to be: T0 = 2.725 ± 0.001 K
[18]. Further, any deviations from a black body spectrum are smaller than 50 parts per
million. Such a perfect Planckian spectrum has made any noncosmological explanation
untenable.
2.2 Successes and Consistency Tests
To sum up, we have determined the basic features of the Universe: the cosmic mat-
ter/energy budget; a self consistent set of cosmological parameters with realistic errors;
and the global curvature. Two of the three key predictions of inflation -- flatness and
nearly scale-invariant, adiabatic density perturbations -- have passed their first significant
tests. Last but not least, the growing quantity of precision data are now testing the
consistency of the Friedmann-Robertson-Walker framework and General Relativity itself.
In particular, the equality of the baryon densities determined from BBN and CMB
anisotropy is remarkable. The first involves nuclear physics when the Universe was seconds
old, while the latter involves gravitational and classical electrodynamics when the Universe
was 400,000 years old.
The entire framework has been tested by the existence of the aforementioned acoustic
peaks in the CMB angular power spectrum. They reveal large-scale motions that have
remained coherent over hundreds of thousands of years, through a delicate interplay of
gravitational and electromagnetic interactions.
Another test of the basic framework is the accounting of the density of matter and
energy in the Universe. The CMB measurement of spatial flatness implies that the matter
and energy densities must sum to the critical density. Measurements of the matter density
indicate ΩM = 0.33 ± 0.04; and measurements of the acceleration of the Universe from
supernovae indicate the existence of a smooth dark energy component that accounts for
ΩX ∼ 0.67.
[The amount of dark energy inferred from the supernova measurements
depends its equation of state; for a cosmological constant, ΩΛ = 0.8 ± 0.16.]
Finally, while cosmology has in the past been plagued by "age crises" -- time back to
5
the big bang (expansion age) apparently less than the ages of the oldest objects within
the Universe -- today the ages determined by very different and completely independent
techniques point to a consistent age of 14 Gyr.
3 Mysteries
Cosmological observations over the next decade will test -- and probably refine -- the New
Cosmology [19]. If we are fortunate, they will also help us to make better sense of it. At
the moment, the New Cosmology has presented us with a number of cosmic mysteries --
opportunities for surprises and new insights. Here I will quickly go through my list, and
save the most intriguing to me -- dark energy -- for its own section.
3.1 Dark Matter
By now, the conservative hypothesis is that the dark matter consists of a new form of
matter, with the axion and neutralino as the leading candidates. That most of the matter
in the Universe exists in a new form of matter -- yet to be detected in the laboratory -- is
a bold and untested assertion.
Experiments to directly detect the neutralinos or axions holding our own galaxy to-
gether have now reached sufficient sensitivity to probe the regions of parameter space
preferred by theory.
In addition, the neutralino can be created by upcoming collider
experiments (at the Tevatron or the LHC), or detected by its annihilation signatures --
high-energy neutrinos from the sun, narrow positron lines in the cosmic rays, and gamma-
ray line radiation [20].
While the CDM scenario is very successful there are some nagging problems. They
may point to a fundamental difficulty or may be explained by messy astrophysics [21].
The most well known of these problems are the prediction of cuspy dark-matter halos
(density profile ρDM → 1/rn as r → 0, with n ≃ 1 − 1.5) and the apparent prediction
of too much substructure. While there are plausible astrophysical explanations for both
problems [22], they could indicate an unexpected property of the dark-matter particle
(e.g., large self-interaction cross section [23], large annihilation cross section [24], or mass
of around 1 keV). While I believe it is unlikely, these problems could indicate a failure of
the particle dark-matter paradigm and have their explanation in a radical modification
of gravity theory [21].
I leave for the "astrophysics to do list" an accounting of the dark baryons. Since
ΩB ≃ 0.04 and Ω∗ ≃ 0.005, the bulk of the baryons are optically dark. In clusters, the
dark baryons have been identified: they exists as hot, x-ray emitting gas. Elsewhere,
the dark baryons have not yet been identified. According to CDM, the bulk of the dark
baryons are likely to exist as hot/warm gas associated with galaxies, but this gas has not
been detected. [Since clusters account for only about 5 percent of the total mass, the bulk
of the dark baryons are still not accounted for.]
6
3.2 Baryogenesis
The origin of quark-based matter is not yet fully understood. We do know that the
origin of ordinary matter requires a small excess of quarks over antiquarks (about a part
in 109) at a time at least as early as 10−6 sec, to avoid the annihilation catastrophe
associated with a baryon symmetric Universe [4]. If the Universe underwent inflation, the
baryon asymmetry cannot be primeval, it must be produced dynamically ("baryogenesis")
after inflation since any pre-inflation baryon asymmetry is diluted away by the enormous
entropy production associated with reheating.
Because we also now know that electroweak processes violate B +L at a very rapid rate
at temperatures above 100 GeV or so, baryogenesis is more constrained than when the
idea was introduced more than twenty years ago. Today there are three possibilities: 1)
produce the baryon asymmetry by GUT-scale physics with B − L 6= 0 (to prevent it being
subsequently washed away by B + L violation); 2) produce a lepton asymmetry (L 6= 0),
which is then transmuted into the baryon asymmetry by electroweak B + L violation
[25]; or 3) produce the baryon asymmetry during the electroweak phase transition using
electroweak B violation [26].
While none of the three possibilities can be ruled out, the second possibility looks
most promising, and it adds a new twist to the origin of quark-based matter: We are
here because neutrinos have mass.
[In the lepton asymmetry first scenario, Majorana
neutrino mass provides the requisite lepton number violation.] The drawback of the first
possibility is the necessity of a high reheat temperature after inflation, TRH ≫ 105 GeV,
which is difficult to achieve in most models of inflation. The last possibility, while very
attractive because all the input physics might be measurable at accelerator, requires new
sources of CP violation at TeV energies as well as a strongly first-order electroweak phase
transition (which is currently disfavored by the high mass of the Higgs) [26].
3.3
Inflation
There are still many questions to be answered about inflation, including the most funda-
mental: did inflation (or something similar) actually take place!
A powerful program is in place to test the inflationary framework. Testing framework
involves testing its three robust predictions: spatially flat Universe; nearly-scale invariant,
nearly power-law spectrum of Gaussian adiabatic, density perturbations; and a spectrum
of nearly scale-invariant gravitational waves.
The first two predictions are being probed today and will be probed much more sharply
over the next decade. The value of Ω0 should be determined to much better than 1 percent.
The spectral index n that characterizes the density perturbations should be measured to
percent accuracy.
Generically, inflation predicts n−1 ∼ O(0.1), where n = 1 corresponds to exact scale
invariance. Likewise, the deviations from an exact power-law predicted by inflation [27],
7
dn/d ln k ∼ 10−4 − 10−2 will be tested. The CMB and the abundance of rare objects
such as clusters of galaxies will allow Gaussianity to be tested.
Inflationary theory has given little guidance as to the amplitude of the gravitational
waves produced during inflation. If detected, they are a smokin' gun prediction. Their
amplitude is directly related to the scale of inflation, hGW ≃ Hinflation/mPl. Together
measurements of n − 1 and dn/ ln k, they can reveal much about the underlying scalar
potential driving inflation. Measuring their spectral index -- a most difficult task -- provides
a consistency test of the single scalar-field model of inflation [28].
3.4 The Dimensionality of Space-time
Are there additional spatial dimensions beyond the three for which we have very firm evi-
dence? I cannot think of a deeper question in physics today. If there are new dimensions,
they are likely to be relevant for cosmology, or at least raise new questions in cosmology
(e.g., why are only three dimensions large? what is going on in the bulk? and so on).
Further, cosmology may well be the best means for establishing the existence of extra
dimensions.
3.5 Before Inflation, Other Big-bang Debris, and Surprises
Only knowing everything there is to know about the Universe would be worse than know-
ing all the questions to ask about it. Without doubt, as our understanding deepens, new
questions and new surprises will spring forth.
The cosmological attraction of inflation is its ability to make the present state of the
Universe insensitive to its initial state. However, should we establish inflation as part of
cosmic history, I am certain that cosmologists will begin asking what happened before
inflation.
Progress in cosmology depends upon studying relics. We have made much of the
handful we have -- the light elements, the baryon asymmetry, dark matter, and the CMB.
The significance of a new relic cannot be overstated. For example, detection of the cosmic
sea of neutrinos would reveal the Universe at 1 second.
Identifying the neutralino as the dark matter particle and determining its properties at
an accelerator laboratory would open a window on the Universe at 10−8 sec. By comparing
its relic abundance as derived from its mass and cross section with its actual abundance
measured in the Universe, one could test cosmology at the time the neutralino abundance
was determined.
And then there may be the unexpected. Recently, a group reported evidence for a
part in 105 difference in the fine-structure constant at redshifts of order a few from its
value today [29]. I remain skeptical, given possible astrophysical explanations, other much
tighter constraints to the variation of α (albeit at more recent times), and the absence of
8
a reasonable theoretical model. For reference, I was also skeptical about the atmospheric
neutrino problem because of the need for large-mixing angles.
4 Dark Energy: Seven Things We Know
The dark energy accounts for 2/3 of the stuff in the Universe and determines its destiny.
That puts it high on the list of outstanding problems in cosmology. Its deep connections
to fundamental physics -- a new form of energy with repulsive gravity and possible impli-
cations for the divergences of quantum theory and supersymmetry breaking -- put it very
high on the list of outstanding problems in particle physics [30, 31].
What then is dark energy? Dark energy is my term for the causative agent for the
current epoch of accelerated expansion. According to the second Friedmann equation,
R
R
= −
4πG
3
(ρ + 3p)
(1)
this stuff must have negative pressure, with magnitude comparable to its energy density,
in order to produce accelerated expansion [recall q = −( R/R)/H 2; R is the cosmic scale
factor]. Further, since this mysterious stuff does not show its presence in galaxies and
clusters of galaxies, it must be relatively smoothly distributed.
That being said, dark energy has the following defining properties: (1) it emits/absorbs
no light; (2) it has large, negative pressure, pX ∼ −ρX ; (3) it is approximately homoge-
neous (more precisely, does not cluster significantly with matter on scales at least as large
as clusters of galaxies); and (4) it is very mysterious. Because its pressure is comparable
in magnitude to its energy density, it is more "energy-like" than "matter-like" (matter
being characterized by p ≪ ρ). Dark energy is qualitatively very different from dark
matter, and is certainly not a replacement for it.
4.1 Two Lines of Evidence for an Accelerating Universe
Two independent lines of reasoning point to an accelerating Universe. The first is the
direct evidence based upon measurements of type Ia supernovae carried out by two groups,
the Supernova Cosmology Project [32] and the High-z Supernova Team [33]. These two
teams used different analysis techniques and different samples of high-z supernovae and
came to the same conclusion: the expansion of the Universe is speeding up, not slowing
down.
The recent serendipitous discovery of a supernovae at z = 1.76 bolsters the case signif-
icantly [34] and provides the first evidence for an early epoch of decelerated expansion[35].
SN 1997ff falls right on the accelerating Universe curve on the magnitude -- redshift di-
agram, and is a magnitude brighter than expected in a dusty open Universe or an open
Universe in which type Ia supernovae are systematically fainter at high-z.
9
The second, independent line of reasoning for accelerated expansion comes from mea-
surements of the composition of the Universe, which point to a missing energy component
with negative pressure. The argument goes like this: CMB anisotropy measurements in-
dicate that the Universe is nearly flat, with density parameter, Ω0 = 1.0 ± 0.04. In a flat
Universe, the matter density and energy density must sum to the critical density. How-
ever, matter only contributes about 1/3 of the critical density, ΩM = 0.33 ± 0.04. (This is
based upon measurements of CMB anisotropy, of bulk flows, and of the baryonic fraction
in clusters.) Thus, two thirds of the critical density is missing! Doing the bookkeeping
more precisely, ΩX = 0.67 ± 0.06 [10].
In order to have escaped detection, this missing energy must be smoothly distributed.
In order not to interfere with the formation of structure (by inhibiting the growth of
density perturbations), the energy density in this component must change more slowly
than matter (so that it was subdominant in the past). For example, if the missing 2/3 of
critical density were smoothly distributed matter (p = 0), then linear density perturba-
tions would grow as R1/2 rather than as R. The shortfall in growth since last scattering
(z ≃ 1100) would be a factor of 30, far too little growth to produce the structure seen
today.
The pressure associated with the missing energy component determines how it evolves:
ρX ∝ R−3(1+w)
⇒ ρX/ρM ∝ (1 + z)3w
(2)
where w is the ratio of the pressure of the missing energy component to its energy density
(here assumed to be constant). Note, the more negative w, the faster the ratio of missing
energy to matter decreases to zero in the past. In order to grow the structure observed
today from the density perturbations indicated by CMB anisotropy measurements, w
must be more negative than about − 1
2 [36].
For a flat Universe the deceleration parameter today is
q0 =
1
2
+
3
2
wΩX ∼
1
2
+ w
Therefore, knowing w < − 1
2 implies q0 < 0 and accelerated expansion. This independent
argument for accelerated expansion and dark energy makes the supernova case all the
more compelling.
4.2 Gravity Can Be Repulsive in Einstein's Theory, But ...
In Newton's theory, mass is the source of the gravitational field and gravity is always
attractive. In General Relativity, both energy and pressure source the gravitational field:
R/R ∝ −(ρ+3p), cf., Eq. 1. Sufficiently large negative pressure leads to repulsive gravity.
While accelerated expansion can be accommodated within Einstein's theory, that does
not preclude that the ultimate explanation lies in a fundamental modification of Einstein's
10
theory. Lacking any good ideas for such a modification, I will discuss how accelerated
expansion fits in the context of General Relativity. If the explanation for the accelerating
Universe ultimately fits within the Einsteinian framework, it will be a stunning new
triumph for General Relativity.
4.3 The Biggest Embarrassment in all of Theoretical Physics
Einstein introduced the cosmological constant to balance the attractive gravity of matter.
He quickly discarded the cosmological constant after the discovery of the expansion of the
Universe.
The advent of quantum field theory made consideration of the cosmological constant
obligatory, not optional: The only possible covariant form for the energy of the (quantum)
vacuum,
T µν
VAC = ρVACgµν,
is mathematically equivalent to the cosmological constant. It takes the form for a perfect
fluid with energy density ρVAC and isotropic pressure pVAC = −ρVAC (i.e., w = −1) and
is precisely spatially uniform. Vacuum energy is almost the perfect candidate for dark
energy.
Here is the rub: the quantum zero-point contributions arising from well-understood
physics (the known particles, integrating up to 100 GeV) sum to 1055 times the present
critical density. (Put another way, if this were so, the Hubble time would be 10−10 sec, and
the associated event horizon would be 3 cm!) This is the well known cosmological-constant
problem [30, 31].
While string theory currently offers the best hope for marrying gravity to quantum
mechanics, it has shed precious little light on the cosmological constant problem, other
than to speak to its importance. Thomas has suggested that using the holographic princi-
ple to count the available number of states in our Hubble volume leads to an upper bound
on the vacuum energy that is comparable to the energy density in matter + radiation [37].
While this reduces the magnitude of the cosmological-constant problem very significantly,
it does not solve the dark energy problem: a vacuum energy that is always comparable to
the matter + radiation energy density would strongly suppress the growth of structure.
The deSitter space associated with the accelerating Universe may pose serious prob-
lems for the formulation of string theory [38]. Banks and Dine argue that all explanations
for dark energy suggested thus far are incompatible with perturbative string theory [39].
At the very least there is high tension between accelerated expansion and string theory.
The cosmological constant problem leads to a fork in the dark-energy road: one path
is to wait for theorists to get the "right answer" (i.e., ΩX = 2/3); the other path is to
assume that even quantum nothingness weighs nothing and something else with negative
pressure must be causing the Universe to speed up. Of course, theorists follow the advice
of Yogi Berra: "When you see a fork in the road, take it."
11
4.4 Parameterizing Dark Energy: For Now, It's w
Theorists have been very busy suggesting all kinds of interesting possibilities for the dark
energy: networks of topological defects, rolling or spinning scalar fields (quintessence and
spintessence), influence of "the bulk", and the breakdown of the Friedmann equations
[31, 41]. An intriguing recent paper suggests dark matter and dark energy are connected
through axion physics [40].
In the absence of compelling theoretical guidance, there is a simple way to parameterize
dark energy, by its equation-of-state w [36].
The uniformity of the CMB testifies to the near isotropy and homogeneity of the
Universe. This implies that the stress-energy tensor for the Universe must take the perfect
fluid form [4]. Since dark energy dominates the energy budget, its stress-energy tensor
must, to a good approximation, take the form
TX
µ
ν ≈ diag[ρX , −pX, −pX , −pX]
(3)
where pX is the isotropic pressure and the desired dark energy density is
ρX = 2.7 × 10−47 GeV4
(for h = 0.72 and ΩX = 0.66). This corresponds to a tiny energy scale, ρ1/4
10−3 eV.
X = 2.3 ×
The pressure can be characterized by its ratio to the energy density (or equation-of-
state):
w ≡ pX /ρX
Note, w need not be constant; e.g., it could be a function of ρX or an explicit function of
time or redshift. (w can always be rewritten as an implicit function of redshift.)
For vacuum energy w = −1; for a network of topological defects w = −N/3 where
N is the dimensionality of the defects (1 for strings, 2 for walls, etc.). For a minimally
coupled, rolling scalar field,
w =
1
2
1
2
φ2 − V (φ)
φ2 + V (φ)
(4)
which is time dependent and can vary between −1 (when potential energy dominates)
and +1 (when kinetic energy dominates). Here V (φ) is the potential for the scalar field.
4.5 The Universe: The Lab for Studying Dark Energy
Dark energy by its very nature is diffuse and a low-energy phenomenon.
It probably
cannot be produced at accelerators; it isn't found in galaxies or even clusters of galaxies.
The Universe itself is the natural lab -- perhaps the only lab -- in which to study it.
12
The primary effect of dark energy on the Universe is determining the expansion rate.
In turn, the expansion rate affects the distance to an object at a given redshift z [≡ r(z)]
and the growth of linear density perturbations. The governing equations are:
H 2(z) = H 2
r(z) = Z z
0
0 (1 + z)3 hΩM + ΩX (1 + z)3wi
du/H(u)
0 = δk + 2H δk − 4πGρM δk
(5)
where for simplicity w is assumed to be constant and δk is the Fourier component of
comoving wavenumber k and overdot indicates d/dt.
The various cosmological approaches to ferreting out the nature of the dark energy
-- all of which depend upon how the dark energy affects the expansion rate -- have been
studied [43]. Based largely upon my work with Dragan Huterer [44], I summarize what
we now know about the efficacy of the cosmological probes of dark energy:
• Present cosmological observations prefer w = −1, with a 95% confidence limit w <
−0.6 [46].
• Because dark energy was less important in the past, ρX /ρM ∝ (1 + z)3w → 0 as
z → ∞, and the Hubble flow at low redshift is insensitive to the composition of the
Universe, the most sensitive redshift interval for probing dark energy is z = 0.2 − 2
[44].
• The CMB has limited power to probe w (e.g., the projected precision for Planck is
σw = 0.25) and no power to probe its time variation [44].
• A high-quality sample of 2000 SNe distributed from z = 0.2 to z = 1.7 could measure
w to a precision σw = 0.05 (assuming an irreducible systematic error of 0.14 mag).
If ΩM is known independently to better than σΩM = 0.03, σw improves by a factor of
three and the rate of change of w′ = dw/dz can be measured to precision σw′ = 0.16
[44].
• Counts of galaxies and of clusters of galaxies may have the same potential to probe
w as SNe Ia. The critical issue is systematics (including the evolution of the intrinsic
comoving number density, and the ability to identify galaxies or clusters of a fixed
mass) [42].
• Measuring weak gravitational lensing by large-scale structure over a field of 1000
square degrees (or more) could have comparable sensitivity to w as type Ia super-
novae. However, weak gravitational lensing does not appear to be a good method
to probe the time variation of w [45]. The systematics associated with weak gravi-
tational lensing have not yet been studied carefully and could limit its potential.
13
With the exception of vacuum energy, all the other possibilities for the dark energy
cluster to some small extent on the largest scales [47]. Measuring this clustering, while
extremely challenging, could rule out vacuum or help to elucidate the nature of the dark
energy. Hu and Okamoto have recently suggested how the CMB might be used to get at
this clustering [48].
While the Universe is likely the lab where dark energy can best be attacked, one should
not rule other approaches. For example, if the dark energy involves a ultra-light scalar
field, then there should be a new long-range force [49].
4.6 The Nancy Kerrigan Problem
A critical constraint on dark energy is that it not interfere with the formation of structure
in the Universe. This means that dark energy must have been relatively unimportant
in the past (at least back to the time of last scattering, z ∼ 1100).
If dark energy is
characterized by constant w, not interfering with structure formation can be quantified
as: w <∼ − 1
2 [36]. This means that the dark-energy density evolves more slowly than R−3/2
(compared to R−3 for matter) and implies
ρX /ρM → 0
for t → 0
ρX /ρM → ∞ for t → ∞
That is, in the past dark energy was unimportant and in the future it will be dominant!
We just happen to live at the time when dark matter and dark energy have comparable
densities. In the words of Olympic skater Nancy Kerrigan, "Why me? Why now?"
Perhaps this fact is an important clue to unraveling the nature of the dark energy.
Perhaps not. I shudder to say this, but it could be at the root of an anthropic explanation
for the size of the cosmological constant: The cosmological constant is as large as it can
be and still allow the formation of structures that can support life [51].
4.7 Dark Energy and Destiny
Almost everyone is aware of the connection between the shape of the Universe and its
destiny: positively curved recollapses, flat; negatively curved expand forever. The link
between geometry and destiny depends upon a critical assumption: that matter dominates
the energy budget (more precisely, that all components of matter/energy have equation
of state w > − 1
3). Dark energy does not satisfy this condition.
In a Universe with dark energy the connection between geometry and destiny is severed
[50]. A flat Universe (like ours) can continue expanding exponentially forever with the
number of visible galaxies diminishing to a few hundred (e.g., if the dark energy is a true
cosmological constant); the expansion can slow to that of a matter-dominated model (e.g.,
14
if the dark energy dissipates and becomes sub-dominant); or, it is even possible for the
Universe to recollapse (e.g., if the dark energy decays revealing a negative cosmological
constant). Because string theory prefers anti-deSitter space, the third possibility should
not be forgotten.
Dark energy is the key to understanding our destiny.
5 A Cosmic Perspective from Rovaniemi
After attending an exciting, enjoyable, and very productive COSMO-01, here is my sum-
mary of the state of cosmology:
• We have discerned the basic features of the Universe, as embodied in The New
Cosmology.
• We can say confidently that events that took place during earliest moments have
had a profound influence on the present state of Universe.
• We have a impressive program in place to test inflation and the particle dark-matter
hypothesis, as well as their corollary, the Cold Dark Matter theory of structure
formation.
• For some years, the front line in cosmology will continue to be phenomenology.
• No doubt, there will be surprises and we will need new ideas to replace (or supple-
ment) those that are driving cosmology today.
• Schramm's Razor -- given two equally interesting new ideas, pursue the one that's
more testable -- is more true than ever.
I am look forward to new results, new ideas and new (as well as old) faces at COSMO-
02 in Chicago.
References
[1] See e.g., M.L. Humason, N.U. Mayall, and A.R. Sandage, Astrophys. J. 61, 97 (1956).
[2] S. Weinberg, Gravitation and Cosmology (Wiley & Sons, NY, 1972).
[3] M.S. Turner, "Grand Unification and the Fundamental Problems of Classical Cos-
mology," in Second Workshop on Grand Unification, eds. J.P. Leveille, L.R. Sulak,
and D.G. Unger (Birkhauser, Boston, 1981)
15
[4] See e.g., P.J.E. Peebles, Physical Cosmology (Princeton University Press, Princeton,
NJ, 1971); or E.W. Kolb and M.S. Turner, The Early Universe (Addison-Wesley,
Redwood City, CA, 1990).
[5] P. de Bernardis et al, Nature 404, 955 (2000); S. Hanany et al, Astrophys. J. 545, L5
(2000); C.B. Netterfield et al, astro-ph/0104460; C. Pryke et al, astro-ph/0104490
[6] D. Tytler et al, Physica Scripta T85, 12 (2000); J.M. O'Meara et al, Astrophys. J.
552, 718 (2001)
[7] S. Burles et al, Phys. Rev. D 63, 063512 (2001); Astrophys. J. 552, L1 (2001)
[8] C.B. Netterfield et al, astro-ph/0104460; C. Pryke et al, astro-ph/0104490
[9] See e.g., J. Mohr et al, Astrophys. J. 517, 627 (1998)
[10] M.S. Turner, astro-ph/0106035
[11] W.J. Percival et al (2dF), astro-ph/0105252; S. Dodelson et al (SDSS), astro-
ph/0107421
[12] O. Lahav et al (2dF), astro-ph/0112162
[13] Y. Fukuda et al, Phys. Rev. Lett. 81, 1562 (1998)
[14] R. Croft, W. Hu and R. Dave, Phys. Rev. Lett. 83, 1092 (1999)
[15] L. Krauss and B. Chaboyer, astro-ph/0111597; I. Ferreres, A. Melchiorri, and J. Silk,
MNRAS 327, L47 (2001)
[16] L. Knox, N. Christensen, and C. Skordis, astro-ph/0109232
[17] W.L. Freedman et al, Astrophys. J. 553, 47 (2001)
[18] J. Mather et al, Astrophys. J. 512, 511 (1999)
[19] M.S. Turner, PASP 113, 653 (2001) (astro-ph/0102057)
[20] See e.g., B. Sadoulet, Rev. Mod. Phys. 71, S197 (1999) or K. Griest and M.
Kamionkowski, Phys. Rep. 333-4, 167 (2000)
[21] See e.g., J. Sellwood and A. Kosowsky, astro-ph/0009074
[22] See e.g., J. Bullock, A. Kravtsov and D. Weinberg, Astrophys. J. 539, 517 (2000);
R.S. Somerville, astro-ph/0107507; A.J. Benson et al, astro-ph/0108218; M.D. Wein-
berg and N. Katz, astro-ph/0110632; A. Klypin, H.-S. Zhao, and R.S. Somerville,
astro-ph/0110390; D. Merritt and F. Cruz, Astrophys. J. 551, L41 (2001); M.
Milosavljevic and D. Merritt, Astrophys. J. 563, 34 (2001); M. Milosavljevic et al,
astro-ph/0110185
16
[23] D.N. Spergel and P.J. Steinhardt, Phys. Rev. Lett. 84, 3760 (2000)
[24] M. Kaplinghat, L. Knox and M.S. Turner, Phys. Rev. Lett. 85 3335 (2000)
[25] W. Buchmuller, hep-ph/0107153
[26] J.M. Cline et al, JHEP 0007 (2000) 018; A. Cohen, D. Kaplan, and A. Nelson, Ann.
Rev. Nucl. Part. Sci. 43, 27 (1994)
[27] A. Kosowsky and M.S. Turner, Phys. Rev. D 52, R1739 (1995)
[28] M.S. Turner, Phys. Rev. D 48, 5539 (1993); J.E. Lidsey et al, Rev. Mod. Phys. 69,
373 (1997)
[29] J.K. Webb et al, Phys. Rev. Lett. 87 091301 (2001)
[30] S. Weinberg, Rev. Mod. Phys. 61, 1 (1989)
[31] http://www.livingreviews.org/Articles/ Volume4/2001-1carroll
[32] S. Perlmutter et al, Astrophys. J. 517, 565 (1999)
[33] A. Riess et al, Astron. J. 116, 1009 (1998)
[34] A. Riess et al, Astrophys. J. 560, 49 (2001)
[35] M.S. Turner and A. Riess, astro-ph/0106051 (Astrophys. J., in press)
[36] M.S. Turner and M. White, Phys. Rev. 56, R4439 (1997)
[37] S. Thomas, hep-th/0010145
[38] E. Witten, hep-th/0106109
[39] T. Banks and M. Dine, hep-th/0106276
[40] S. Barr and D. Seckel, astro-ph/0106239
[41] M.S. Turner, Physica Scripta T85, 210 (2000)
[42] See e.g., J. Newman and M. Davis, Astrophys. J. 534, L11 (2000); G.P. Holder et
al, Astrophys. J. 553, 545 (2001); S. Podariu and B. Ratra, astro-ph/0106549
[43] http://supernova.lbl.gov/∼evlinder/ sci.html
[44] D. Huterer and M.S. Turner, Phys. Rev. D 64, 123527 (2001)
[45] D. Huterer, astro-ph/0106399 (submitted to Phys. Rev. D)
17
[46] S. Perlmutter, M.S. Turner, and M. White, Phys. Rev. Lett. 83, 670 (1999)
[47] K. Coble, S. Dodelson, and J. Frieman, Phys. Rev. D 55, 1851 (1997)
[48] W. Hu and T. Okamoto, astro-ph/0111606
[49] S. Carroll, Phys. Rev. Lett. 81, 3067 (1998)
[50] L. Krauss and M.S. Turner, Gen. Rel Grav. 31, 1453 (1999)
[51] H. Martel, P. Shapiro, and S. Weinberg, Astrophys. J. 492, 29 (1998)
18
|
0802.2118 | 2 | 0802 | 2008-03-06T01:27:51 | A Long Look at the Be/X-Ray Binaries of the Small Magellanic Cloud | [
"astro-ph"
] | We have monitored 41 Be/X-ray binary systems in the Small Magellanic Cloud over ~9 years using PCA-RXTE data from a weekly survey program. The resulting light curves were analysed in search of orbital modulations with the result that 10 known orbital ephemerides were confirmed and refined, while 10 new ones where determined. A large number of X-ray orbital profiles are presented for the first time, showing similar characteristics over a wide range of orbital periods. Lastly, three pulsars: SXP46.4, SXP89.0 and SXP165 were found to be misidentifications of SXP46.6, SXP91.1 and SXP169, respectively. | astro-ph | astro-ph |
A Long Look at the Be/X-Ray Binaries of the Small Magellanic
Cloud
J.L. Galache
Harvard-Smithsonian Center for Astrophysics, 60 Garden Street, Cambridge, MA 02138
R.H.D. Corbet1
NASA Goddard Space Flight Center, Greenbelt, MD 20771
M.J. Coe
School of Physics and Astronomy, University of Southampton, Southampton SO17 1BJ, UK
Harvard-Smithsonian Center for Astrophysics, 60 Garden Street, Cambridge, MA 02138
S. Laycock
School of Physics and Astronomy, University of Southampton, Southampton SO17 1BJ, UK
M.P.E. Schurch
C. Markwardt2
NASA Goddard Space Flight Center, Greenbelt, MD 20771
F.E. Marshall
NASA Goddard Space Flight Center, Greenbelt, MD 20771
and
J. Lochner3
NASA Goddard Space Flight Center, Greenbelt, MD 20771
ABSTRACT
1University of Maryland, Baltimore County, MD 21250
2Department of Astronomy, University of Maryland, College Park, MD 20742
3Universities Space Research Association
-- 2 --
We have monitored 41 Be/X-ray binary systems in the Small Magellanic
Cloud over ∼ 9 years using PCA-RXTE data from a weekly survey program.
The resulting light curves were analysed in search of orbital modulations with
the result that 10 known orbital ephemerides were confirmed and refined, while
10 new ones where determined. A large number of X-ray orbital profiles are
presented for the first time, showing similar characteristics over a wide range
of orbital periods. Lastly, three pulsars: SXP46.4, SXP89.0 and SXP165 were
found to be misidentifications of SXP46.6, SXP91.1 and SXP169, respectively.
Subject headings: galaxies: individual (Small Magellanic Cloud) -- pulsars: gen-
eral -- X-rays: binaries
1.
Introduction
The Small Magellanic Cloud (SMC) has become a fertile orchard of High-Mass X-
ray Binaries (HMXBs), with 49 confirmed systems (Coe et al. 2005; McGowan et al. 2007).
Given that, from extrapolation of the Milky Way's population (and even correcting for the
higher Be/B ratio in the SMC (Maeder et al. 1999)), one would expect to find only 3 -- 4
systems, it is clear that the SMC is a special place where recent star formation has provided
an abundance of HMXBs; indeed, the SFR/M (Star Formation Rate/Mass) of the SMC is 150
times that of the Milky Way (Grimm et al. 2003). Of particular significance is the fact that
only one of these binary systems is not a Be/X-ray binary (SMC X-1 is the sole supergiant
system discovered so far). This large number of Be/X-ray binary systems, conveniently
located within the 3◦× 3◦area of the SMC, provides an unrivaled opportunity to study this
population as a whole, as well as individually. With a 2◦ FWZI field of view, high timing
resolution (1 µs), and sensitive enough to detect the 1036 -- 1038 erg s−1 luminosities typical of
these systems when in outburst, the PCA instrument (Jahoda et al. 1996, 2005) on board
RXTE is well suited for a long-term monitoring survey of the SMC.
The different types of X-ray activity displayed by Be/X-ray transient systems were
classified by Stella et al. (1986) into the following categories:
-- Persistent low-luminosity X-ray emission (Lx . 1036 erg s−1) or none detectable (in
which case the system is said to be in quiescence).
-- Type I outbursts: Outbursts of moderate intensity (Lx ≃ 1036 -- 1037 erg s−1) and short
duration (a few days) generally recurring with the orbital period of the system and
taking place at, or close to the time of periastron passage.
-- 3 --
-- Type II outbursts: Giant X-ray outbursts (Lx & 1037 erg s−1) lasting for weeks or
months that generally show no correlation with orbital phase.
The data presented in this paper spans November 1997 -- November 2006, and builds
upon the work of Laycock et al. (2005), who analysed the first 4 years of data. Following is
a brief description of the survey so far, which is still ongoing as of June 2007.
1.1. The survey
The initial observations of the SMC with RXTE began in 1997. The first observation
took place in November of that year when an outburst detected by the ASM was missiden-
tified as SMC X-3. It was RXTE's second year of operation, and only SMC X-1, 2 and 3,
SXP2.76 and SXP8.88 were known. From these initial observations it soon became apparent
that there were more than 5 X-ray pulsars in the SMC. The observations carried out within
the next year brought about the discovery of 5 new systems: SXPs 46.6, 59.0, 74.7, 91.1 and
169. Other pulsars were also discovered or detected with EINSTEIN, ASCA, BeppoSAX and
ROSAT.
1999 marked the beginning of a coordinated survey of the SMC using the PCA. The
PCA's 2◦ field of view provides coverage of a wide area of the SMC, which allows many
pulsars to be monitored with just one pointing. A number of different pointing positions
have been used throughout the years and are given in Table 1 (see also Fig. 1); some of
the less observed ones never received a name and are omitted from the table. The most
frequently observed is Position 1 (later renamed to A), which has been the main pointing
position since AO4, except during AO6 and AO7, when Position 5 was the main target. In
total, the collected data spans ∼ 9 years.
The survey has gone through various phases characterised by different observing posi-
tions and/or observing modalities. Phases 1 -- 4 have already been described in Laycock et al.
(2005) but they are outlined here again, together with the two latest phases not included in
previous studies.
Phase 1 (AO2 -- AO3): These observations used Positions 1a, 1b and 1c and are described
in Lochner et al. (1999a,b). Only 30 observations were carried out, their main purpose
being to monitor the 5 newly discovered pulsars in those regions.
Phase 2 (AO4): Positions 1 -- 4 were defined; Position 1 overlaps most of Positions 1a -- c
and contained the new pulsars, Positions 2 and 4 cover the rest of the bar of the SMC
-- 4 --
while Position 3 covers the area of the wing containing SMC X-1. A continued survey
began at this point in time, with ∼ 3 ks observations being made once a week, mostly
of Position 1, with some occasional looks at the other positions.
Phase 3 (AO5 -- AO6): Only Position 5 was monitored, and was done weekly so as not to
create gaps in the data; the majority of the most active systems located in Position 1
also fell within the field of view of Position 5. Time allotted for the project was
increased to an average of ∼ 5 ks per observation, thus providing better sensitivity to
longer period pulsations.
Phase 4 (AO7 -- AO9): Weekly monitoring returned to Position 1, now renamed Position A,
with additional observations of the other positions (B, C and D, which are very close
to 2, 3 and 4, respectively) being made once a month. These monthly pointings were
∼ 15 ks, while the weekly ones were increased to ∼ 6 -- 7 ks.
Phase 5 (AO10): The time available for the monthly observations of alternate positions
was invested into increasing the length of the weekly observations of Position A to
∼ 10 ks, this being the only monitored position.
Phase 6 (AO11): Having mainly monitored the central bar of the SMC, it was decided to
move to another location at the northeastern tip of the bar, near position B. This new
location, Position X, was monitored for ∼ 10 ks weekly for 18 weeks, but was abandoned
due to lack of pulsar activity. Observations then changed to Position D.
1.2. Data reduction
Cleaning of the raw light curves was achieved via a pipeline employing various FTOOLS1
routines (Blackburn 1995). Initial filtering required the data to come from observations offset
from the target no more than 0.004◦, with an elevation above the Earth's horizon of more
than 5◦, and not taken during times of high background. Good Xenon mode data from the
top anode layer only were used, within an energy range of 3 -- 10 keV. Data were binned at
0.01 s intervals, while background files were generated in 16 s bins. For each time-step, the
net flux was divided by the number of active PCUs to give the counts PCU−1 s−1. Then,
the light curves time tags were corrected to the Solar System barycenter, removing timing
variations caused by the satellite's motion around the Earth and Sun. Finally, short light
1http://heasarc.gsfc.nasa.gov/ftools/
-- 5 --
Fig. 1. -- Map of the SMC H I distribution with the 5 main pointing positions of RXTE
during the survey shown as numbered white circles. For each pointing, the inner circle has
a diameter of 1◦, the outer of 2◦. For clarity, Positions A, B, C, D and X have not been
plotted, as they are very close to Positions 1, 2, 3, 4 and 2, respectively. Pulsars with known
positions are marked by small circles. H I data from Stanimirovic et al. (1999).
-- 6 --
curves belonging to the same observation (which had been split up due to SAA passage,
Earth occultation or flares) were pieced together into a long light curve spanning the whole
observation.
1.3. Collimator correction
The collimator of each PCU consists of a number of hexagonal cells joined in a hon-
eycomb structure. The hexagonal tubes are not perfectly parallel with the result that all
the PCUs are slightly off center. Hence, the resulting collimator response pattern has an
elliptical hexagon shape with the maximum throughput being slightly off center.
In order to account for differences in observed flux from a pulsar when observed at
different pointing positions, a collimator correction was applied a posteriori to each pulsar's
count rate. A look-up table approach was used, where each pulsar had a collimator response
calculated for each of the pointing positions used in the survey (see Table 1).
A pulsars with unknown position cannot be collimator corrected, so significant detec-
tions may not rise above the noise level in the long-term light curve unless the outbursts are
very bright and/or the pulsar is located close to the center of the field of view. Type I out-
bursts may only appear bright when the pulsar is close to the center of the field of view (as
with SXP59.0 in Position 1/A, Fig. 17), in which case the collimator correction is small any-
way. To overcome this limitation pulsars with unknown positions had coordinates "guessed"
for them on the basis of which observing positions they had been detected in throughout the
mission; the coordinates assigned were approximately at the center of the region formed by
the overlap of the observing positions in which it had been previously detected.
1.4. Data analysis
Once the cleaned light curve for an observation was obtained, it was run through a timing
analysis package using the Lomb-Scargle method (Lomb 1976; Scargle 1982) and numerically
implemented following the prescription by Press & Rybicki (1989). Two different power
spectra were generated for each light curve, each spanning different period spaces (from Pmin
to Pmax), at different timing resolutions (∆f ), and pulsars were later searched for in their
corresponding group according to their pulse period. The parameters used for each group
are listed in Table 2. The reason for creating two groups was to obtain densely sampled
periodograms at long periods without creating excessively large files. The parameters were
chosen such that both groups would contain approximately the same number of independent
-- 7 --
frequencies, M.
Once the periodograms were calculated, it was necessary to establish an objective way
to determine which peaks in the power spectra were real; this is covered in Appendix A.
1.4.1. Pulsar Detection
Pulsars were searched for within the data using an algorithm specifically created for the
task. Each light curve's two power spectra were scanned to look for the pulsars in groups 1
and 2; the steps used were as follows:
Step 1: The highest peak (at a period P ) above 90% global significance in the power spec-
trum is found and, either identified as a known pulsar, or flagged as an "unknown".
Step 2: The light curve is folded at period P to produce a pulse profile.
Step 3: Using the pulse profile as a template, the pulsations are then subtracted from the
light curve and a power spectrum of the cleaned light curve is produced.
Step 4: This power spectrum is subtracted from the previous one to create a pulsar-specific
power spectrum (or P2S2), which shows only the contribution of the individual pulsar
to the power spectrum. This method allows one to see the possible harmonics that
may have been lost in the noise or confused with the fundamental of another pulsar.
The power of the fundamental and harmonic(s) peaks in the P2S2 is measured and a
pulse amplitude estimated using Eq. (A5). The significance of the detection is taken
to be the local significance of a peak of power equal to that which the detected signal
would have if the power in the harmonics were concentrated in the fundamental.
Step 5: Repeat previous steps until all signals with peaks above 90% significance have been
removed.
Step 6: To account for the remaining, dim pulsars, a stretch of the power spectrum centered
on each pulsar's nominal frequency, and with a total width which is different for each
pulsar (and depends on the pulsar's past period history), is searched for a peak. If no
peak is found, then the significance of the detection is set to 0% and the amplitude to
that of the average power within the region. If there is a peak, then the significance of
the detection is set to the local significance and the amplitude of the pulsar calculated
from the power in the peak according to Eq. (A4). No harmonics are searched for in
this case.
-- 8 --
Table 1. RXTE's SMC survey pointing positions
Name
RA (◦)
Dec (◦)
1a
1b
1c (SMC X-3)
1/A
2
3
4
5
B
C
D
X
SMC X-2
00 52 07.8 −72 25 43.3
00 51 04.0 −72 13 44.0
00 54 54.8 −72 26 40.9
00 53 53.0 −72 26 42.0
01 05 00.0 −72 06 00.0
01 15 00.0 −73 06 00.0
00 50 44.6 −73 16 04.8
00 50 00.0 −73 06 00.0
01 05 00.0 −72 06 21.6
01 15 00.0 −73 24 59.8
00 50 00.0 −73 06 00.0
01 05 00.0 −72 06 00.0
00 54 33.3 −73 41 04.2
Table 2. Parameters used for the Lomb-Scargle periodogram
Group 1 Group 2
Pmin (s)
Pmax (s)
∆f (Hz)
Search range (s)
0.5
1000
10−5
0.5 -- 40
10
3000
5 × 10−7
40 -- 3000
-- 9 --
1.4.2. Light Curve Generation and Porb Calculations
Once the amplitude of the pulsed flux for every pulsar in every observation had been
measured, a long-term light curve was created for each system showing its X-ray activity
over the course of the survey. It is important to note that these light curves do not show the
absolute flux, but rather the pulsed flux. Without knowing the pulsar's pulse fraction for
each observation it is impossible to convert the pulsed flux amplitude into an absolute flux
value. These light curves were then searched for periodicities, as the X-ray emission could
show orbital modulations. Again, the Lomb-Scargle technique was employed, searching
for periods in the appropriate range for each pulsar.
In cases where Type II outbursts
contaminated the Lomb-Scargle periodogram, these were removed from the data. In general,
all bright outbursts were initially removed and only added back into the light curve if they
coincided with the calculated ephemerides and their inclusion increased the significance of
the calculated orbitla period. It is important to note that the collimator response correction
was only applied to the data points whose local significance upon detection was ≥ 99%,
anything below this detection threshold was considered noise, and as such was not collimator
corrected. § 2 shows the results of these analyses.
1.4.3. Orbital Profile Generation
Based on the ideas of de Jager et al. (1988) and Carstairs (1992), we used the Phase-
Independent Folding (PIF) technique to create the orbital profiles. The method is as follows:
the folded light curve is obtained from m sets of n-binned folded light curves. To begin with,
the light curve is folded at the desired period and the time values converted to phase space
(ranging from 0 to 1). This "raw" folded light curve is then binned into n bins (of width
1/n) in the standard way, with each bin starting at phase a/n (with a = 0, 1, 2, ...n − 1).
This step is repeated again, but this time the bins start at phase a/n + 1/(n × m). In this
way we create m folded light curves from the same data, each consisting of n bins, and only
differing in the starting phase of these bins2. The general expression for the phase at which
each bin begins is a/n + m/(n × m). Now each folded light curve is further divided into
l = n × m sub-bins, such that in each light curve there will be n groups of m consecutive
bins with the same flux value. The final folded light curve will have l bins, each of which is
the average of the bins from the m sets of light curves. The error in each bin will be given
by the standard error calculated from the m values that were averaged for each l-bin. In the
present work we have used n × m = 10 × 5.
2We define φ = 0 at the time of the first data point in each observation.
-- 10 --
The PIF method provides an efficient way of generating folded light curves from poorly
sampled data as their shape will not depend on the starting point at which they are folded
and, although only every m th bin will be independent, spurious flux values within bins will
be evened out while real features will remain. Thus, we obtain the benefits of both wide
bins (sufficient counts in each bin for good statistics) and narrow bins (sensitivity to narrow
features in the profile).
2. Results
Following are the results obtained from the light curves of the observed SMC X-ray
pulsars.
In the triple-panel plots (generally the (a) plots), the top panel shows the X-
ray activity of each pulsar through the amplitude of the pulsed flux, with each solid line
representing one observation. When an orbital period has been found, the ephemerides for
the dates of maximum flux are over plotted as dashed vertical lines; where the orbital period
is less than 25 days, only every other line is plotted for clarity. The middle panel shows the
period at which the pulsar was detected, with the horizontal dashed line denoting the center
of the pulsar's search range (note that only significant detections have their periods plotted).
Finally, the bottom panel displays the significance of the neutron star's pulsations for each
observation, with the three horizontal dashed lines marking, from bottom to top, the levels
of 99, 99.99 and 99.9999% local significance. We considered a pulsar had been detected when
its local significance was ≥ 99%. In the (b) plots we show the Lomb-Scargle periodogram of
the pulsed flux light curve (top) and the light curve folded at the orbital period (bottom).
When horizontal dashed lines are plotted on the periodogram, these represent different levels
of significance, which are (from bottom to top) 99, 99.9, 99.99, and 99.999%. The coordinates
given for each system are the most accurate at time or writing. When a Be/X-ray system
has a confirmed optical counterpart, the coordinates of the counterpart are used. When no
optical counterpart is known, but the system has been detected by Chandra and/or XMM,
then it is the X-ray coordinates that were given. For some of the pulsars without an exact
position, the coordinates provided by scans with RXTE are given; these are the least precise
and have errors of up to several arc minutes.
2.1. SXP0.92
PSR J0045−7319
RA 00 45 35, dec −73 19 02
-- 11 --
History: First discovered as a radio pulsar by Ables et al. (1987) with a period of 0.926499± 0.000003 s
using the Parkes 64 m radio telescope. Kaspi et al. (1993) observed Doppler shifts in the
pulse period which where consistent with a 51 d binary orbit with a companion star hav-
ing mass > 4 M⊙. Optical observations of the field revealed a 16th magnitude, 11 M ⊙, B1
main-sequence star, which is likely the companion (Bell 1994).
Survey Results: SXP0.92 received ∼ 2 years of uninterrupted coverage during AO5 and
AO6 (see Fig. 2), during which time it was only once detected above 99% local significance
(MJD 52334). The power spectrum does not show any significant periods.
2.2. SXP2.16
XTE SMC2165 RA 01 19 00, dec −73 12 27
History: Discovered during the course of this survey by Corbet et al. (2003a) at 2.1652± 0.0001 s.
Survey Results: Due to its position near SMC X-1 it has only been within the field of view
on 3 occasions: MJD 51220, 51263 and 51310. It was on this last date that it was discovered
(its only significant detection).
2.3. SXP2.37
SMC X-2
RA 00 54 34, dec −73 41 03
History: SMC X-2 was discovered in SAS-3 observations of the SMC carried out by
Clark et al. (1978). Further outbursts were also observed by HEAO 1 and ROSAT, but
no pulsations were detected until it was observed in the present survey by RXTE during
a long outburst lasting from January through to May, 2000 (Corbet et al. 2001; Laycock
2002).
Survey Results: After the aforementioned outburst, SMC X-2 was only detected on 3
further occasions (MJD 51974, 52025 and 52228), but at a much lower flux of F xpul <
1 counts PCU−1 s−1(see Fig. 3). Lomb-Scargle analysis of the data following the major out-
burst shows no clear periods.
-- 12 --
Fig. 2. -- SXP0.92, X-ray amplitude light curve.
-- 13 --
Fig. 3. -- SXP2.37, X-ray amplitude light curve.
-- 14 --
2.4. SXP2.76
RX J0059.2−7138
RA 00 59 11.7, dec −71 38 48
History: First reported by Hughes (1994) from a ROSAT observation showing 2.7632 s
pulsations that varied greatly with energy. In the low-energy band of the ROSAT PSPC
(0.07 -- 0.4 keV) the source appears almost unpulsed while in the high-energy band (1.0 --
2.4 keV) the flux is ∼ 50% pulsed. Its optical counterpart was confirmed as a 14th magnitude
Be star by Southwell & Charles (1996). Schmidtke et al. (2006) report a period of 82.1 d
with a maximum at MJD 52188.9 from analysis of OGLE III data.
Survey Results: This source is near the edge of the field of view of Position 1/A so we
would expect to detect only the brighter outbursts. Only 2 detections were made (MJD
52527 and 53327) and no orbital period can be extracted from the X-ray data (see Fig. 4).
We find no power at the reported optical period and note that the two X-ray detections
occured ∼ 10 days after and ∼ 11 days before the optical ephemeris's predicted maximum.
2.5. SXP3.34
AX J0105−722, RX J0105.3−7210
RA 01 05 02, dec −72 11 00
History: Was reported as an ASCA source with 3.34300± 0.00003 s pulsations by Yokogawa & Koyama
(1998c). Its optical counterpart is proposed to be [MA93] 1506 (Coe et al. 2005), who also
find an 11.09 d modulation in MACHO data. Although this could be the orbital period of
the system (as it falls within the expected range on the Corbet diagram (Corbet 1984)),
Schmidtke & Cowley (2005a) report the find of a strong 1.099 d period in MACHO data and
that the 11.09 d value is an alias of this main period. They attribute the modulation to
non-radial pulsation in the Be star.
Survey Results: There have been no significant detections of the pulsar during this survey
and timing analysis of the light curve reveals no clear periodicities.
2.6. SXP4.78
XTE J0052−723
RA 00 52 06.6, dec −72 20 44
-- 15 --
Fig. 4. -- SXP2.76, X-ray amplitude light curve.
-- 16 --
History: Was discovered by the present survey in late December 2000 and reported in
Laycock et al. (2003), where [MA93] 537 was proposed as the optical counterpart. Another
possible counterpart is suggested in Coe et al. (2005) as the star AzV129 is found to have
a 23.9± 0.1 d period in both MACHO colours, which would agree with the expected orbital
period inferred from the Corbet diagram.
Survey Results: SXP4.78 was detected on one occasion after its initial outburst (MJD
52729) at a much weaker F xpul≃ 0.8 counts PCU−1 s−1.
It then began a relatively long,
bright (peaking at ∼ 1.2 counts PCU−1 s−1) outburst on December 21, 2005 (MJD 53725)
that lasted ∼ 7 weeks (see Fig. 5). Despite the long outburst, no orbital modulation is
apparent. Timing analysis using all the data finds no periods, while analysis of the data
outside the two bright outbursts suggests a weak period at 34.1 d
A 1 ks observation with Chandra was carried out on March 3rd 2006 in an attempt to
establish the pulsar's coordinates. Unfortunately, the outburst had ended and no source was
detected within the RXTE error box provided by Laycock et al. (2003).
2.7. SXP6.85
XTE J0103−728
RA 01 02 53.1, dec = −72 44 33
History: First detected in April 2003 with a pulse period of 6.8482± 0.0007 s (Corbet et al.
2003b). It was detected in outburst with XMM on October 2nd 2006, which provided a more
accurate position (Haberl et al. 2007). This allowed its optical counterpart to be identified
as a Be star with V = 14.59, with a possible 24.82 d periodicity (Schmidtke & Cowley 2007).
Survey Results: It has been detected on 3 other occasions (circa MJD 52885, 53440 and
53677) at varying fluxes (F xpul ≈ 0.8 -- 1.8 counts PCU−1 s−1), but high above the noise level
(see Fig. 6). Lomb-Scargle analysis of the data outside the 4 bright detections shows no
significant period. When including these outbursts, we find a significant period of 112.5 d;
this is the minimum time lapse between any two outbursts and probably drives the result.
Given its pulse period, and based on the spin-orbit relation, we do not discard the possibility
that the real orbital period be 1
3 of this value. The current X-ray ephemeris is MJD
52318.5± 7.9 + n× 112.5± 0.5 d. We note that the outburst detected on October 2nd 2006
(MJD 54010) is consistent with the proposed ephemeris.
2 or 1
-- 17 --
Fig. 5. -- SXP4.78, X-ray amplitude light curve.
-- 18 --
Fig. 6. -- SXP6.85. a) Top: X-ray amplitude light curve. b) Middle: Lomb-Scargle power
-- 19 --
2.8. SXP7.78
SMC X-3
RA 00 52 05.8, dec −72 26 03.2
History: Originally detected by Clark et al. (1978), it was not identified with the previously
detected RXTE 7.78 s source until 2004 (Edge et al. 2004a). Corbet et al. (2003c, 2004b) pro-
posed an orbital period from a series of recurrent X-ray outbursts (in the present survey) of
45.1± 0.4 d. An optical modulation in MACHO data was reported by Cowley & Schmidtke
(2004) (44.86 d) and Coe et al. (2005) (44.6± 0.2 d); Edge (2005d) also found a strong
44.8± 0.2 d modulation in the OGLE counterpart, present even when there was no significant
X-ray activity.
Survey Results: It has displayed in the past 10 years 3 distinct periods of outbursts (F xpul ≈
0.3 counts PCU−1 s−1), lasting ∼ 200 -- 400 d (see Fig. 7(a)). Timing analysis of the complete
light curve reveals a clear period at 44.92 d; the ephemeris we derive is MJD 52250.9± 1.4
+ n× 44.92± 0.06 d. The folded light curve may show evidence of detections at apastron.
During the longer of the outburst episodes (circa MJD 52500) the pulsar shows some spin
up, about P = 3.7 × 10−10 s s−1, implying Lx ≥ 2.3× 1037 erg s−1 (B ≤ 3.6× 1012 G). This
pulsar is unique in that, despite the spin up observed during each of the individual outburst
episodes, the overall spin evolution seems to show a long-term spin down.
2.9. SXP8.02
CXOU J010042.8−721132
RA 01 00 41.8, dec −72 11 36
History: Proposed as the first anomalous X-ray pulsar (AXP) in the SMC by Lamb et al.
(2002), they found the source to have displayed little variability in the past 20 years. It is
characterised by a very soft spectrum and low luminosity (∼ 1.5× 1035 erg s−1).
Survey Results: During AO5 and AO6 it was outside the field of view of our observations.
Coverage from AO7 onwards has been good and timing analysis of these dates shows a
possible periodicity at ∼ 23.2 d, which is likely driven by the only two clear detections,
separated by that time range (see Fig. 8). This period disappears when removing these two
detections from the data. If SXP8.02 is truly an anomalous pulsar, it is not expected to show
periodicity in its X-ray emission. The two significant detections were observed at around Lx
= 5.0× 1036 erg s−1 (unabsorbed, assuming a 50% pulse fraction and a power law spectrum
of γ = 1).
-- 20 --
Fig. 7. -- SXP7.78. a) Top: X-ray amplitude light curve. b) Middle: Lomb-Scargle power
-- 21 --
Fig. 8. -- SXP8.02, X-ray amplitude light curve.
-- 22 --
2.10. SXP8.88
RX J0051.8−7231, 1E0050.1−7247, 1WGA J0051.8−7231
RA 00 51 52.0, dec −72 31 51.7
History: Israel et al. (1997) detected this source several times between 1979 and 1993. A
number of optical counterparts have been proposed, and Haberl & Sasaki (2000) believe
[MA93] 506 to be the correct one. From a number of outbursts during July 2003 -- May 2004
Corbet et al. (2004a) derived an orbital ephemeris of MJD 52850± 2 + n× 28.0± 0.3 d.
Schmidtke & Cowley (2006) find an optical period of 33.4 d in OGLE and MACHO data.
Note: This system was incorrectly named SXP8.80 in Coe et al. (2005).
Survey Results: Two Type II outbursts together with the aforementioned series of Type I
outbursts have been detected (see Fig. 9). A period of ∼ 28 d is found using Lomb-Scargle
analysis on the data not containing the Type II outbursts, similar to the value from Corbet et al.
(2004a). The new ephemeris from the survey data is MJD 52392.2± 0.9 + n× 28.47± 0.04 d.
We note that the two Type II outbursts began roughly at the times of expected maximum
flux, lasted about 1 orbit, and peaked 0.5P orb later. This could be explained by a number of
scenarios: a) The neutron star forms an accretion disc around periastron, which it consumes
throughout the orbit; b) an accretion disc is formed around periastron which is accreted
onto the neutron star along the orbit at the same time as wind accretion (which would peak
around apastron) is taking place; c) the Be star ejects matter forcefully enough that it moves
outwards as an annulus and the neutron star is able to capture material along its orbit as it
moves away from the Be star, but capture decreases as soon as it passes apastron.
None of these scenarios completely explain the behaviour of the X-ray emission during
the Type II outbursts. If a) is correct, then why would accretion become greatest at apastron
on both occasions. Scenario c) suffers a similar drawback as it would require the annulus
to move at similar speeds on both outbursts in order for the outburst to peak precisely
at apastron. Scenario b) would require a very dense wind in order for its accretion to be
comparable to disc accretion. Simultaneous observations of this system in the X-ray and the
optical during an outburst would help clarify this behaviour.
2.11. SXP9.13
AX J0049−732, RX J0049.2−7311
RA 00 49 13.6, dec −73 11 39
History: Discovered during an ASCA observation in November 1997 (MJD 50765), pulsa-
-- 23 --
Fig. 9. -- SXP8.88. a) Top: X-ray amplitude light curve. b) Middle: Lomb-Scargle power
-- 24 --
tions were detected at 9.1321± 0.0004 s and Lx = 3.3× 1035 erg s−1 (Imanishi et al. 1998).
There has been much debate as to which is the correct optical counterpart to this source,
Schmidtke et al. (2004) and Filipovi´c et al. (2000) identify it with the ROSAT source RX
J0049.5−7310 but Coe et al. (2005) conclude that it is a Hα source coincident with RX
J0049.2−7311. Timing analysis of the OGLE data shows a peak at 40.17 d and its probable
harmonic at 20.08 d (Edge 2005d). It should be noted that Schmidtke et al. (2004) find a
91.5 d or possibly an ∼ 187 d period for RX J0049.5−7310. It was detected by ASCA on one
further occasion on MJD 51645 at Lx = 1.9× 1035 erg s−1 (Yokogawa et al. 2003).
Survey Results: Although coverage of this source has been very complete, it has only been
detected in outburst 3 times (circa MJD 53545, 53700 and 53780). Lomb-Scargle analysis
of data prior to this date shows power at a period of 77.4 d; when analysing the whole data
set, the most significant peak is at 77.2 d (Fig. 10(b)). In both cases the ephemerides agree
with the 3 outbursts (Fig. 10(a)), which strongly suggests this could be the orbital period
of the system. However, it should be noted that this value is different from the reported
optical periods. The ephemeris we derive is MJD 52380.5± 2.3 + n× 77.2± 0.3 d. The above
reported ASCA detections did not find the system in outburst and cannot be used to ratify
this ephemeris.
2.12. SXP15.3
RX J0052.1−7319
RA 00 52 14, dec −73 19 19
History: Lamb et al. (1999) found 15.3 s pulsations in ROSAT and BATSE data from 1996.
They estimate the luminosity to be ∼ 1037 erg s−1 with a pulse fraction of ∼ 27%. Edge
(2005d) finds an ephemeris of MJD 50376.1 + n× 75.1± 0.5 d describes the modulation in
the MACHO and OGLE light curves. It should be noted that this ephemeris is likely driven
by the large 1996 X-ray outburst which is clearly visible in the OGLE data; this Type II
outburst peaked around MJD 50375.
Survey Results: There was a very minor detection of SXP15.3 in February 2000 (MJD
51592), and a more significant one in August 2003 (MJD 52883) at F xpul ≃ 1.6 counts PCU−1 s−1
(see Fig. 11). Two more bright detections separated by 13 days occurred in March 2005 (circa
MJD 53445); then, on July 12th (MJD 53564) a very bright outburst began, lasting until
October 17th (MJD 53661, ∼ 100 days long). During this time the pulsed flux oscillated
between ∼ 2.0 and ∼ 5.6 counts PCU−1 s−1. A clear spin up is detected up until September
15th, when the period started increasing. The maximum period change was ∆P = 3.3× 10−2
-- 25 --
Fig. 10. -- SXP9.13. a) Top: X-ray amplitude light curve. b) Middle: Lomb-Scargle power
-- 26 --
over the course of 56.68 d, giving a P = 6.74× 10−9 s s−1. We derive the expected luminosity
(see Appendix B for method) from such a level of spin up (if it were all intrinsic with no
orbital contribution) to be Lx ≥ 8.6× 1037 erg s−1 (B ≤ 1.5× 1013 G).
It is likely this outburst lasted for more than one orbital cycle (expected to be ∼ 30 --
50 d from the Corbet diagram) so some orbital modulation might be visible in the period
data. An attempt was made to fit these data to an orbital model with constant global spin
up, and also with a piece-wise approach where the spin up varies throughout the outburst
(following the method outlined in Wilson et al. (2003)). Although no definite value was
found, variations in the period curve suggest a period of ∼ 28 d.
Timing analysis of the data outside the long outburst revealed no periodicities. Fur-
thermore, the optical ephemeris proposed by Edge (2005d) does not agree with any of the
detections for this source.
2.13. SXP16.6
XTE J0050−731
No position available
History: Discovered with a deep 121 ks observation taken for this survey in September
2000. It was initially misidentified as RX J0051.8−7310 but later disproved Yokogawa et al.
(2002), and so SXP16.6 remains unassociated with any known source, although it is often
still mistakenly referred to as RX J0051.8−7310.
Survey Results: There have been a large number of detections of SXP16.6, and Lomb-
Scargle analysis finds a strong modulation at 33.72 d (see Fig. 12(b)), which we propose as
the orbital period of the system. The ephemeris is MJD 52373.5± 1.0 + n× 33.72± 0.05 d.
2.14. SXP18.3
XTE J0055−727
RA 00 50, dec −72 42
History: Reported by Corbet et al. (2003e) from observations in November/December
2003. An approximate position was established from scans with RXTE's PCA (Corbet et al.
2003e). No optical counterpart has yet been identified. A number of further outbursts be-
tween May -- October 2004 provide an ephemeris of MJD 53145.7± 1.3 + n× 34.6± 0.4 d
from O-C (observed vs. calculated) orbital calculations (Corbet et al. 2004d).
-- 27 --
Fig. 11. -- SXP15.3, X-ray amplitude light curve.
-- 28 --
Fig. 12. -- SXP16.6. a) Top: X-ray amplitude light curve. b) Middle: Lomb-Scargle power
-- 29 --
Survey Results: A long, bright outburst began around MJD ∼ 53925 (July 2006), which
was ongoing as of January 2007 (see Fig. 13(a)). Lomb-Scargle analysis of data prior to this
outburst finds a peak at 17.73 d, which is ∼ 1
2 the period proposed by Corbet et al. (2004d).
We find no significant power at this period and note that some of the significant detections
do not agree with it. Thus, the new ephemeris for this system is MJD 52275.6± 0.9 +
n× 17.73± 0.01 d.
2.15. SXP22.1
RX J0117.6−7330
RA 01 17 40.5, dec −73 30 52.0
History: Discovered with ROSAT as an X-ray transient in September 1992 (Clark et al.
1996), its companion was identified as a 14.2 magnitude OB star by Charles et al. (1996). It
was observed with ROSAT again in October, but no pulsations were detected (Clark et al.
1997) The companion's classification as a Be (B1-2) star was confirmed by Coe et al. (1998).
22 s pulsations were finally detected in ROSAT and BATSE data by Macomb et al. (1999).
Survey Results: There are only 3 observations of this pulsar's region and it was not
significantly detected in any of them.
2.16. SXP31.0
XTE J0111.2−7317, AX J0111.1−7316
RA 01 11 09.0, dec −73 16 46.0
History: Was simultaneously discovered by RXTE and BATSE during a giant outburst that
began late October 1998 (Chakrabarty et al. 1998; Wilson & Finger 1998). Schmidtke et al.
(2006) report a 90.4 d periodicity in OGLE III data.
Survey Results: Being in the same field as SMC X-1, it was only observed 3 times, the first
during the end of the aforementioned giant outburst. The two detections are ∼ 43 days apart
and a spin up of 0.4 s is measured, which implies a luminosity of Lx ≥ 2.7× 1038 erg s−1. As
Laycock et al. (2005), in an in depth spectral analysis of these observations, derive a value of
Lx ≥ 4.6× 1037 erg s−1, we can assume we are seeing Doppler shifted periods due to orbital
motion. No further information can be derived from these observations.
-- 30 --
Fig. 13. -- SXP18.3. a) Top: X-ray amplitude light curve (for clarity, only every other
-- 31 --
2.17. SXP34.1
RX J0055.4−7210
RA 00 55 26.9, dec −72 10 59.9
History: Discovered in archival Chandra data at 34.08± 0.03 s and lying 0.6 arcsec from a
known ROSAT source (Edge et al. 2004d,c).
Survey Results: Only two significant detections are seen (MJD 50777 and 53690) and no
clear periods can be found with timing analysis (see Fig. 14).
2.18. SXP46.6
XTE J0053−724, 1WGA 0053.8−7226
RA 00 53 58.5, dec −72 26 35
History: Discovered in the first observation of this survey (November 25th 1998) with
a period of 46.63± 0.04 s while it was undergoing a long outburst (Corbet et al. 1998).
Laycock et al. (2005) derive a period of 139± 6 d from 6 X-ray outbursts in the earlier part
of this survey. Two candidates for the optical counterpart were proposed by Buckley et al.
(2001). Schmidtke et al. (2007) confirm it is star B and find two periods in OGLE data:
69.2± 0.3 d and 138.4± 0.9 d; they suggest the possibility that the orbital period is the
shorter or the two values.
Survey Results: The source was thought to be inactive after January 2002. In the mean-
time a new SMC pulsar with a 46.4 s period was announced (Corbet et al. 2002). Lomb-
Scargle analysis of both pulsars revealed the same orbital periods and very similar ephemeris,
suggesting they were the same pulsar which has been slowly spinning up (Galache et al.
P = 1.05× 10−9 s s−1 during MJD
2005). The estimated luminosity required for a spin up of
50800 -- 51900 is Lx ≥ 9.9× 1035 erg s−1 (B ≤ 6.0× 1012 G), and for a P = 4.68 × 10−10 s s−1
during MJD, Lx ≥ 4.4× 1035 erg s−1 (B ≤ 4.0× 1012 G). The ephemeris of the outbursts is
now best described by MJD 52293.9± 1.4 + n× 137.36± 0.35 d (see Fig. 15(a)). Although
the folded light curve shows a small increase in flux half a phase away from maximum, we
do not believe this is evidence that the orbital period is half the calculated value. Given
the clear, regular outbursts experienced by this system throughout the survey we believe the
true orbital period is ∼ 137 d.
-- 32 --
Fig. 14. -- SXP34.1, X-ray amplitude light curve.
-- 33 --
Fig. 15. -- SXP46.6. a) Top: X-ray amplitude light curve. b) Middle: Lomb-Scargle power
-- 34 --
2.19. SXP51.0
SMC 51
No position available
History: Was erroneously proposed as a new 25.5 s pulsar in Lamb et al. (2002) from a
deep 121 ks observation. Laycock (2002) identifies the 25.5 s peaks in the power spectrum as
harmonics of SXP51.0's true pulse period. No position is available; it lies within Position 4,
and most likely in the overlap between Positions 1/A and 4.
Survey Results: Despite numerous significant detections (see Fig. 16), timing analysis
finds no strong periodicities in the light curve.
2.20. SXP59.0
XTE J0055−724, RX J0054.9−7226, 1WGA J0054.9−7226
RA 00 54 56.6, dec −72 26 50
History: Discovered in RXTE observations of the vicinity of SMC X-3 (Marshall et al.
1998), it showed 4 bright, very similar, outbursts during January 1998 -- September 1999 from
which Laycock et al. (2005) derived an orbital period of 123± 1 d. The optical counterpart
was established by Stevens et al. (1999); Schmidtke & Cowley (2005b) found a 60.2± 0.8 d
period from timing analysis of OGLE and MACHO data which they propose as the orbital
period.
Survey Results: SXP59.0 remained undetected from September 1999 until a bright out-
burst in mid 2002 (circa MJD 52520) kicked off a series of 5 outbursts (see Fig. 17(a)). From
the whole data range we extract an ephemeris of MJD 52306.1± 3.7 + n× 122.10± 0.38 d.
We note that this period is twice the optical period but we detect no significant flux half a
phase from maximum that would suggest the orbital period is not 122 days.
P = 4.7 × 10−9 s s−1 and
During the 1998 -- 99 outbursts a spin up was detected of
P = 5.9 × 10−9 s s−1. The
again throughout the 5 outbursts of 2002 -- 2003, with a value of
luminosities associated with these spin ups are, respectively, Lx ≥ 2.6× 1036 erg s−1 (B ≤
1.3× 1013 G) and Lx ≥ 3.3× 1036 erg s−1 (B ≤ 1.4× 1013 G). In between both groups of out-
bursts SXP59.0 was observed to have spun down during the ∼ 1100 days it was undetected.
The average spin down was P = −4.2 × 10−9 s s−1, which would be associated with an esti-
mated Lx ≥ 2.9× 1036 erg s−1. Although SXP59.0 was further away from the center of the
field of view during AO5 -- AO6 (MJD 51600 -- 52300), it should have been picked up as a
number of the detections during the second outburst season were made when RXTE was
-- 35 --
Fig. 16. -- SXP51.0, X-ray amplitude light curve.
-- 36 --
pointing at Position D (essentially the same coordinates as Position 5). In view of this, the
spin down mechanism for SXP59.0 must be something other than reverse accretion torque
and is likely due to the propeller effect.
2.21. SXP65.8
CXOU J010712.6-723533
RA 01 07 12.63, dec −72 35 33.8
History: Discovered as part of a Chandra survey of the SMC wing that is reported in
McGowan et al. (2007). They detected a source at Lx = 3× 1036 erg s−1 (37± 5% pulse frac-
tion) with pulsations at 65.78± 0.13 s. Its position was found to coincide with the emission
line star [MA93] 1619, a V = 16.6 Be star. A 110.6 d period has been observed in the
MACHO data for this object (Schmidtke & Cowley 2007b).
Survey Results: Due to its location outside the heavily observed positions, this pulsar has
very little data. We find no significant detections and no periodic modulation is apparent in
the X-ray light curve.
2.22. SXP74.7
RX J0049.1−7250, AX J0049−729
RA 00 49 04, dec −72 50 54
History: Discovered in the first observation of this survey with a period of 74.8± 0.4 s
(Corbet et al. 1998). Kahabka & Pietsch (1998) identified it with the ROSAT source RX
J0049.1−7250 and Stevens et al. (1999) found a single Be star within the ROSAT error radius
which they proposed as the optical counterpart. Only 3 X-ray outbursts where observed
in the early stages of this survey (before MJD 52300), from which Laycock et al. (2005)
derived a possible orbital period of 642± 59 d based on the separation between the outbursts.
Schmidtke & Cowley (2005b) and Edge (2005d) find a 33.4 d periodicity in OGLE data.
Survey Results: Lomb-Scargle analysis of the data finds a period at ∼ 62 d, which although
not strong, shows a convincing profile (see Fig. 18(b)); its ephemeris is MJD 52319.0± 3.1 +
n× 61.6± 0.2 d. There is a weak peak in the power spectrum at 33.2 d, close to the reported
optical period, but folding the light curve at this period does not produce a clear modulation.
-- 37 --
Fig. 17. -- SXP59.0. a) Top: X-ray amplitude light curve. b) Middle: Lomb-Scargle power
-- 38 --
Fig. 18. -- SXP74.7. a) Top: X-ray amplitude light curve. b) Middle: Lomb-Scargle power
-- 39 --
2.23. SXP82.4
XTE J0052−725
RA 00 52 09, dec −72 38 03
History: First observed by RXTE in this survey (Corbet et al. 2002), its position was
determined from archival Chandra observations (Edge et al. 2004b). OGLE data show a
strong modulation at ∼ 380 d (Edge 2006, private communication).
Survey Results: The Lomb-Scargle periodogram for the light curve shows a very significant
peak at ∼ 362 d with a number of harmonics (see Fig. 19(b)); the ephemeris derived is MJD
52089.0± 3.6 + n× 362.3± 4.1 d. Although this period is longer than would be expected
given its P s position in the Corbet diagram, its similarity to the optical period would seem
to confirm it is the actual orbital period of the system. We note that in the aforementioned
Chandra observation of this puslar on MJD 52459, it was detected at Lx = 3.4× 1036 erg s−1
(Edge et al. 2004c); this date is ∼ 8 days after our predicted periastron passage and the
luminosity exhibited is consistent with a Type I outburst (which was also detected with
RXTE see Fig. 19(a)).
2.24. SXP89.0
XTE SMC pulsar
No position available
History: Reported by Corbet et al. (2004b) from observations in March 2002, it is located
within the field of view of Position 1/A.
Survey Results: The first outburst is a single detection in February 2000 (MJD 51592), 2
years before the official discovery; 4 others occurred 2 years later in a short space of time --
within ∼ 260 days (see Fig. 20(a)). They follow a high-low-high-low brightness pattern, with
very similar high/low countrates. The separation between them is ∼ 88 d which could be
expected to be the orbital period. In fact, timing analysis finds a strong period at 87.6 d,
with an ephemeris of MJD 52337.5± 6.1 + n× 87.6± 0.3 d.
2.25. SXP91.1
AX J0051−722, RX J0051.3−7216
RA 00 50 55, dec −72 13 38
-- 40 --
Fig. 19. -- SXP82.4. a) Top: X-ray amplitude light curve. b) Middle: Lomb-Scargle power
-- 41 --
Fig. 20. -- SXP89.0. a) Top: X-ray amplitude light curve. b) Middle: Lomb-Scargle power
-- 42 --
History: Discovered in the first observation in this survey with a period of 92± 1.5 s
(Marshall et al. 1997), further analysis improved this measurement to 91.12± 0.05 s (Corbet et al.
1998). An orbital period of 115± 5 d was derived by Laycock et al. (2005) from early
survey data (before MJD 52200). Stevens et al. (1999) identified the optical counterpart;
Schmidtke et al. (2004) find an 88.25 d period in their analysis of MACHO data for this star.
Survey Results: Corbet et al. (2004b) reported the discovery of a new 89 s pulsar from
XTE observations in March 2002, it was located within the field of view of Position 1/A.
After studying the long-term lightcurves of these two pulsars, we believe SXP89.0 is actually
SXP91.1 after having spun up (see Fig. 21(a)). The Lomb-Scargle periodogram of the whole
light curve returns a period of ∼ 101 d, which is slightly shorter than the period found by
Laycock et al. (2005). Timing analysis of the light curve post MJD 52300 shows no clear
periods. The ephemeris we derive is MJD 52197.9± 8.2 + n× 117.8± 0.5 d. The average
spin up during MJD 50750 -- 51550 is calculated to be P = 1.8× 10−8 s s−1, with a luminosity
of Lx ≥ 3.6× 1036 erg s−1 (B ≤ 2.5× 1013 G).
Given the proximity in spin period between SXP91.1 and SXP89.0, and the fact that
the optical orbital period of SXP91.1 is so similar to the X-ray period of SXP89.0, it is
possible that they may be one and the same pulsar. In Fig. 22(a) we show the consolidated
light curve for both pulsars showing the possibility that SXP91.1 spun up sufficiently to be
later detected as a separate pulsar. We are hesitant to claim they are one single system
because the orbital ephemeris of this consolidated data set does not coincide with either
system's ephemeris: MJD 52301.2± 3.0 + n× 101.3± 0.4 d. However, the orbital profile
(see Fig. 22(b)) is convincing and appears similar to that of SXP8.88 (Fig. 9(b)) or SXP18.3
(Fig. 13(b)).
2.26. SXP95.2
SMC 95
RA 00 52, dec −72 45
History: Was discovered in March 1999 in data from this survey (Laycock et al. 2002);
an approximate position was obtained with PCA scans over the source and has a large
uncertainty. The X-ray orbital period suggested from the two bright outbursts before MJD
52300 is 283± 8 d (Laycock et al. 2005).
Survey Results: Only 3 other marginal detections are available in the data (see Fig. 23).
Lomb-Scargle analysis does not return any clear period, either including or excluding the
two bright outbursts, although the former has its highest power at 71.3 d, an ephemeris that
-- 43 --
Fig. 21. -- SXP91.1. a) Top: X-ray amplitude light curve. b) Middle: Lomb-Scargle power
-- 44 --
Fig. 22. -- Consolidated light curve of SXP89.0 and SXP91.1. a) Top: X-ray amplitude light
-- 45 --
is consistent with the source's significant detections and may reflect its orbital period.
2.27. SXP101
RX J0057.3−7325, AX J0057.4−7325
RA 00 57 26.8, dec −73 25 02
History: Discovered in an ASCA observation at a period of 101.45± 0.07 s (Yokogawa et al.
2000a), and identified also as a ROSAT source. Two optical counterparts were suggested
by Edge & Coe (2003); from a Chandra observation McGowan et al. (2007) pinpoint the
counterpart as a V = 14.9 star that exhibits a 21.9 d periodicity in both OGLE III and
MACHO data. Schmidtke & Cowley (2007b) find the same period.
Survey Results: This pulsar lies in the SE edge of the wing and was in the field of view
of Positions 4 and 5, so coverage is only continuous throughout AO5 and AO6 (see Fig. 24);
during this time only 3 outbursts of low brightness were observed (MJD 51814, 51863 and
52137). Timing analysis returns no significant periods, although the highest peak in the
periodogram is at 25.2 d, similar to the optical period.
2.28. SXP138
CXOU J005323.8−722715
RA 00 53 23.8, dec −72 27 15.0
History: Discovered in archival Chandra data (Edge et al. 2004b), the optical counterpart
is [MA93] 667 (Edge 2005d). The MACHO light curves for the companion star reveal peaks
at ∼ 125.1 d intervals (stronger in the red band). Schmidtke & Cowley (2006) report finding
a weak periodicity in the 122 -- 123 s region in MACHO data.
Survey Results: X-ray data from this survey show two brighter detections ∼ 112 d apart
(see Fig. 2.28(a)). Lomb-Scargle analysis finds no significant periods; however, if these two
detections are removed from the data, a period of 103.6 d appears in the power spectrum. Its
ephemerides (shown by the vertical, dashed lines in Fig. 2.28(a)) do not correspond with the
two detections, and the profile (see bottom panel of Fig. 2.28(b)) looks almost sinusoidal,
unlike the profiles exhibited by other systems in the SMC. The maxima for this period occur
at MJD 52400.2± 5.2 + n× 103.6± 0.5 d.
-- 46 --
Fig. 23. -- SXP95.2, X-ray amplitude light curve.
-- 47 --
Fig. 24. -- SXP101, X-ray amplitude light curve.
-- 48 --
Fig. 25. -- SXP138. a) Top: X-ray amplitude light curve. b) Middle: Lomb-Scargle power
-- 49 --
2.29. SXP140
XMMU J005605.2−722200, 2E0054.4−7237
RA 00 56 05.7, dec −72 22 00
History: Discovered in XMM observations by Sasaki et al. (2003). The optical counterpart
is believed to be [MA93] 904 (Haberl & Pietsch 2004). Schmidtke & Cowley (2006) find a
197± 5 d period in MACHO data.
Survey Results: None of the detections has been longer than 1 week, with only 2 of
them showing significant brightness (see Fig. 26). Timing analysis returns no clear periods,
although the periodogram of the data, excluding the two bright detections, does show some
power at ∼ 197 d. As a number of different periods have similar power, this may only be
coincidence. The folded light curve at 197 d does not show a typical orbital profile.
2.30. SXP144
XTE SMC pulsar
No position available
History: Detected in observations from this survey in April 2003 by Corbet et al. (2003b),
who later reported an ephemeris of MJD 52779.2± 2.9 + n× 61.2± 1.6 d (Corbet et al.
2003d).
Survey Results: Although there are a few minor detections before the initial discovery,
April 2003 (∼ MJD 52750) saw the beginning of a regular pattern of outbursts which contin-
ued until February 2006 (∼ MJD 53800, see Fig. 27(a)). The neutron star has displayed an ex-
tremely linear and constant spin down during this time, with an average P = 1.6×10−8 s s−1,
from which we derive a Lx ≥ 1.1× 1036 erg s−1 (B ≤ 2.4× 1013 G). The improved outburst
ephemeris from Lomb-Scargle analysis is MJD 52368.9± 1.8 + n× 59.38± 0.09 d. This pe-
riod is shorter than might have have expected from the pulse-orbit relationship, but what
is most unusual about this pulsar is that it is spinning down, moving it even further away
from the Be group on the Corbet diagram.
2.31. SXP152
CXOU J005750.3−720756
RA 00 57 49, dec −72 07 59
-- 50 --
Fig. 26. -- SXP140, X-ray amplitude light curve.
-- 51 --
Fig. 27. -- SXP144. a) Top: X-ray amplitude light curve. b) Middle: Lomb-Scargle power
-- 52 --
History: Haberl & Sasaki (2000) suggested that this object is a Be binary pulsar based on
the Hα-emitting object [MA93] 1038, although ROSAT observations of this source had not
detected any pulsations. These were found in a long Chandra observation by Macomb et al.
(2003) at a period of 152.098± 0.016 s (they report a very high pulse fraction of 64± 3% and a
Lx = 2.6× 1035 erg s−1), and in an XMM observation by Sasaki et al. (2003) at 152.34± 0.05 s.
Survey Results: The observational history is shown in Fig. 28. The periodogram of the
light curve shows no clear orbital period, although the highest power peak is at ∼ 107 d,
which would agree with the expected value from the Corbet diagram. The lack of periodic
outbursts, despite its clear X-ray activity, may point towards a low eccentricity system.
This would limit accretion onto the neutron star to times when the Be star ejecta are dense
enough, and would be independent of the orbital phase. Analysis of the optical light curve
of the companion Be star is needed.
2.32. SXP169
XTE J0054−720, AX J0052.9−7158, RX J0052.9−7158
RA 00 52 54.0, dec −71 58 08.0
History: First detected by RXTE in December 1998 at a period of 169.30 s (Lochner et al.
1998). Laycock et al. (2005) suggested a possible orbital period of 200± 40 d. Galache et al.
(2005) reported an orbital period of 68.6± 0.2 d based on data from this survey while
Schmidtke et al. (2006) found a period of 67.6± 0.3 d in OGLE -III data.
Survey Results: Corbet et al. (2004b) announced a new SMC pulsar at 164.7 s, with
an unknown position. After comparing the long term light curves and the ephemerides
from timing analysis it became apparent that SXP165 and SXP169 were the same source.
A consolidated light curve is presented here (Fig. 29(a)), where the spin up of SXP169
throughout the survey is apparent. The estimated spin up during MJD 50800 -- 51500 is P =
2.5 × 10−8 s s−1, implying Lx ≥ 1.2× 1036 erg s−1 (B ≤ 3.0× 1013 G); for the remaining data
the spin up is P = 2.0×10−8 s s−1, implying Lx ≥ 9.6× 1035 erg s−1 (B ≤ 2.6× 1013 G). Lomb-
Scargle analysis provides a clear period, and the outbursts are described by the ephemeris
MJD 52240.1± 2.1 + n× 68.54± 0.15 d, in agreement with the reported optical period.
2.33. SXP172
AX J0051.6−7311, RX J0051.9−7311
RA 00 51 52, dec −73 10 35
-- 53 --
Fig. 28. -- SXP152, X-ray amplitude light curve.
-- 54 --
Fig. 29. -- SXP169. a) Top: X-ray amplitude light curve. b) Middle: Lomb-Scargle power
-- 55 --
History: Found in an ASCA observation (Torii et al. 2000), it was identified with the
ROSAT source RX J0051.9−7311, which has a Be optical counterpart (Cowley et al. 1997).
Laycock et al. (2005) suggest a possible orbital period of ∼ 67 d based on the X-ray activity
up until MJD 52350. Schmidtke & Cowley (2006) report an optical period of 69.9± 0.6 d in
OGLE II data. This pulsar has been detected on 17 occasions by EINSTEIN, ROSAT and
ASCA, but never above Lx = 7.8× 1035 erg s−1 (Yokogawa et al. 2000c).
Survey Results: SXP172 underwent a phase of intense, semi-regular, activity during MJD
51600 -- 52400 (see Fig. 30). Lomb-Scargle analysis was carried using only data from this
period, and also the whole data set, with no conclusive outcome. However, it is clear that
the series of outbursts during the aforementioned dates are separated by ∼ 70 days (consistent
with the optical period). It is possible we are seeing contamination from another pulsar of
similar pulse period, maybe located within a different pointing position. As past missions
did not detect this pulsar in outburst, it is hard to confirm an X-ray ephemeris.
2.34. SXP202
XMMU J005920.8−722316
RA 00 59 20.8, dec −72 23 16
History: Detected in a number of archival XMM observations and reported in Majid et al.
(2004); the authors found an early B type star at the X-ray coordinates and classified it as
a HMXB.
Survey Results: This source has shown little bright activity throughout the survey except
for the outburst in January 2006 (circa MJD 53740, see Fig. 31). Lomb-Scargle analysis of
the data returns no clear period, but we note that a ∼ 91 d orbital period would agree with
the 6 outburst detections since MJD 53000.
2.35. SXP264
XMMU J004723.7−731226, RX J0047.3−7312, AX J0047.3−7312
RA 00 47 23.7, dec −73 12 25
History: It was initially reported by Yokogawa et al. (2003) from ASCA observations, al-
though it had previously been detected (yet remained undiscovered) in earlier XMM observa-
tions (Ueno et al. 2004). It had originally been proposed as a Be/X-ray binary candidate by
Haberl & Sasaki (2000) based on its X-ray variability. Edge et al. (2005a) identified the com-
-- 56 --
Fig. 30. -- SXP172, X-ray amplitude light curve.
-- 57 --
Fig. 31. -- SXP202, X-ray amplitude light curve.
-- 58 --
panion as the emission line star [MA93] 172 and found an optical period of 48.8± 0.6 d, which
they propose as the orbital period of the system. Further analysis of the OGLE light curve
finds an ephemeris of MJD 50592± 2 + n× 49.2± 0.2 d (Edge 2005d). Schmidtke & Cowley
(2005b) found a 49.1 d period in OGLE data.
Survey Results: Due to its location this pulsar was only observed consistently during AO5
and AO6 (see Fig. 32). No major outbursts were detected and the power spectrum shows no
significant peak at the optical period, nor at any other. We note that the optical ephemeris
does not agree with the X-ray activity.
2.36. SXP280
RX J0057.8−7202, AX J0058−72.0
RA 00 57 48.2, dec −72 02 40
History: Discovered in March 1998 in an ASCA observation by Yokogawa & Koyama
It is identified with the Be star [MA93] 1036, and
(1998a) at a period of 280.4± 0.3 s.
Schmidtke et al. (2006) find a 127.3 d period in its OGLE data with an epoch of maximum
brightness at MJD 52194.7. This pulsar was observed on 15 occasions by EINSTEIN, ROSAT
and ASCA, but never detected above Lx = 6.0× 1035 erg s−1 (Tsujimoto et al. 1999).
Survey Results: There are only 5 clear detections of this source throughout the survey,
and the power spectrum shows no significant periods, although the highest peak is at 64.8 d,
which is ∼ 1
2 of the optical period. For this reason, the ephemeris lines and folded light
curve are shown in Fig. 33. The tentative ephemeris of the X-ray period is MJD 52312± 6
+ n× 64.8± 0.2 d, which is within ∼ 12 d of the epoch of peak optical flux. A lack of earlier
detections of this system in outburst makes it difficult to firmly establish an X-ray ephemeris.
2.37. SXP293
XTE J0051−727
No position available
History: Reported by Corbet et al. (2004c) from observations during this survey of the
outburst at MJD 53097.
Survey Results: Only one of the 6 outbursts lasts longer than one week (Fig. 34(a)), which
could imply this is a Type II outburst, and thus not expected to coincide with periastron
passage. For this reason, it was removed from the data after initial timing analysis produced
-- 59 --
Fig. 32. -- SXP264, X-ray amplitude light curve.
-- 60 --
Fig. 33. -- SXP280. a) Top: X-ray amplitude light curve. b) Middle: Lomb-Scargle power
-- 61 --
no results. The resulting power spectrum, while not showing any significant periods, does
have its strongest peak at 151 d, an ephemeris that agrees well with the remaining 5 detec-
tions. We suggest this is the likely orbital period of the system, with an ephemeris of MJD
52327.3± 4.5 + n× 151± 1 d.
2.38. SXP304
RX J0101.0−7206, CXOU J010102.7−720658
RA 01 01 01.7, dec −72 07 02
History: Discovered in Chandra observations at 304.49± 0.13 s, the optical counterpart is
identified as the emission line star [MA93] 1240 (Macomb et al. 2003). The authors mea-
sured an unusually high pulse fraction of 90± 8% at a luminosity of Lx = 1.1× 1034 erg s−1.
Schmidtke & Cowley (2006) suggest there may be a 520± 12 d period in MACHO data of
the optical counterpart [MA93] 1240.
Survey Results: This source was out of the field of view during AO5 and AO6, so it has
only been adequately covered from AO7 onwards. A number of small outbursts were detected
during MJD 52600 -- 53000 (see Fig. 35), and it was not detected again until recently, in MJD
53747, when it began a 2 -- 3 week outburst, the longest and brightest observed so far in this
survey. Lomb-Scargle analysis of the data outside of this outburst finds moderate power at
∼ 50 d and no significant power at the optical period. It should be noted that this source
displays X-ray activity on timescales much shorter than the reported 520 d optical period.
2.39. SXP323
RX J0050.7−7316, AX J0051−733
RA 00 50 44.8, dec −73 16 06.0
History: Pulsations at 323.2± 0.4 s were detected in November 1997 (Yokogawa & Koyama
1998a) at the coordinates of the ROSAT source RX J0050.7−7316. Cowley et al. (1997)
identified the optical counterpart as a Be star. This system has been found to exhibit optical
and IR variability at periods of ∼ 0.7 and 1.4 d (Coe et al. 2002) and 1.695 d (Coe et al. 2005).
These periods are too short to be the orbital period of the system and are most likely non-
radial pulsations in the Be star. Laycock et al. (2005) suggest an orbital period of 109± 18 d
from X-ray data earlier in this survey.
Survey Results: This pulsar showed quite regular, bright activity during the survey (see
-- 62 --
Fig. 34. -- SXP293. a) Top: X-ray amplitude light curve. b) Middle: Lomb-Scargle power
-- 63 --
Fig. 35. -- SXP304, X-ray amplitude light curve.
-- 64 --
Fig. 36(a)). We found that the outburst circa MJD 52960 skewed the timing analysis results,
which we attribute to it being a Type II outburst; for this reason it was excluded from
the analysis. The ephemeris found for the remaining outbursts is MJD 52336.9± 3.5 +
n× 116.6± 0.6 d, which is consistent with the orbital period proposed by Laycock et al.
(2005).
2.40. SXP348
RX J0102−722, SAX J0103.2 −7209, AX J0103.2−7209
RA 01 03 13.0, dec −72 09 18.0
History: Pulsations at 345.2± 0.1 s were detected in BeppoSAX observations in July 1998
(Israel et al. 1998). Subsequently, pulsations at 348.9± 0.1 s were found in archival ASCA
data from May 1996 implying a P = −5.39 × 10−8 s s−1 (Yokogawa & Koyama 1998b). The
ROSAT source lies in a supernova remnant with a Be counterpart; a weak 93.9 d periodicity
has been reported from OGLE data (Schmidtke & Cowley 2006). Israel et al. (2000) suggest
this is a persistent, low luminosity X-ray system.
Survey Results: Although this source has been detected by a number of different instru-
ments, it was always at low luminosities (.1036 erg s−1), so it is not surprising that there are
not many detections in the survey Fig. 37. Furthermore, it has also been detected at a wider
range of periods than other pulsars (from 340 to 348 s), which makes it more difficult to pick
out in the periodograms from weekly observations. Timing analysis reveals no significant
periods.
2.41. SXP452
RX J0101.3−7211
RA 01 01 19.5, dec −72 11 22
History: Was initially proposed as an X-ray binary by Haberl et al. (2000). Pulsations
were detected in XMM observations during 2001 at 455± 2 s and in 1993 ROSAT data at
450 -- 452 s (Sasaki et al. 2001), implying a P = 2.3 × 10−8 s s−1; the optical counterpart was
identified as a Be star Sasaki et al. (2001). Schmidtke et al. (2004) propose an orbital period
of 74.7 d for this system based on its optical variability.
Survey Results: With only 5 detections throughout the survey (Fig. 38), this source's
periodogram shows no periodicities and there is no evidence for the reported 74.7 d optical
-- 65 --
Fig. 36. -- SXP323. a) Top: X-ray amplitude light curve. b) Middle: Lomb-Scargle power
-- 66 --
Fig. 37. -- SXP348, X-ray amplitude light curve.
period.
-- 67 --
2.42. SXP504
RX J0054.9−7245, AX J0054.87244, CXOU J005455.6−724510
RA 00 54 55.6, dec −72 45 10
History: Discovered in archival Chandra data with 503.5± 6.7 s pulsations (Edge et al.
2004d,c). Also reported by Haberl et al. (2004) from an XMM observation at 499.2± 0.7 s
and Lx = 4.3× 1035 erg s−1. An orbital period of 268.0± 1.4 d was determined from op-
tical (OGLE) data, and corroborated by preliminary X-ray data from the present sur-
vey (Edge et al. 2005b). The optical ephemeris was later refined to MJD 50556± 3 +
n× 268.0± 0.6 d (Edge et al. 2005c). Schmidtke & Cowley (2005b) found a period of 273 d
in OGLE data.
Survey Results: The light curve for this system can be seen in Fig. 39(a). Lomb-Scargle
analysis of the entire survey (with or without the bright outbursts circa MJD 52440 and
52980) returns a slightly shorter orbital period to the optical one previously reported, but
is consistent within errors. Furthermore, Edge et al. (2005c) find that the epochs of maxi-
mum X-ray flux are coincident with the optical outbursts. The ephemeris we find is MJD
52167.4± 8.0 + n× 265.3± 2.9 d. This source displays a lot of activity in between periastron
passages, which might be indicative of a low eccentricity orbit and makes timing analysis
more difficult.
2.43. SXP565
CXOU J005736.2−721934
RA 00 57 36.2, dec −72 19 34
History: Discovered in Chandra observations at 564.81± 0.41 s with a pulse fraction of
48± 5% (Lx = 5.6× 1034 erg s−1), the optical counterpart is idenified as the emission line
star [MA93] 1020 (Macomb et al. 2003). An optical period of 95.3 d has been reported for
this system (Schmidtke et al. 2004), but this period is not seen in OGLE data (Edge 2005d).
Survey Results: This source shows a lot of variability, but no clear outbursts (see Fig. 40(a)).
The power spectrum returns two significant peaks, with only the higher one providing a con-
vincing orbital profile. The ephemeris for this period is MJD 52219.0± 13.7 + n× 151.8± 1.0 d.
We note there is no power at the reported optical period nor does the optical ephemeris agree
-- 68 --
Fig. 38. -- SXP452, X-ray amplitude light curve.
-- 69 --
Fig. 39. -- SXP504. a) Top: X-ray amplitude light curve. b) Middle: Lomb-Scargle power
-- 70 --
with the brightest X-ray detections of this source. In view of this, it is possible that the
optical counterpart has been misidentified; however, there is no other bright candidate star
within 10 arcsec of the very precise Chandra position. The folded light curve has a unique
shape, showing significant activity around periastron. The asymmetry of the periastron
peak is likely due to the presence of an accretion disk around the neutron star that is not
completely consumed by the time it reaches apastron, where its lower orbital velocity allows
it to "top up" its disk from the Be star's wind.
2.44. SXP700
CXOU J010206.6-714115
RA 01 02 06.69, dec −71 41 15.8
History: Discovered as part of the aforementioned Chandra survey of the SMC wing re-
ported in McGowan et al. (2007). They detected a source with Lx = 6.0× 1035 erg s−1
(35± 5% pulse fraction) with pulsations at 700.54± 34.53 s and found its position to co-
incide with the emission line star [MA93] 1301, a V = 14.6 O9 star. The OGLE data
available for this object show a periodic modulation at 267.38± 15.10 d (McGowan et al.
2007).
Survey Results: Due to its proximity in period to SXP701, limitations of the current anal-
ysis script (which cannot work on two pulsars of such similar spin periods) have not allowed
the extraction of an X-ray light curve for this pulsar. However, it has not contaminated
the data presented for SXP701 because, even though they are ∼ 66 -- arcmin away from each
other, they were never within the same field of view.
2.45. SXP701
RX J0055.2−7238, XMMU J005517.9−723853
RA 00 55 17.9, dec −72 38 53
History: First observed during an XMM TOO observation at 701.6± 1.4 s and located
within the error circle of a ROSAT object (Haberl et al. 2004). Fabrycky (2005) finds optical
periods of 6.832 and 15.587 h, which are attributed to stellar pulsations. A weak 412 d period
has been seen in MACHO data (Schmidtke & Cowley 2005b).
Survey Results: Similar to SXP565, it displays much X-ray variability with no bright
outbursts (see Fig. 41). Unfortunately, timing analysis provides no clear periodicities.
-- 71 --
Fig. 40. -- SXP565. a) Top: X-ray amplitude light curve. b) Middle: Lomb-Scargle power
-- 72 --
Fig. 41. -- SXP701, X-ray amplitude light curve.
-- 73 --
2.46. SXP756
RX J0049.6−7323, AX J0049.4−7323
RA 00 49 42.42, dec −73 23 15.9
History: Pulsations at 755.5± 0.6 s were detected in a very long (∼ 177 ks) ASCA observa-
tion of the SMC in April 2000, the source was detected at a luminosity of Lx = 5× 1035 erg s−1
(Yokogawa et al. 2000b). The optical counterpart was later established as a V = 15 Be
star (Edge & Coe 2003). Laycock (2002) and Laycock et al. (2005) find an X-ray period of
396± 5 d while Cowley & Schmidtke (2003); Schmidtke et al. (2004) report recurring out-
bursts at ∼ 394 d intervals in R and V MACHO data.
Survey Results: Coverage of this pulsar has been sparse due to its southern location in
the SMC, only visible to observations of Positions 4 and 5. Despite this, it has been detected
various times in outburst, the brightest of which was a 3 week outburst (Fig. 42(a)). We
decided to remove the brightest point from this outburst so as not to skew the Lomb-Scargle
power spectrum. A strong period is found at 194.9 d which we believe to be the first harmonic
of the orbital period. Looking at the light curve it is evident that the bright detections are
spaced ∼ 400 d apart3. Moreover, this would be in agreement with the optical period of
the system, so we use twice the value of the harmonic as the orbital period and derive the
ephemeris MJD 52196.1± 3.9 + n× 389.9± 7.0 d. This ephemeris places the first outburst
in the data at MJD 51416, which is consistent with the last optical outburst available in the
MACHO data (Cowley & Schmidtke 2003).
2.47. SXP1323
RX J0103.6−7201
RA 01 03 37.5, dec −72 01 33.2
History: This is the source with the longest known pulse period in the SMC, SXP1323
was reported by Haberl & Pietsch (2005) in a number of archival XMM observations. The
authors identify the emission line star [MA93] 1393 (V ≃ 14.6) as the optical counterpart.
Schmidtke & Cowley (2006) report three strong periods for this source from OGLE II data:
0.41 d, 0.88 d and 26.16 d; they attribute the first two to non-radial pulsations of the Be star
but suggest the latter might be the orbital period.
3Also note that only one peak appears in the orbital profile in Fig. 42(b).
If 194.9 d were the actual
orbital period, the plot would show two peaks per orbital cycle.
-- 74 --
Fig. 42. -- SXP756. a) Top: X-ray amplitude light curve. b) Middle: Lomb-Scargle power
-- 75 --
Survey Results: This source is difficult to detect due to its long period (requiring ob-
servations with a baseline longer than ∼ 4 ks) and its location near the edge of Position
1/A. Despite these limitations, a number of bright outbursts have been detected but no
periodicities can be found in the sparse data (see Fig. 43).
3. Discussion
Table 3 presents a summary of the timing analysis results for the SMC pulsars, providing
orbital periods and ephemerides where available, together with periods observed at other
wavelengths. In cases when there is only weak evidence for an orbital period in the X-ray
data, the number is given in parentheses.
Those systems with unequivocal orbital periods have allowed us to study and compare
a selection of well-defined orbital profiles. In particular, we consider the profiles of the fol-
lowing systems that have shown regular activity at, or around, periastron pasage: SXP6.85,
SXP7.78, SXP46.6, SXP59.0, SXP82.4, SXP144, SXP169 and SXP756. It is notable that
their profiles all share a similar shape, even though their orbital periods cover a wide range
(from ∼ 45 -- 390 d): they are relatively narrow (except for SXP756, all widths are between
0.40 and 0.55 wide in orbital phase), and they are very symmetrical in shape (suggesting ac-
cretion takes place fluidly and increases smoothly as the neutron star approaches the deccre-
tion disk, to decrease in an equally smooth manner after periastron). Other systems display
different characteristics, some show sharp rises with long declines (e.g.,SXP323, Fig. 36(b)),
the opposite (e.g., SXP16.6, Fig. 12(b)), or even long, plateau-like profiles (e.g.,SXP280,
Fig. 33(b)).
A number of profiles show a small increase in flux half an orbital phase away from
the maximum (e.g., SXP7.78, SXP89.0, SXP144, SXP504 and SXP565), suggesting that
accretion is taking place on a smaller scale, possibly from the Be star's wind, when the
neutron star is travelling slowly enough to accrete from it. Given the range of orbit sizes
in the systems that exhibit this behaviour (judging from the orbital periods), it may be the
eccentricity of the orbit that determines whether apastron accretion takes place (because the
more eccentric the orbit, the slower the neutron star will be travelling at apastron).
Throughout the length of the survey there have been some pulsars whose spin periods
have undergone slow, persistent evolution, e.g., SXP7.78 (spin down), SXP46.6 (spin up),
SXP59.0 (spin up) or SXP144 (spin down), while others have changed their pulse period
over the course of a few weeks, e.g., SXP15.3 (spin up). And still, some other systems have
not changed their spin periods significantly despite bright and/or persistent activity, e.g.,
-- 76 --
Fig. 43. -- SXP1323, X-ray amplitude light curve.
-- 77 --
SXP2.37, SXP4.78 or SXP8.88. This suggests there are a number of pulsars that must be
rotating close to, or at, their equilibrium spin periods, as suggested by Corbet (1984). A
lack of modulation in the spin period throughout a long outburst (as in SXP18.3's outburst
circa MJD 54000) further reveals systems that are likely being observed at a low inclination
(face on). Unfortunately, despite the number of orbital periods known, it has still not been
possible to calculate the complete set of orbital parameters of any SMC system (with the
exception of SXP0.92 (Kaspi et al. 1993)), so orbital inclination angles and their effects on
pulse arrival times remain a matter of speculation.
An interesting case is presented by SXP348; Israel et al. (2000) suggest this is a persis-
tent, low luminosity system, and recent Chandra observations (McGowan et al. 2007) tend
to support this notion after detecting the pulsar at its shortest recorded pulse period until
then, 337.5 s. Given that it was observed in 1996 at 349 s in ASCA data, we can deduce
that it has spun down 11.5 s in ∼ 10 years, implying a P = 3.6 × 10−8 s s−1. An estimate
for the accretion necessary to produce the required torque on this timescale implies a time-
averaged X-ray luminosity of just Lx ≥ 3.3× 1035 erg s−1, which would explain why it was
rarely detected during the present survey. However, RXTE detected it a number of times
at longer pulse periods than would have been expected; a clear occasion was during the
short outburst circa MJD 53740, when the system was detected at a high significance on 2
consecutive weeks with periods of 345.0 s and 336.2 s, respectively. If this spin-up were all
intrinsic, the expected luminosity would be Lx ≥ 1.1× 1038 erg s−1, which is not inconceiv-
able for a Type II outburst. McGowan et al. (2007) detected it 30 d later (MJD 53777) at
Lx = 4.2× 1037 erg s−1, but RXTE did not detect it 2 days prior to that (MJD 53774). So
what is happening in this system? It is not clear whether we are detecting orbitally-induced
Doppler-shifted pulse periods or if SXP348 is actually undergoing massive spin up every time
it outbursts. Given the pulse periods of the detections in Fig. 37, it is unclear whether the
pulsar is, in fact, spinning down (as proposed by Israel et al. (2000)); it could very well be
spinning up.
Another case involving changes in spin is that of SXP59.0 (Fig. 17(a)). It went unde-
tected by RXTE during MJD 51550 -- 52500, during which time its pulse period increased
∼ 0.45 s. If this spin-down were a product of reverse accretion torques, we would have ex-
pected to see outbursts at Lx ≥ 2.6× 1036 erg s−1, comparable in strength with those detected
during the survey. Although SXP59.0 was not close to the center of the field of view during
the spin-down period (as it was for most of the rest of the survey), it should still have been
detectable. This suggests that SXP59.0 was spun down through mechanisms other than
reverse accretion torques, possibly through the propeller effect.
Yet another example of strange spin behaviour is that of SXP144. Given its pulse period,
-- 78 --
and based on the Corbet diagram, its orbital period should be ∼ 90 d; so with its clear 59.4 d
orbital period, the expectation is that it would spin up during outburst. However, quite the
opposite is what RXTE has detected. This system went into outburst at least once every
orbit over a period of ∼ 1160 d, during which time it spun down ∼ 1.7 s. It would seem that
factors other than just the orbital and spin periods govern the behaviour and evolution of
these binary systems, and they need to be taken into account. Orbital eccentricity and/or
the magnetic field strength of the pulsar are likely to be important factors in determining
the relationship between the spin and orbital periods.
4. Conclusion
We have presented the most comprehensive colection of pulsed light curves for 41 prob-
able Be/X-ray systems in the SMC and have determined an X-ray (likely orbital) ephemeris
for 21 of them while presenting possible periods for 6 others. 10 of these ephemerides are
from new X-ray periods, while 6 others are improvements of previously known ephemerides.
Of the systems exhibiting periodicities in the X-ray, 19 also show optical periodicities, al-
though it is noteworthy that the optical and X-ray periods only agree on 7 occasions. Of
the remaining 12, 3 show an optical period that is half the X-ray period while 1 exhibits
an optical modulation that is twice the X-ray period. However, there are still 11 systems
with a known optical period that show no modulation in X-rays. The systems exhibiting
a large discrepancy between the optical and X-ray periods suggest the possibility that the
optical counterpart to the X-ray system has been missidentified, or that the interaction be-
tween the Be star's disk and the neutron star is more complicated than originally thought,
although optical periods can be affected by pulsations of the primary star, and X-ray periods
by Type II outbursts. Nonetheless, for extragalactic X-ray binaries, long-term monitoring
has proven to be the most successful method of study.
-- 79 --
Table 3. X-ray binary systems in the SMC
Object
P s (s)
RA
dec
X-ray T P (MJD) X-ray P orb (days)
Other P orb (days)
SXP0.09
SXP0.72
SXP0.92
SXP2.16
SXP2.37
SXP2.76
SXP3.34
SXP4.78
SXP6.85
SXP7.78
SXP8.02
SXP8.88
SXP9.13
SXP15.3
SXP16.6
SXP18.3
SXP22.1
SXP31.0
SXP34.1
SXP46.6
SXP51.0
SXP59.0
SXP65.8
SXP74.7
SXP82.4
SXP89.0
SXP91.1
SXP95.2
SXP101
SXP138
SXP140
SXP144
SXP152
SXP169
SXP172
SXP202
SXP264
SXP280
SXP293
0.087
0.716
0.92
2.165
2.374
2.763
3.34
4.782
6.848
7.780
8.020
8.880
9.130
15.30
16.52
18.37
22.07
31.01
34.08
46.59
51.00
58.86
65.78
74.70
82.40
89.00
91.10
95.20
101.40
138.00
140.99
144.00
152.10
167.35
172.40
202.00
263.60
280.40
293.00
00 42 35.0
-73 40 30.0
01 17 05.2
-73 26 36.0
00 45 35.0
-73 19 02.0
01 19 00
-73 12 27
00 54 34.0
-73 41 03.0
00 59 11.7
-71 38 48.0
01 05 02.0
-72 11 00.0
00 52 06.6
-72 20 44
--
--
--
--
--
--
--
--
01 02 53.1
-72 44 33
52318.5 ± 7.9
00 52 06
-72 26 06
52250.9 ± 1.4
01 00 41.8
-72 11 36
--
--
--
--
--
--
--
--
(34.1)
112.5 ± 0.5
44.92 ± 0.06
--
00 51 52.0
-72 31 51.7
52392.2 ± 0.9
28.47 ± 0.04
--
--
51a
--
--
82.1b
--
23.9c
24.82d
44.86e , 44.6f
, 44.8g
--
33.4h
00 49 13.6
-73 11 39
52380.5 ± 2.3
77.2 ± 0.3
40.17g , 91.5i
00 52 14
-73 19 19
--
--
00 50
--
-72 42
52373.5 ± 1.0
52275.6 ± 0.9
(28)
33.72 ± 0.05
17.73 ± 0.01
01 17 40.5
-73 30 52.0
01 11 09.0
-73 16 46.0
00 55 26.9
-72 10 59.9
--
--
--
--
--
--
00 53 58.5
-72 26 35.0
52293.9 ± 1.4
137.36 ± 0.35
--
--
--
--
00 54 56.6
-72 26 50
52306.1 ± 3.7
122.10 ± 0.38
01 07 12.63
-72 35 33.8
--
00 49 04
00 52 09
--
00 50 55
00 52 00
-72 50 54
-72 38 03
--
-72 13 38
-72 45 00
00 57 26.8
-73 25 02
52319.0 ± 3.1
52089.0 ± 3.6
52337.5 ± 6.1
52197.9 ± 8.2
--
--
--
61.6 ± 0.2
362.3 ± 4.1
87.6 ± 0.3
117.8 ± 0.5
(71.3)
(25.2)
75.1j
--
34.6k
--
90.4b
69.2l
--
, 138.4l
--
60.2m
110.6n
33.4o
∼ 380p
--
88.25q
--
21.9r
00 53 23.8
-72 27 15.0
52400.5 ± 5.2
103.6 ± 0.5
∼ 125s , 122 -- 123t
00 56 05.7
-72 22 00.0
--
--
--
--
52368.9 ± 1.8
59.38 ± 0.09
00 57 49
-72 07 59
--
--
00 52 54.0
-71 58 08.0
52240.1 ± 2.1
68.54 ± 0.15
00 51 52
-73 10 35
00 59 20.8
-72 23 17
00 47 23.7
-73 12 25
--
--
--
00 57 48.2
-72 02 40
52312 ± 6
--
--
52327.3 ± 4.5
(70)
(91)
--
64.8 ± 0.2
151 ± 1
197t
--
--
67.6u
69.9v
--
49.2g , 49.1w
127.3u
--
-- 80 --
Table 3 -- Continued
Object
P s (s)
RA
dec
X-ray T P (MJD) X-ray P orb (days) Other P orb (days)
520t
--
93.9v
74.7x
268.0y , 273w
95.3q
267z
412aa
394ab
26.16v
304.49
321.20
344.75
452.01
503.50
564.83
01 01 01.7
-72 07 02
--
--
00 50 44.8
-73 16 06.0
52336.9 ± 3.5
116.6 ± 0.6
01 03 13.0
-72 09 18.0
01 01 19.5
-72 11 22
--
--
--
--
00 54 55.6
-72 45 10.0
52167.4 ± 8.0
00 57 36.2
-72 19 34.0
52219.0 ± 13.7
265.3 ± 2.9
151.8 ± 1.0
700.5
01 02 06.69
-71 41 15.8
00 55 17.9
-72 38 53.0
--
--
--
--
SXP304
SXP323
SXP348
SXP452
SXP504
SXP565
SXP700
SXP701
SXP756
696.37
755.50
00 49 42.42
-73 23 15.9
52196.1 ± 3.9
389.9 ± 7.0
SXP1323
1323.20
01 03 37.5
-72 01 33.2
--
--
aFrom radio observations (Kaspi et al. 1993).
bFrom OGLE III data (Schmidtke et al. 2006).
cFrom MACHO data of the possible optical counterpart (Coe et al. 2005).
dFrom OGLE and MACHO data (Schmidtke & Cowley 2007).
eFrom MACHO data (Cowley & Schmidtke 2004).
f From MACHO data (Coe et al. 2005).
gFrom OGLE data (Edge 2005d).
hFrom OGLE and MACHO data (Schmidtke & Cowley 2006).
iFrom OGLE data (Schmidtke et al. 2004).
jFrom OGLE and MACHO data (Edge 2005d).
kFrom RXTE data (Corbet et al. 2004d).
lFrom OGLE data (Schmidtke et al. 2007).
mFrom OGLE and MACHO data (Schmidtke & Cowley 2005b).
nFrom MACHO data (Schmidtke & Cowley 2007b).
oFrom OGLE data (Schmidtke & Cowley 2005b; Edge 2005d).
pFrom OGLE data (Edge 2006, private communication).
qFrom MACHO data (Schmidtke et al. 2004).
rFrom OGLE and MACHO data (McGowan et al. 2007; Schmidtke & Cowley 2007b).
sFrom MACHO data (Edge 2005d).
tFrom MACHO data (Schmidtke & Cowley 2006).
uFrom OGLE data (Schmidtke et al. 2006).
vFrom OGLE data (Schmidtke & Cowley 2006).
wFrom OGLE data (Schmidtke & Cowley 2005b).
xFrom OGLE and MACHO data (Schmidtke et al. 2004).
-- 81 --
yFrom OGLE and MACHO data (Edge et al. 2005c).
zFrom OGLE data (McGowan et al. 2007).
aaFrom MACHO data (Schmidtke & Cowley 2005b).
abFrom MACHO data (Cowley & Schmidtke 2003; Schmidtke et al. 2004).
-- 82 --
We thank A. Cowley for pointing out some glaring omissions in the final manuscript
and also the anonymous referee for many useful comments that made this paper clearer.
Facilities: RXTE (PCA).
A. Determining signal significance
Scargle (1982) proposes the normalisation of the periodogram to the variance, σ2
n, of
the signal-free data, where σn is the standard deviation; as such, Gaussian noise will have
a power of 14. Furthermore, the probability function Prob associated with the periodogram
will be exponentially distributed5, and it can be shown that the probability that a periodic
signal with power of Z is due to noise is
FAP = 1 − (1 − e−Z)M
(A1)
which is the False Alarm Probability, with M being the number of independent frequencies,
which we (rather conservatively) define as
M = 2 × nf × ∆f × τ
(A2)
where nf is the number of scanned frequencies, ∆f is the frequency interval used when
calculating the periodogram, and τ is the duration of the observation.
A more useful number may be the significance of a detection, or how sure we are that
it is real; this is simply 1 − FAP, expressed as a percentage. This is the value that will be
used throughout this paper to estimate the importance and believability of a signal, and is
given by
Sig(%) = 100 × (1 − e−Z)M
(A3)
4It is clear that most of the data from observations made by RXTE during the survey contain contributions
from a number of sources in the field of view, and their variance should not be used in the calculations.
However, after analysing a large number of observations, it was found that the average power within the
calculated power spectra was essentially 1 (likely due to the low S/N), which justifies our use of the light
curve variance (including all the pulsar signals) instead of the variance of the noise, which would have been
difficult to obtain.
5It will be of the form Prob = e−Z, which is the probability of detecting a peak in the periodogram above
a certain power, Z.
-- 83 --
In the Lomb-Scargle periodogram, the peak-to-trough amplitude A of the modulation
in the signal is related to the power PLS through
A = 4rPLS σ2
N
n
(A4)
where N is the number of data points (Scargle 1982).
If the signal detected has any harmonics, its total power will be divided between the
individual harmonic peaks in the power spectrum. Using only the fundamental to estimate
the amplitude of the signal's pulsations could then severely underestimate it if there were
considerable power in any of the harmonics (which is often the case). If the amplitude of all
the harmonics is known, it can be shown that the total amplitude of the signal will be given
by
Atotal =sXi
A2
i
(A5)
From the error in the angular frequency detected at a certain power in the Lomb-Scargle
periodogram (Horne & Baliunas 1986), we derive the error on the period as
dP =
3
4(cid:18) P 2 σn√N τ A(cid:19)
(A6)
where A is the Lomb-Scargle amplitude given by Eq. (A5), and σn is the standard deviation
of the noise, although the standard deviation of the actual data is used, as explained earlier.
Apart from the global significance of a detection, we define an additional quantity, the
local significance, as the significance of a peak at frequency f , within a region of frequency
space extending 5% of f to either side of it.
B. Luminosity and magnetic field estimation
If all the matter accreting onto a neutron star is converted into energy, the luminosity
that will result is simply the gravitational energy lost by the in-falling mass (Frank et al.
2002):
-- 84 --
Lx =
GM M
R
(B1)
where M is the mass transfer rate. Some manipulation and substitution can provide a more
manageable expression:
Lx37 = 8.4 × 109 Mn M
Rn !
(B2)
which will give us the luminosity in terms of 1037 erg s−1, and where Mn is the mass of the
M is the mass accretion rate in units of M ⊙ yr−1, and Rn is
neutron star in units of M ⊙,
the radius of the neutron star in km.
The angular momentum of a neutron star is given by
where P s is the spin period, and the moment of inertia is given by
Ln =
2πIn
Ps
In =
2
5
MnR2
n
with Mn and Rn being the mass and radius of the neutron star in standard units.
The torque experienced by a neutron star spinning up or down is given by
(B3)
(B4)
(B5)
= 2πIn
Ps
P 2
s
τ ≡(cid:12)(cid:12)(cid:12)(cid:12)
dLn
dt (cid:12)(cid:12)(cid:12)(cid:12)
with P s being the rate of change of the spin period (Frank et al. 2002).
For an accreting pulsar undergoing steady spin up/down, the applied torque will depend
M , and the angular momentum of matter in the accretion disc
on the mass accretion rate,
at the magnetospheric radius, rm. This torque is given by
τ = MpGMnrm
(B6)
-- 85 --
The maximum torque possible will occur when rm = rco, where the corotation radius6
is given by
1
3
rco =(cid:18)GMnP 2
4π2 (cid:19)
s
(B7)
Substituting this value in Eq. (B6) will provide the expression for the maximum torque
possible:
τmax = M(cid:20) G2M 2
2π
nPs
1
3
(cid:21)
(B8)
Clearly, τ ≤ τmax, so using Eqs. B5 and B8, and substituting the expression for the moment
of inertia from Eq. (B4), we find the accretion rate will be
M ≥
2
5
R2
n
1
3
Ps(cid:20) 16π4Mn
s G2 (cid:21)
P 7
kg s−1
(B9)
substituting this value in Eq. (B1) we finally obtain the lower limit on the luminosity that
will be produced through accretion:
Lx37 ≥
2Rn Ps
5 × 1030 (cid:20)16π4GM 4
P 7
s
n
1
3
(cid:21)
(B10)
which will be in units of 1037 erg s−1 if S.I. units are used (and the neutron star mass is in
M ⊙). This equation will allow the estimation of the luminosity associated with an outburst
if the average spin up/down is measured.
One further value that can be estimated is the magnetic field of the neutron star.
Rearranging Eq. (6.24) of Frank et al. (2002), and using the period in place of the frequency,
a constraint can be placed on its value:
B12 ≤(cid:20)3.4 × 10−4 R−2
n M
−10
7
n L
6
7
x37
−7
2
P 2
s
Ps (cid:21)
(B11)
6The corotation radius is defined as the radius at which matter in the disc is moving at the same speed
as the neutron star's surface.
-- 86 --
where the magnetic field will be in units of 1012 G if Mn is in M ⊙ and Rn in metres. In the
present work we assume values for the neutron star's radius and mass of Rn = 104 m and
Mn = 1.4 M⊙, respectively.
REFERENCES
Ables J. G., Jacka C. E., Hall P. J., Hamilton P. A., McConnell D., McCulloch P. M., 1987,
IAU Circ., 4422, 1
Bell J. F., 1994, Proceedings of the Astronomical Society of Australia, 11, 81
Blackburn J. K., 1995, in Shaw R. A., Payne H. E., Hayes J. J. E., eds, Astronomical Data
Analysis Software and Systems IV Vol. 77 of Astronomical Society of the Pacific
Conference Series, FTOOLS: A FITS Data Processing and Analysis Software Package.
pp 367 -- +
Buckley D. A. H., Coe M. J., Stevens J. B., van der Heyden K., Angelini L., White N.,
Giommi P., 2001, Mon. Not. R. Astron. Soc., 320, 281
Carstairs I. R., 1992, Ph.D. Thesis, University of Southampton
Chakrabarty D., Levine A., Clark G., Takeshima T., Wilson C., Finger M., 1998, IAU Circ.,
7048, 1
Charles P. A., Southwell K. A., O'Donoghue D., 1996, IAU Circ., 6305, 2
Clark G., Doxsey R., Li F., Jernigan J. G., van Paradijs J., 1978, Astrophys. J., Lett., 221,
L37
Clark G., Remillard R., Woo J., 1996, IAU Circ., 6282, 1
Clark G. W., Remillard R. A., Woo J. W., 1997, Astrophys. J., Lett., 474, L111+
Coe M. J., Edge W. R. T., Galache J. L., McBride V. A., 2005, Mon. Not. R. Astron. Soc.,
356, 502
Coe M. J., Haigh N. J., Laycock S. G. T., Negueruela I., Kaiser C. R., 2002, Mon. Not. R.
Astron. Soc., 332, 473
Coe M. J., Stevens J. B., Buckley D. A. H., Charles P. A., Southwell K. A., 1998, Mon. Not.
R. Astron. Soc., 293, 43
-- 87 --
Corbet R. H. D., 1984, Astron. Astrophys., 141, 91
Corbet R., Marshall F. E., Lochner J. C., Ozaki M., Ueda Y., 1998, IAU Circ., 6803, 1
Corbet R. H. D., Marshall F. E., Coe M. J., Laycock S., Handler G., 2001, Astrophys. J.,
Lett., 548, L41
Corbet R., Markwardt C. B., Marshall F. E., Laycock S., Coe M., 2002, IAU Circ., 7932, 2
Corbet R., Markwardt C. B., Marshall F. E., Coe M. J., Edge W. R. T., Laycock S., 2003a,
IAU Circ., 8064, 4
Corbet R. H. D., Markwardt C. B., Marshall F. E., Coe M. J., Edge W. R. T., Laycock S.,
2003b, The Astronomer's Telegram, 163, 1
Corbet R. H. D., Edge W. R. T., Laycock S., Coe M. J., Markwardt C. B., Marshall F. E.,
2003c, AAS/High Energy Astrophysics Division, 7, 1
Corbet R. H. D., Laycock S., Marshall F. E., Markwardt C. B., Coe M. J., 2003d, The
Astronomer's Telegram, 209, 1
Corbet R. H. D., Markwardt C. B., Coe M. J., Edge W. R. T., Laycock S., Marshall F. E.,
2003e, The Astronomer's Telegram, 214, 1
Corbet R. H. D., Coe M. J., Edge W. R. T., Laycock S., Markwardt C. B., Marshall F. E.,
2004a, The Astronomer's Telegram, 277, 1
Corbet R. H. D., Laycock S., Coe M. J., Marshall F. E., Markwardt C. B., 2004b, in AIP
Conf. Proc. 714: X-ray Timing 2003: Rossi and Beyond Monitoring and Discovering
X-ray Pulsars in the Small Magellanic Cloud. pp 337 -- 341
Corbet R. H. D., Markwardt C. B., Coe M. J., Edge W. R. T., Laycock S., Marshall F. E.,
2004c, The Astronomer's Telegram, 273, 1
Corbet R. H. D., Markwardt C. B., Marshall F. E., Coe M. J., Edge W. R. T., Galache J. L.,
Laycock S., 2004d, The Astronomer's Telegram, 347, 1
Cowley A. P., Schmidtke P. C., 2003, Astrophys. J., 126, 2949
Cowley A. P., Schmidtke P. C., 2004, Astrophys. J., 128, 709
Cowley A. P., Schmidtke P. C., McGrath T. K., Ponder A. L., Fertig M. R., Hutchings J. B.,
Crampton D., 1997, Publ. Astron. Soc. Pac., 109, 21
-- 88 --
de Jager O. C., Raubenheimer B. C., North A. R., Nel H. I., van Urk G., 1988, Astrophys.
J., 329, 831
Edge W. R. T., Coe M. J., 2003, Mon. Not. R. Astron. Soc., 338, 428
Edge W. R. T., Coe M. J., Corbet R. H. D., Markwardt C. B., Laycock S., 2004a, The
Astronomer's Telegram, 225, 1
Edge W. R. T., Coe M. J., Corbet R. H. D., Markwardt C. B., Laycock S., Marshall F. E.,
2004b, The Astronomer's Telegram, 216, 1
Edge W. R. T., Coe M. J., Galache J. L., McBride V. A., Corbet R. H. D., Markwardt C. B.,
Laycock S., 2004c, Mon. Not. R. Astron. Soc., 353, 1286
Edge W. R. T., Coe M. J., McBride V. A., 2004d, The Astronomer's Telegram, 217, 1
Edge W. R. T., Coe M. J., Galache J. L., 2005a, The Astronomer's Telegram, 405, 1
Edge W. R. T., Coe M. J., Galache J. L., McBride V. A., Corbet R. H. D., Markwardt C. B.,
Laycock S., Marshall F. E., 2005b, The Astronomer's Telegram, 426, 1
Edge W. R. T., Coe M. J., Galache J. L., McBride V. A., Corbet R. H. D., Okazaki A. T.,
Laycock S., Markwardt C. B., Marshall F. E., Udalski A., 2005c, Mon. Not. R. Astron.
Soc., 361, 743
Edge W. R. T., 2005d, Ph.D. Thesis, University of Southampton
Fabrycky D., 2005, Mon. Not. R. Astron. Soc., 359, 117
Filipovi´c M. D., Haberl F., Pietsch W., Morgan D. H., 2000, Astron. Astrophys., 353, 129
Frank J., King A., Raine D. J., 2002, Accretion Power in Astrophysics: Third Edition.
Cambridge University Press
Galache J. L., Corbet R. H. D., Coe M. J., Laycock S., Markwardt C. B., Marshall F. E.,
2005, The Astronomer's Telegram, 674, 1
Grimm H.-J., Gilfanov M., Sunyaev R., 2003, Mon. Not. R. Astron. Soc., 339, 793
Haberl F., Filipovi´c M. D., Pietsch W., Kahabka P., 2000, Astron. Astrophys. Suppl. Ser.,
142, 41
Haberl F., Pietsch W., 2004, Astron. Astrophys., 414, 667
-- 89 --
Haberl F., Pietsch W., Kahabka P., 2007, The Astronomer's Telegram, 1095, 1 Haberl F.,
Pietsch W., 2005, Astron. Astrophys., 438, 211
Haberl F., Pietsch W., Schartel N., Rodriguez P., Corbet R. H. D., 2004, The Astronomer's
Telegram, 219, 1
Haberl F., Sasaki M., 2000, Astron. Astrophys., 359, 573
Horne J. H., Baliunas S. L., 1986, Astrophys. J., 302, 757
Hughes J. P., 1994, Astrophys. J., Lett., 427, L25
Imanishi K., Yokogawa J., Koyama K., 1998, IAU Circ., 7040, 2
Israel G. L., Campana S., Covino S., Dal Fiume D., Gaetz T. J., Mereghetti S., Oosterbroek
T., Orlandini M., Parmar A. N., Ricci D., Stella L., 2000, Astrophys. J., Lett., 531,
L131
Israel G. L., Stella L., Angelini L., White N. E., Giommi P., Covino S., 1997, Astrophys. J.,
Lett., 484, L141+
Israel G. L., Stella L., Campana S., Covino S., Ricci D., Oosterbroek T., 1998, IAU Circ.,
6999, 1
Jahoda K., Markwardt C. B., Radeva Y., Rots A., Stark M. J., Swank J. H., Strohmayer
T. E., Zhang W., 2005, ArXiv Astrophysics e-prints, astro-ph/0511531
Jahoda K., Swank J. H., Giles A. B., Stark M. J., Strohmayer T., Zhang W., Morgan E. H.,
1996, in Proc. SPIE Vol. 2808, p. 59-70, EUV, X-Ray, and Gamma-Ray Instrumen-
tation for Astronomy VII, Oswald H. Siegmund; Mark A. Gummin; Eds. In-orbit
performance and calibration of the Rossi X-ray Timing Explorer (RXTE) Propor-
tional Counter Array (PCA). pp 59 -- 70
Kahabka P., Pietsch W., 1998, IAU Circ., 6840, 1
Kaspi V. M., Johnston S., Manchester R. N., Bailes M., Bell J. F., Bessell M., Lyne A. G.,
D'Amico N., 1993, Bulletin of the American Astronomical Society, 25, 1434
Lamb R. C., Fox D. W., Macomb D. J., Prince T. A., 2002, Astrophys. J., Lett., 574, L29
Lamb R. C., Macomb D. J., Prince T. A., Majid W. A., 2002, Astrophys. J., Lett., 567,
L129
Lamb R. C., Prince T. A., Macomb D. J., Finger M. H., 1999, IAU Circ., 7081, 4
-- 90 --
Laycock S., Corbet R. H. D., Coe M. J., Marshall F. E., Markwardt C., Edge W., 2003,
Mon. Not. R. Astron. Soc., 339, 435
Laycock S., Corbet R. H. D., Coe M. J., Marshall F. E., Markwardt C., Lochner J., 2005,
Astrophys. J., Suppl. Ser., 161, 96
Laycock S., Corbet R. H. D., Perrodin D., Coe M. J., Marshall F. E., Markwardt C., 2002,
Astron. Astrophys., 385, 464
Laycock S. G. T., 2002, Ph.D. Thesis, University of Southampton
Lochner J. C., Marshall F. E., Whitlock L. A., Brandt N., 1998, IAU Circ., 6814, 1
Lochner J. C., Whitlock L. A., Corbet R. H. D., Marshall F. E., 1999a, American Astro-
nomical Society Meeting Abstracts, 194th Meeting
Lochner J. C., Whitlock L. A., Corbet R. H. D., Marshall F. E., 1999b, Bulletin of the
American Astronomical Society, 31, 742
Lomb N. R., 1976, Astrophys. Space. Sci., 39, 447
Macomb D. J., Finger M. H., Harmon B. A., Lamb R. C., Prince T. A., 1999, Astrophys.
J., Lett., 518, L99
Macomb D. J., Fox D. W., Lamb R. C., Prince T. A., 2003, Astrophys. J., Lett., 584, L79
Maeder A., Grebel E. K., Mermilliod J.-C., 1999, Astron. Astrophys., 346, 459
Majid W. A., Lamb R. C., Macomb D. J., 2004, Astrophys. J., 609, 133
Marshall F. E., Lochner J. C., Santangelo A., Cusumano G., Israel G. L., dal Fiume D.,
Orlandini M., Frontera F., Parmar A. N., Corbet R. H. D., 1998, IAU Circ., 6818, 1
Marshall F. E., Lochner J. C., Takeshima T., 1997, IAU Circ., 6777, 2
McGowan K. E., Coe M. J., Schurch M., McBride V. A., Galache J. L., Edge W. R. T.,
Corbet R. H. D., Laycock S., Udalski U., Buckley D. A. H., 2007, to appear in Mon.
Not. R. Astron. Soc.
Press W. H., Rybicki G. B., 1989, Astrophys. J., 338, 277
Sasaki M., Haberl F., Keller S., Pietsch W., 2001, Astron. Astrophys., 369, L29
Sasaki M., Pietsch W., Haberl F., 2003, Astron. Astrophys., 403, 901
-- 91 --
Scargle J. D., 1982, Astrophys. J., 263, 835
Schmidtke P. C., Cowley A. P., Levenson L., Sweet K., 2004, Astron. J., 127, 3388
Schmidtke P. C., Cowley A. P., 2005a, The Astronomer's Telegram, 648, 1
Schmidtke P. C., Cowley A. P., 2005b, Astron. J., 130, 2220
Schmidtke P. C., Cowley A. P., 2006, Astron. J., 132, 919
Schmidtke P. C., Cowley A. P., Udalski A., 2006, Astron. J., 132, 971
Schmidtke P. C., Cowley A. P., 2007, The Astronomer's Telegram, 1181, 1
Schmidtke P. C., Cowley A. P., 2007, American Astronomical Society Meeting Abstracts,
211, #03.06
Schmidtke P. C., Cowley A. P., Udalski A., 2007, The Astronomer's Telegram, 1316, 1
Southwell K. A., Charles P. A., 1996, Mon. Not. R. Astron. Soc., 281, L63
Stanimirovic S., Staveley-Smith L., Dickey J. M., Sault R. J., Snowden S. L., 1999, Mon.
Not. R. Astron. Soc., 302, 417
Stella L., White N. E., Rosner R., 1986, Astrophys. J., 308, 669
Stevens J. B., Coe M. J., Buckley D. A. H., 1999, Mon. Not. R. Astron. Soc., 309, 421
Torii K., Yokogawa J., Imanishi K., Koyama K., 2000, IAU Circ., 7428, 3
Tsujimoto M., Imanishi K., Yokogawa J., Koyama K., 1999, Publ. Astron. Soc. Jpn., 51,
L21
Ueno M., Yamaguchi H., Takagi S.-I., Yokogawa J., Koyama K., 2004, Publ. Astron. Soc.
Jpn., 56, 175
Wilson C., Finger M., 1998, IAU Circ., 7048, 1
Wilson C. A., Finger M. H., Coe M. J., Negueruela I., 2003, Astrophys. J., 584, 996
Yokogawa J., Imanishi K., Koyama K., Nishiuchi M., Mizuno N., 2002, Publ. Astron. Soc.
Jpn., 54, 53
Yokogawa J., Imanishi K., Tsujimoto M., Koyama K., Nishiuchi M., 2003, Publ. Astron.
Soc. Jpn., 55, 161
-- 92 --
Yokogawa J., Imanishi K., Ueno M., Koyama K., 2000, Publ. Astron. Soc. Jpn., 52, L73
Yokogawa J., Torii K., Imanishi K., Koyama K., 2000, Publ. Astron. Soc. Jpn., 52, L37
Yokogawa J., Koyama K., 1998a, IAU Circ., 6853, 2
Yokogawa J., Koyama K., 1998b, IAU Circ., 7009, 3
Yokogawa J., Koyama K., 1998c, IAU Circ., 7028, 1
Yokogawa J., Torii K., Kohmura T., Imanishi K., Koyama K., 2000, Publ. Astron. Soc. Jpn.,
52, L53
This preprint was prepared with the AAS LATEX macros v5.2.
|
0802.0151 | 1 | 0802 | 2008-02-01T16:07:11 | Boxy/peanut and discy bulges : formation, evolution and properties | [
"astro-ph"
] | The class `bulges' contains objects with very different formation and evolution paths and very different properties. I review two types of `bulges', the boxy/peanut bulges (B/Ps) and the discy bulges. The former are parts of bars seen edge-on, have their origin in vertical instabilities of the disc and are somewhat shorter in extent than bars. Their stellar population is similar to that of the inner part of the disc from which they formed. Discy bulges have a disc-like outline, i.e., seen face-on they are circular or oval and seen edge-on they are thin. Their extent is of the order of 5 times smaller than that of the boxy/peanut bulges. They form from the inflow of mainly gaseous material to the centre of the galaxy and from subsequent star formation. They thus contain a lot of young stars and gas. Bulges of different types often coexist in the same galaxy. I review the main known results on these two types of bulges and present new simulation results. B/Ps form about 1Gyr after the bar, via a vertical buckling. At that time the bar strength decreases, its inner part becomes thicker -- forming the peanut or boxy shape -- and the ratio $\sigma_z^2/\sigma_r^2$ increases. A second buckling episode is seen in simulations with strong bars, also accompanied by a thickening of the peanut and a weakening of the bar. The properties of the B/Ps correlate strongly with those of the bar: stronger bars have stronger peanuts, a more flat-topped vertical density distribution and have experienced more bucklings. I also present simulations of disc galaxy formation, which include the formation of a discy bulge. Decomposition of their radial density profile into an exponential disc and a Sersic bulge gives realistic values for the disc and bulge scale-lengths and mass ratios, and a Sersic shape index of the order of 1. | astro-ph | astro-ph | Proceedings Title IAU Symposium
Proceedings IAU Symposium No. 245, 2008
A.C. Editor, B.D. Editor & C.E. Editor, eds.
c(cid:13) 2008 International Astronomical Union
DOI: 00.0000/X000000000000000X
Boxy/peanut and discy bulges : formation,
evolution and properties
E. Athanassoula
Laboratoire d'Astrophysique de Marseille, Observatoire Astronomique de Marseille Provence,
2 place Le Verrier, 13248 Marseille cedex 04, France
Abstract. The class 'bulges' contains objects with very different formation and evolution paths
and very different properties. I review two types of 'bulges', the boxy/peanut bulges (B/Ps)
and the discy bulges. The former are parts of bars seen edge-on, have their origin in vertical
instabilities of the disc and are somewhat shorter in extent than bars. Their stellar population is
similar to that of the inner part of the disc from which they formed. Discy bulges have a disc-like
outline, i.e., seen face-on they are circular or oval and seen edge-on they are thin. Their extent is
of the order of 5 times smaller than that of the boxy/peanut bulges. They form from the inflow
of mainly gaseous material to the centre of the galaxy and from subsequent star formation. They
thus contain a lot of young stars and gas. Bulges of different types often coexist in the same
galaxy. I review the main known results on these two types of bulges and present new simulation
results.
z/σ2
B/Ps form about 1Gyr after the bar, via a vertical buckling. At that time the bar strength
decreases, its inner part becomes thicker -- forming the peanut or boxy shape -- and the ratio
σ2
r increases. A second buckling episode is seen in simulations with strong bars, also accom-
panied by a thickening of the peanut and a weakening of the bar. The properties of the B/Ps
correlate strongly with those of the bar: stronger bars have stronger peanuts, a more flat-topped
vertical density distribution and have experienced more bucklings.
I also present simulations of disc galaxy formation, which include the formation of a discy
bulge. Decomposition of their radial density profile into an exponential disc and a S´ersic bulge
gives realistic values for the disc and bulge scale-lengths and mass ratios, and a S´ersic shape
index of the order of 1.
It is thus clear that classical bulges, B/P bulges and discy bulges are three distinct classes
of objects and that lumping them together can lead to confusion. To avoid this, the two latter
could be called B/P features and inner discs, respectively.
Keywords. galaxies: bulges, galaxies: evolution, galaxies: formation, galaxies: kinematics and
dynamics, stellar dynamics, methods: N-body simulations
1. Introduction
What is a bulge? Three different definitions have been used so far, based on morphol-
ogy, photometry, or kinematics, respectively. According to the morphological definition,
a bulge is the component of a disc galaxy that swells out of the central part of a disc
viewed edge-on. Based on photometry, a bulge is the extra light in the central part of the
galaxy, over and above the exponential profile fitting the remaining (non central) part of
the disc. The third definition is based on kinematics, and in particular on the value of
V /σ, or, more specifically, on the location of the object on the (V /σ, ellipticity) diagram
(often referred to as the Binney diagram (Binney 1978, 2005)). These three definitions
are compared and discussed in Athanassoula & Martinez-Valpuesta (2007).
The lack of a single, clear-cut definition of a bulge, although historically understand-
able, has led to considerable confusion and to the fact that bulges are an inhomoge-
neous class of objects. For this reason, Kormendy (1993; see also Kormendy & Kenni-
1
8
0
0
2
b
e
F
1
]
h
p
-
o
r
t
s
a
[
1
v
1
5
1
0
.
2
0
8
0
:
v
i
X
r
a
2
E. Athanassoula
cutt 2004, hereafter KK04) distinguished classical bulges from pseudo-bulges. However,
pseudo-bulges by themselves are also an inhomogeneous class of objects, as argued by
Athanassoula (2005a, hereafter A05), who distinguishes three types of objects which are,
according to the above definitions, classified as bulges. Classical bulges are formed
by gravitational collapse or hierarchical merging of smaller objects and corresponding
dissipative gas processes. Their morphological, photometrical and kinematical proper-
ties are similar to those of ellipticals. They are discussed extensively in other papers in
these proceedings and are not the subject of this review. The two other types of bulges
are boxy/peanut bulges (B/P), and discy bulges, which will be discussed here. As
stressed in A05, different types of bulges often co-exist and it is possible to find all three
types of bulges in the same simulation, or in the same galaxy.
2. Boxy/peanut bulges
Viewed edge-on, disc galaxies often have a central component which swells out of the
disc and whose outline is not elliptical, but has a boxy, or peanut, or even 'X' shape. Due
to the morphological definition of a bulge, such components have been called bulges, or,
more specifically, boxy/peanut bulges (B/Ps).
The formation of B/Ps has been witnessed in a large number of numerical simulations
(Combes & Sanders 1981; Combes et al. 1990; Raha et al. 1991; Athanassoula & Misiriotis
2002, hereafter AM02; Athanassoula 2003, hereafter A03; A05; O'Neil & Dubinski 2003;
Debattista et al. 2004; Martinez-Valpuesta & Shlosman 2004; Debattista et al. 2006;
Martinez-Valpuesta, Shlosman & Heller 2006, etc). It is linked to the vertical instability
of parts of the main family of periodic orbits constituting the bar, widely known as the
x1 family (Binney 1981; Pfenniger 1984; Skokos, Patsis & Athanassoula 2002; Patsis,
Skokos & Athanassoula 2002). The stability of the x1 family can be followed from the
corresponding stability diagram (see e.g. figures 3 and 4 of Skokos et al. 2002) which shows
that, at the positions where the x1 becomes unstable, other families bifurcate. These are
linked to the n : 1 vertical resonances and extend well outside the disc equatorial plane.
As shown by Patsis et al. (2002), some of them are very good building blocks for the
formation of peanuts, because they are stable and because their orbits have the right
shape, extent and location. Studies of these orbits reproduced many of the B/P properties
and helped explaining crucial aspects of B/P formation and evolution. For example, an
analysis of the orbital families that constitute peanuts predicts that B/Ps should be
shorter than bars. This is indeed found to be the case both in N -body simulations and
in real galaxies (Lutticke, Dettmar & Pohlen 2000; A05; Athanassoula & Beaton 2006).
2.1. Time evolution
The time evolution of the bar, of the buckling and of the peanut strengths are plotted in
Fig. 1 for a simulation which develops a strong bar. The time is given in Gyrs, using the
calibration proposed in AM02. The initially unbarred disc forms a bar roughly between
times 3 and 4 Gyrs (lower panel). I define as bar formation time the time at which the
bar-growth is maximum (i.e. when the slope of the bar strength as a function of time is
maximum) and indicate it by the first vertical line in Fig. 1. The bar strength reaches a
maximum at a time indicated by the second vertical line, and then decreases considerably
over ∼ 1 Gyr. The time at which the bar amplitude decrease is maximum is given by the
third vertical line. Subsequently, the bar strength reaches a minimum, at a time shown
by the fourth vertical line, and then starts increasing again at a rate much slower than
that during bar formation.
The upper panel shows the buckling strength, i.e. the vertical asymmetry as a function
Boxy/peanut and discy bulges : formation, evolution and properties
3
Figure 1. Time evolution of three peanut-, or bar-related quantities, namely the buckling
strength (i.e. the vertical asymmetry; upper panel), the peanut strength (i.e. its vertical extent;
middle panel) and the bar strength (lower panel). The solid vertical lines mark characteristic
times linked to bar formation and evolution. From left to right, these are the bar formation
time, the maximum amplitude time, the bar decay time and the bar minimum amplitude time
(see text). The vertical dashed line marks the time of the buckling.
of time. The disc is vertically symmetric before and during bar formation and the first
indications of asymmetry occur only after the bar amplitude has reached a maximum.
The asymmetry then grows very abruptly to a strong, clear peak and then drops equally
abruptly. The time of the buckling (dashed vertical line) is given by the peak of this curve
and is very clearly defined. It is important to note that, to within the measuring errors,
it coincides with the time of bar decay (third vertical line). This is not accidental. I
verified it for a very large number of simulations and thus can establish the link between
the buckling episode and the decay of the bar strength (Raha et al. 1991; Martinez-
Valpuesta & Shlosman 2004).
The middle panel shows the strength of the peanut, i.e. its vertical extent, again as a
function of time. This quantity grows abruptly after the bar has reached its maximum
amplitude and during the time of the buckling. This abrupt growth is followed by a much
slower increase over a longer period of time. Taken together, the three panels of Fig. 1
show that the bar forms vertically thin, and only after it has reached a maximum strength
does the buckling phase occur. During the buckling time the bar strength decreases
significantly, while the the B/P strength increases. The time intervals during which bar
formation, peanut formation, or buckling occur are all three rather short, of the order of
a Gyr, and they are followed by a longer stretch of time during which the bar and B/P
evolve much slower. This later evolution is often referred to as secular evolution.
4
E. Athanassoula
This particular simulation has a second, weaker buckling episode shortly after 8 Gyrs.
This occurs very often in simulations developing strong bars and was discussed first by
Athanassoula (2005b) and Martinez-Valpuesta et al. (2006). It is seen clearly in all three
panels and has characteristics similar to those of the first buckling.
2.2. Peanut formation and collective effects
As already discussed, orbital structure theory explains B/P formation by the vertical
instabilities of the main family of bar-supporting periodic orbits. An alternative approach
explains the buckling and the peanut formation as due to the bending, or fire-hose,
instability, studied analytically in the linear regime (Toomre 1966; Araki 1985). These
studies assign a critical value to the ratio Rσ = σ2
r igniting the onset of the instability,
which is around 0.1. A number of simulations, however, have shown that the vertical
instability sets in at much larger values of Rσ (e.g. Merritt & Sellwood 1994; Sotnikova
& Rodionov 2003).
z /σ2
Figure 2. Time evolution of the ratio Rσ = σ2
r . The thin vertical lines mark two character-
istic times linked to bar formation and evolution. From left to right, these are the bar maximum
and minimum amplitude times, corresponding to the first buckling episode (see Sect. 2.1 and
Fig. 1).
z/σ2
To test this hypothesis, I calculate the radial and z components of the disc velocity
dispersion as a function of radius (averaging over azimuth and height). I then find the
minimum value of their ratio Rσ and plot its time evolution in Fig. 2. The thin vertical
lines mark two characteristic times linked to bar formation and evolution -- namely the
bar maximum and minimum amplitude times -- as found from Fig. 1. Their location
is clearly linked to changes in behaviour of Rσ. This, however, does not necessarily
imply that the changes in Rσ are the cause of the buckling, but can also be seen as
its consequence. Indeed, as the bar forms σr increases drastically, so that Rσ decreases.
Then the bar amplitude reaches its maximum and starts decreasing, while the peanut
starts forming. During this time, σr decreases while σz increases. As a result, the ratio
Rσ reaches a minimum when the bar amplitude is maximum and then increases again, as
is indeed seen in Fig. 2. Then the bar amplitude reaches a minimum, which corresponds
to a minimum of σr and therefore to a maximum of Rσ. This is followed by a slower
decrease of Rσ, which is stopped by the second buckling episode. The value of Rσ at
which this instability sets in is much less extreme than that predicted by the above
mentioned analytical works, but is in good agreement with other N -body simulations.
More work is necessary before we fully understand the respective roles of the orbital
structure results and of the velocity anisotropy effects on the formation and evolution of
B/P structures. Both explain part of the story, but many aspects of their interplay are
still unclear. Orbital structure results tell us whether the appropriate building blocks are
available, or not, and this is essential, since the lack of such building blocks prohibits the
Boxy/peanut and discy bulges : formation, evolution and properties
5
formation of a given structure. Furthermore, studies of the properties of the building-
block orbits are essential for understanding the properties of the B/P structures. Orbital
structure theory, however, can not tell us how much matter is trapped around a given
orbit or family. Furthermore, it is necessary to group all these building blocks into one co-
herent, self-consistent unit and here collective effects are essential. Like orbital structure,
they also can set limits on the formation of B/P structures, as well as give information
on their properties. The respective input from the two methods will be discussed further
elsewhere.
2.3. Comparison with observations
The fact that B/Ps are just parts of bars seen edge-on was not immediately accepted
(see e.g. Kormendy 1993). The main arguments against it were, however, refuted in
A05, with the help of orbital structure results. Furthermore, considerable observational
evidence argues in its favour, particularly detailed comparisons between observations and
simulations.
Radial density profiles from simulations, taken along slits on, or parallel to, the equa-
torial plane when the galaxy is seen edge-on (AM02; A05) have the same characteristic
signatures as the corresponding radial light profiles (Lutticke, Dettmar & Pohlen 2000;
Bureau et al. 2006). Similarly, density profiles along cuts perpendicular to the equato-
rial plane (AM02; A05) show similar characteristics to analogous observed light profiles
(Aronica et al. 2003 and this volume). Further tests come from comparisons of median
filtered images of B/P systems (Bureau et al. 2006) to similar images of N -body bars
(A05). These show the same types of characteristic features, namely four extensions out
of the equatorial plane, which form an X-like shape, except that the four extensions do
not necessarily cross the centre. Another common feature is maxima of the density along
the equatorial plane, away from the centre and diametrically opposite. Starting from the
centre of the galaxy and going outwards along the equatorial plane, the projected surface
density first drops, then increases again to reach a local maximum and then decreases
again to the edge of the disc.
Considerable evidence was also accumulated using kinematical observations. Cylindri-
cal rotation, witnessed in a number of B/P galaxies (KK04 and references therein), is
also seen in velocity fields of strong N -body bars viewed edge-on (Combes et al. 1990;
AM02). Emission line spectroscopy of boxy/peanut galaxies (Kuijken & Merrifield 1995;
Bureau & Freeman1999; Merrifield & Kuijken 1999) shows that their major axis posi-
tion velocity diagrams (PVDs) have a number of interesting features, well reproduced by
gas flow simulations (Athanassoula & Bureau 1999). In particular, the shocks along the
leading edges of the bar and the corresponding inflow lead to a characteristic gap in the
PVDs, between the signature of the nuclear spiral (whenever existent) and the signature
of the disc.
Comparison of long-slit absorption line spectra (Chung & Bureau 2004) of galaxies
with B/Ps to similar 'observations' of N -body bars viewed edge-on reveals that the two
have the same characteristic features (Bureau & Athanassoula 2005). The integrated light
along the slit (equivalent to a major-axis light profile) has a quasi-exponential central
peak and a plateau at intermediate radii, followed by a steep drop. The rotation curve
(V (r)) has a characteristic double hump. The velocity dispersion has a central peak,
which in the centre-most part may be rather flat or may even have a central minimum.
At intermediate radii there can be a plateau which sometimes ends on either side with
a shallow maximum before dropping steeply at larger radii. h3 (i.e. the coefficient of the
third order term in a Gauss-Hermite expansion of the line of sight velocity distribution)
correlates with V over most of the bar length, contrary to what is expected for a fast
6
E. Athanassoula
rotating disc. All these features are spatially correlated and are seen, more or less strongly,
both in the observations and in the simulations (Bureau & Athanassoula 2005).
2.4. The effect of the halo
Figure 3. Histograms of number of bucklings, bar strength, B/P strength and B/P shape for a
sample of fully self-consistent high-resolution simulations. The histograms include simulations
with halo cores of all sizes and the hatched areas includes only simulations with small halo cores.
In collaboration with Martinez-Valpuesta, I made an extensive statistical study of a
few hundred simulations which I had run for different purposes (AM02; A03 etc). We
measured the strength of the peanut (from its thickness) and its shape (from the shape
of the density profile on cuts perpendicular to the equatorial plane) and found that
both correlate well with the bar strength. Thus, stronger bars have stronger peanuts and
more flat-topped vertical density profiles. I also find that the type of halo plays a major
role in determining the properties of the B/P. Fig. 3 shows histograms of the number
of bucklings that have occurred, of the bar strength, of the B/P strength and of the
B/P shape, distinguishing between simulations with small halo cores and simulations
with large halo cores. The two populations are indeed very different. Simulations with
small halo cores have stronger bars, stronger peanuts, more flat-topped vertical density
profiles and have experienced more bucklings than simulations with large cores. The few
simulations with small cores which have weak bars, weak B/Ps and did not buckle have
either a very hot halo or a very hot disc. Simulations with cuspy haloes (not shown here)
have yet weaker B/Ps and smaller number of bucklings and will be discussed elsewhere.
The above can be explained by the fact that the halo plays a major role in determining
the properties of the bar (AM02; A03). Athanassoula (2002, hereafter A02) showed that
angular momentum is primarily emitted by near-resonant material in the bar region
and absorbed by near-resonant material in the outer disc and, particularly, in the halo,
Boxy/peanut and discy bulges : formation, evolution and properties
7
while A03 showed that bars grow stronger when more angular momentum is exchanged
within the galaxy. Furthermore, as shown in AM02 and as explained in A02 and A03,
the size of the halo core strongly influences the bar evolution. Haloes with a small core
have a lot of mass in the inner regions and thus, provided their velocity dispersion is
not too high, can provide substantial angular momentum sinks and lead to considerable
angular momentum exchange between the near-resonant particles in the bar region and
the near-resonant particles in the halo. Such models grow strong bars (long, thin and
massive) with rectangular-like isodensities (AM02). Viewed side-on (i.e. edge-on with
the line-of-sight along the bar minor axis) they exhibit a strong peanut, or even X-like
shape. If, however, the velocity dispersion in the disc and/or halo is too high, the angular
momentum exchange is hindered and the bar and peanut will be weak (A03). Haloes with
large cores have considerably less material in the inner parts and are thus capable of less
angular momentum exchange. Bars grown in such environments are less strong, have
elliptical-like isodensities when viewed face-on and boxy-like when viewed side-on. All
these considerations explain the results found in Fig. 3, namely the difference between
the histograms for simulations with small halo cores and simulations with large halo
cores. They also explain the weak bars and B/Ps found in some simulations with small
halo cores.
3. Disc-like bulges
Disc-like bulges form from inflow of (mainly) gas material to the centre of the galaxy
and subsequent star formation. This inflow is due to the torques exerted by a non-
axisymmetric component, usually a large-scale bar, as witnessed in hydrodynamic simu-
lations (e.g. Athanassoula 1992; Friedli & Benz 1993; Heller & Shlosman 1994; Wada &
Habe 1995). The high density of the gas accumulated in the inner regions triggers very
strong star formation. Such inflow, however, is also seen in N -body simulations (AM02;
Valenzuela & Klypin 2003), which represent the old stellar population. Thus, this inner
region is not only a region of increased density for the gas and the young stars, but also for
the older stellar populations. This should lead to the formation of an inner, central com-
ponent of disc-like shape, whose extent is of the order of a kpc and which is constituted
mainly of gas and young stars, but also of older stars. This was named disc-like bulge, or,
for short, discy bulge in A05 and is often observed in disc galaxies. Due to its disc-like
shape, it often has spirals or inner bars (KK04 and references therein). It stands out very
clearly in radial photometric profiles, whose decomposition shows it is well represented
by a S´ersic law (S´ersic 1968). Contrary to classical bulges, however, it does not swell
out of the galactic plane. This is not the only difference between disc-like and classical
bulges. Disc-like bulges have a S´ersic index of the order of 1, i.e. much smaller than the
values found for classical bulges (KK04 and references therein). They also have different
kinematics, like that of discs, a higher fraction of young stars and a higher gas content.
A lot of data on such bulges has been collected over the last few years, but still much
work, particularly theoretical, is necessary before we fully understand these objects.
In order to describe adequately the formation and evolution of disc-like bulges, simu-
lations should include gas, star formation and feedback, all in a realistic way. It would,
furthermore, be preferable if they started from cosmological or cosmologically-motivated
initial conditions, since the properties of pre-existing discs may influence the proper-
ties of the disc-like bulges. I will briefly describe here results from simulations following
this outline (Athanassoula, Heller & Shlosman, in preparation). For information on the
numerical techniques used in these simulations and an initial discussion of some of the
results see Heller, Shlosman & Athanassoula (2007a) and (2007b).
8
E. Athanassoula
Figure 4. Properties of a simulated disc-like bulge. Left : Radial projected density profile in
arbitrary units. Radii are measured in kpc. The dots give the simulation results and the straight
line the fit by an exponential disc and a S´ersic component. Right : Measure of the vertical height
of the material near the equatorial plane (see text), as a function of radius, measured in kpc.
Several non-axisymmetric components -- such as a triaxial halo, oval disc, inner and
outer bar -- form during these simulations. Their interactions give very interesting dy-
namical phenomena (Heller et al. 2007a; b), while they induce considerable inflow and
gaseous high density inner discs. As in the sketchy outline in the beginning of this sec-
tion, the high gas concentration in the central area triggers considerable star formation,
resulting in a disc-like central, high-density object, which, seen face-on, is often some-
what oval. It has many properties similar to those of discy-bulges. For example, it has, in
many cases, sub-structures, like an inner bar. In order to further assess the properties of
the disc galaxy formed in these simulations and to better establish the link with disc-like
bulges, I chose a characteristic specific snapshot, i.e. a characteristic specific simulation
and time, and examine its mass distribution in order to compare best with the observed
light distribution in galaxies. An analysis of the kinematics, together with a statistical
treatment, including other times and other simulations, will be given elsewhere. The
radial projected surface density profile of the snapshot under consideration is given in
Fig. 4, together with a fit by an exponential disc and a S´ersic component. Note that
the fit is excellent, all the way to the outer parts of the disc, roughly at 10 kpc. In this
example, the disc scale-length is ∼2.7 kpc, i.e. very realistic, while the S´ersic index is ∼1,
in good agreement with observed discy bulges.
Figure 5. Edge-on view of the stellar component of the simulation with a disc and a disc-like
bulge. The projected density is given by grey-scale and also by five isocontours whose level is
picked so as to show best the features under consideration.
Fig. 5 shows the snapshot seen edge-on. The three outer isodensity curves show clearly
that the shape and aspect ratio of the disc component is very realistic. The two innermost
contours (within 1 kpc) reveal the existence of a small, central, disc-like object, the
Boxy/peanut and discy bulges : formation, evolution and properties
9
vertical height of which we need to quantify. Measuring the average thickness would not
be useful, since this is due to both the external big disc and the small inner component, so
I proceeded differently. I divided the 'stars' in the snapshot into circular annuli, according
to their distance from the centre and, in each annulus, sorted them as a function of their
distance from the equatorial plane (z). Since, statistically, the 'stars' in the disc-like
inner component will have smaller z values than the ones in the outer disc, I plot in
Fig. 4 the z component of the 'star' with rank 0.3Nan, where Nan is the total number
of 'stars' in the annulus. This shows a deep minimum in the central region, as one would
expect due to the existence of an inner disc with a shorter vertical extent than the outer
one. It also shows that the region where the inner disc is contributing significantly is of
the order of 1 kpc, in good agreement with the radial density profile (Fig. 4). Finally,
the aspect ratio of the inner and the outer discs are similar.
To summarise, in our fiducial simulation, as well as in several others, we witness inflow
of mainly gas material to the central regions and strong subsequent star formation. Thus,
an inner disc is formed, composed of both stars and gas. Its radial extent is of the order
of a kpc and its vertical extent much smaller than that of the outer disc. This disc can
harbour spiral structure, or an inner bar. Its contribution to the radial projected density
profile is well fitted by a S´ersic law with S´ersic index ∼1. It is thus very likely that this
simulation describes correctly the formation of discy-like bulges in galaxies.
4. Summary and discussion
I briefly reviewed the formation, evolution and properties of boxy/peanut bulges and of
disc-like bulges. These two types of objects have very different formation and evolutionary
histories and very different properties. B/Ps form from vertical instabilities and their
building blocks are the 3D families associated with the 3D bifurcations of the x1 family.
Discy bulges form from the inflow of (mainly) gas material and from the ensuing enhanced
star formation. Thus B/Ps are mainly constituted of inner disc stars, while the discy
bulges have a very large contribution from gas and young stars. Since the formation of
discy bulges relies on the gas inflow, it is expected that they will be found mainly in
late type disc galaxies, as is indeed the case. The face-on extent of the B/Ps is of the
order of five times larger than that of the discy bulges and, seen edge-on, they extend
well outside the equatorial plane, while the discy bulges are thin. Their kinematics and
their contribution to the radial photometric profiles are different from those of discy
bulges. Thus, one should clearly distinguish between B/Ps and discy-bulges and not
lump together them in a single category.
Once this has become clear, one may also wish to revise the existent nomenclature
in order to avoid some of the present confusion. Boxy/peanut bulges could be called
boxy/peanut features (or structures), or simply peanuts, as proposed in A05. This would
make it clearer that they are just a part or a feature of the bar and not an independent
entity. Similarly, discy bulges could be simply called inner discs. Then the name 'bulge'
would be reserved for classical bulges. This change, however, will also necessitate changing
the bulge definitions described in Sect. 1.
Acknowledgements
I thank my collaborators, A. Bosma, A. Aguerri, C. Heller, I. Martinez-Valpuesta and
I. Shlosman, for interesting and fruitful discussions. This work was partially supported
by grant ANR-06-BLAN-0172 and by the Gruber foundation.
10
E. Athanassoula
References
Araki, S. 1985 PhD thesis, MIT
Aronica, G., Athanassoula, E., Bureau, M., Bosma, A., Dettmar, R.-J., Vergani, D., Pohlen, M.
2003, Ap&SS, 284, 753
Athanassoula, E. 1992, MNRAS, 259, 345
Athanassoula, E. 2002, ApJL, 569, 83 (A02)
Athanassoula, E. 2003, MNRAS, 341, 1179 (A03)
Athanassoula, E. 2005a, MNRAS, 358, 1477 (A05)
Athanassoula, E. 2005b, in Planetary Nebulae as Astronomical Tools, eds. R. Szczerba, G.
Stasi´nska & S. K. G´orny, AIP Conf. Proc. 804, Melville, New York, 333
Athanassoula, E., Beaton, R. L. 2006, MNRAS, 370, 1499
Athanassoula, E., Bureau, M. 1999, ApJ, 522, 699
Athanassoula, E., Misiriotis, A. 2002, MNRAS, 330, 35 (AM02)
Athanassoula, E., Martinez-Valpuesta 2007, in Pathways through an eclectic Universe, eds. J.
H. Knapen, T. J. Mahoney, and A. Vazdekis, ASP Conf. Ser., 2007, in press and astro-
ph/0710.1518
Binney, J. 1978, MNRAS, 183, 501
Binney, J. 1981, MNRAS, 196, 455
Binney, J. 2005, MNRAS, 363, 937
Bureau, M., Aronica, G., Athanassoula, E., Dettmar, R.-J., Bosma, A., Freeman, K. C. 2006,
MNRAS, 370, 753
Bureau, M., Athanassoula, E. 2005, ApJ, 626, 159
Bureau, M., Freeman, K. C. 1999, AJ, 118, 126
Chung, A., Bureau, M. 2004, AJ, 127, 3192
Combes, F., Sanders, R. H. 1981, A&A, 96, 164
Combes, F., Debbasch, F., Friedli, D., Pfenniger, D. 1990, A&A, 233, 82
Debattista, V. P., Carollo, M., Mayer, L., Moore, B. 2004, ApJ, 604, L93
Debattista, V. P., Carollo, M., Mayer, L. Moore, B., Wadsley, J., Quinn, T. 2006, ApJ, 645, 209
Friedli, D., Benz, W. 1993, A&A, 268, 65
Heller, C. H., Shlosman, I. 1994, ApJ, 424, 84
Heller, C.H., Shlosman, I., Athanassoula, E. 2007a, ApJ, 657, L65
Heller, C.H., Shlosman, I., Athanassoula, E. 2007b, ApJ, in press and astro-ph 0706.3895
Kormendy, J. 1993, in Galactic Bulges, eds. H. Dejonghe and H. J. Habing, Kluwer Academic
Publ., IAU Symposium 153, 209
Kormendy, J., Kennicutt, R. C. 2004, ARA&A, 42, 603 (KK04)
Kuijken, K., Merrifield, M. R. 1995, ApJ, 443, L13
Lutticke, R., Dettmar, R.-J., Pohlen, M. 2000, A&A, 362, 435
Martinez-Valpuesta, I., Shlosman, I. 2004, ApJ, 613, 29
Martinez-Valpuesta, I., Shlosman, I., Heller, C. 2006, ApJ, 637, 214
Merrifield, M. R., Kuijken, K. 1999, A&A, 345, L47
Merritt, D., Sellwood, J. A. 1994, ApJ, 425, 551
O'Neill, J. K., Dubinski, J. 2003, MNRAS, 346, 251
Patsis, P., Skokos, Ch., Athanassoula, E. 2002, MNRAS, 337, 578
Pfenniger, D. 1984, A&A, 134, 373
Raha, N., Sellwood, J. A., James, R. A., Kahn, F. D. 1991, Nature, 352, 411
S´ersic, J. 1968, Atlas de Galaxias Australes, Obs. Astron. Cordoba
Skokos, H., Patsis, P., Athanassoula, E. 2002, MNRAS, 333, 847
Sotnikova, N. Ya., Rodionov, S. A. 2003, AstL, 29, 321
Toomre, A. 1966, Geophys. Fluid Dyn., N66-46, 111
Valenzuela, O., Klypin 2003, MNRAS, 345, 406
Wada, K., Habe, A. 1995, MNRAS, 277, 433
|
astro-ph/0112505 | 3 | 0112 | 2002-02-25T23:44:05 | Coherent Radio Pulses From GEANT Generated Electromagnetic Showers In Ice | [
"astro-ph",
"hep-ph"
] | Radio Cherenkov radiation is arguably the most efficient mechanism for detecting showers from ultra-high energy particles of 1 PeV and above. Showers occuring in Antarctic ice should be detectable at distances up to 1 km. We report on electromagnetic shower development in ice using a GEANT Monte Carlo simulation. We have studied energy deposition by shower particles and determined shower parameters for several different media, finding agreement with published results where available. We also report on radio pulse emission from the charged particles in the shower, focusing on coherent emission at the Cherenkov angle. Previous work has focused on frequencies in the 100 MHz to 1 GHz range. Surprisingly, we find that the coherence regime extends up to tens of Ghz. This may have substantial impact on future radio-based neutrino detection experiments as well as any test beam experiment which seeks to measure coherent Cherenkov radiation from an electromagnetic shower. Our study is particularly important for the RICE experiment at the South Pole. | astro-ph | astro-ph |
Coherent Radio Pulses From GEANT Generated
Electromagnetic Showers In Ice
Soebur Razzaque,a Surujhdeo Seunarine,b
David Z. Besson,a Douglas W. McKay,a John P. Ralston, a
and David Seckel c
aDepartment of Physics & Astronomy
University of Kansas, Lawrence, KS 66045, USA
b Department of Physics
University of Canterbury, Private Bag 4800, Christchurch, New Zealand
c Bartol Research Institute
University of Delaware, Newark, DE 19716, USA
Abstract
Radio Cherenkov radiation is arguably the most efficient mechanism for detecting
showers from ultra-high energy particles of 1 PeV and above. Showers occuring in
Antarctic ice should be detectable at distances up to 1 km. We report on electromag-
netic shower development in ice using a GEANT Monte Carlo simulation. We have
studied energy deposition by shower particles and determined shower parameters for
several different media, finding agreement with published results where available. We
also report on radio pulse emission from the charged particles in the shower, focusing
on coherent emission at the Cherenkov angle. Previous work has focused on frequen-
cies in the 100 MHz to 1 GHz range. Surprisingly, we find that the coherence regime
extends up to tens of Ghz. This may have substantial impact on future radio-based
neutrino detection experiments as well as any test beam experiment which seeks to
measure coherent Cherenkov radiation from an electromagnetic shower. Our study is
particularly important for the RICE experiment at the South Pole.
Contents
1 Introduction
2 Shower Description
2.1 Energy Loss Mechanisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2 Longitudinal Profile . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.3 Lateral Spread . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3 Shower Simulations
3.1 Radiation Length . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1
2
4
4
5
6
7
7
8
3.2 Moliere radius . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
9
3.3 Energy Loss in the Medium . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.4 Critical Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
3.5 Track Lengths . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
3.6 Particle yield . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
3.7
1-D Model Study . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
4 Shower Fluctuation
4.1 Fit to Simulations
19
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
5 Electromagnetic Pulse Theory
21
5.1 Fraunhoffer Limit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
5.2 Parametrization Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
5.3 Radiated Power and The Form Factor . . . . . . . . . . . . . . . . . . . . . 26
6 Electric Field Pulse Calculation
28
6.1 Vector Superposition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
6.2 Phase Coherence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
6.3 Factorization of the Electric Field, and the Discrete Form Factor . . . . . . 31
6.4 The Discrete Form Factor and the Frequency Spectrum . . . . . . . . . . . . 32
6.5 The Analytic Form Factor and the Frequency Spectrum . . . . . . . . . . . . 32
6.6 Direct Calculation: Monte Carlo Field Spectrum . . . . . . . . . . . . . . . 34
6.7 Related Issues . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
6.8 Angular Pulse Distribution . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
7 Summary of Results and Conclusions
39
7.1 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
1
Introduction
Ultra high energy (UHE) neutrinos can travel without scattering over large distances.
These may prove to be useful cosmological and astrophysical probes. They also present
themselves as candidate high energy particles with which we can test the Standard Model
of electro-weak theory beyond the energy regime of current accelerators.
In an UHE
electron-neutrino charged current interaction, the neutrino gives most (≈ 80%) of its
energy to the secondary electron, which can then initiate an electromagnetic cascade or
shower.
It was predicted that an electromagnetic shower generated by a high energy
primary could develop a charge excess which would emit Cherenkov radiation coherently
[1]. For ultra high energy primaries the Cherenkov radiation would be coherent in the
radio region of the spectrum; this long wavelength radiation might then be detected using
radio antennas [2, 3, 4]. Given the small predicted flux of ultra high energy neutrinos
[5, 6, 7], a suitable experiment to detect ultra high energy neutrinos using this method
requires a large, dense (and radio-transparent) volume. Antarctic ice is suitable for this
purpose. A detailed analysis of all aspects of such an experiment was done by Frichter,
2
Ralston and McKay (FRM) [8], using a simulation developed by Zas, Halzen and Stanev
(ZHS) [9]. FRM concluded that a modest array of optimally designed antennas could
detect many events per year. The sensitivity of radio detection peaks above 1 PeV, which
compliments an optical array such as AMANDA [10].
The Radio Ice Cherenkov Experiment (RICE) at the South pole [11] is designed as a
prototype detector of ultra high energy neutrinos with energy ≥PeV using this method.
One basic requirement for such an experiment is a reliable Monte Carlo simulation of the
shower development, Cherenkov radiation, detector, and data acquisition system. One
can also test the idea of coherent Cherenkov emission at accelerator facilities by dumping
a beam of photons or electrons into a dense target like sand or salt or any other suitable
medium. Such tests have begun with experiments at Argonne and SLAC [12]. A Monte
Carlo simulation which can be easily adapted to such a test beam experiment, where
Fresnel and possibly near zone radiation is important, and one that can include hadron
showers conveniently, is clearly desirable. The ZHS simulation, designed for electromag-
netic showers and Fraunhoffer (far zone) detection, has been a powerful tool. However
an expansion and update with extensive testing, offering applications to test beam and
neutrino astronomy, is currently needed.
We have written a GEANT-based Monte Carlo simulation to study coherent Cherenkov
emission in ice, salt, or a beam dump. GEANT 3.21 [13] is a well known and widely used
simulation and detection Monte Carlo package in particle physics1.
It allows access to
all details of the simulation such as controls of various processes, definition of target and
detector media, and a complete history of all events simulated. GEANT can be used to
simulate all dominant processes in 10 keV - 10 TeV energy range, although it has not been
extensively verified for energies >100 GeV, where the extrapolation of well-known lower-
energy electromagnetic cross-sections becomes large, and other effects (LPM, e.g. [14, 15])
become significant. For electron energies above 10 GeV, GEANT uses screened Bethe-
Heitler cross-section for bremsstrahlung together with the Migdal corrections [16, 17]. The
first Migdal correction is important for energy ≥1 TeV, reducing the cross-section. The
second correction reduces the differential cross-section for soft photon emission and is
effective even at much lower energies, in the 100 MeV - 1 GeV range. The LPM effect
in the context of UHE electromagnetic shower development and radio emission has been
discussed in [18, 19, 20].
GEANT is used to simulate electromagnetic showers inside materials from which we
extract detailed track information including particle type, coordinates, energy and inter-
action time. From this track information, we investigate shower properties like radiation
length, Moliere radius, critical energy and energy deposition in the material. We also de-
termine the resulting radio pulse using standard electrodynamic calculations from charged
particles' tracks and by parametrizing the shower. Ultimately we will consider hadronic
shower information and GEANT provides the flexibility to expand our analysis to this
case. We note that physics results presented thus far by RICE have neglected the hadronic
shower contribution and are, in this respect, conservative.
1Differences between GEANT 3.21 and GEANT 4 are primarily at energies below the threshold for
emission of Cherenkov radiation, and therefore do not affect the results presented here.
3
The organization of this report is as follows. In section two, we discuss various aspects
of electromagnetic showers and define quantities which characterize the shower. We present
results on the shower structure from the Monte Carlo simulation in section three and
compare them with established values in standard materials such as iron, lead and carbon.
Our analysis includes the detailed breakdown of the radiation-generating charge imbalance
into energy bins, the direct evaluation from energy considerations of Moliere radius and
the determination from dE/dx of radiation and energy deposition of the critical energy.
We discuss shower-to-shower fluctuations by parametrizing the showers in section four. In
section five, we review the theory behind coherent emission of an electromagnetic pulse
from the shower [21]. In section six, we calculate the electric pulse from the shower using
track information from GEANT. We summarize our results in section seven, making a
number of comparisons with ZHS, and discuss future work.
2 Shower Description
When a high energy electron or photon hits a material target, an electromagnetic shower
develops longitudinally inside the material. At the beginning of the shower, bremsstrahlung
and pair production are the dominant processes. Due to these processes, the number
of particles increases exponentially and the shower is created. Due to the dominance
of bremsstrahlung as the primary energy loss mechanism for high energy electrons, the
population of photons quickly dominates that of electrons or positrons. The energy of the
initial particle is divided among the secondaries. The exponential production of particles
is halted when the charged particles reach the critical energy (Ec). This is the transition
region where ionization loss overtakes radiation loss as the most important electron energy
loss mechanism. The particle population also reaches its maximum ("shower max") at this
point. After reaching the critical energy, charged particles lose their energy predominantly
inelastically by ionization, resulting in a subsequent decline in the number of particles in
the shower as they degrade in energy and range out in the medium.
Multiple Coulomb scattering is responsible for the transverse spread of the shower. The
shower core is populated by the highest energy particles. There is a long tail composed
mostly of the coulomb- scattered low-energy component.
Other processes - Compton, Moller and Bhabha and positron annihilation build up a
net charge (more electrons than positrons) in the shower as atomic electrons in the target
medium are swept up into the forward moving shower.
2.1 Energy Loss Mechanisms
Energy loss of an electron due to radiation is well approximated by the Bethe-Heitler
formula [22]
−(cid:18) dE
dx(cid:19)rad
= E
4NoZ 2r2
e
137A
ln(cid:18) 183
Z 1/3(cid:19)
(1)
where E is the initial energy of the electron, x is the distance in g-cm−2 units, No is
Avogadro's number, A is the mass number of the nuclei of the medium, Z is the atomic
4
number of the medium and r2
e = e4/m2c4 is the classical electron radius. For a composite
medium like ice (H2O), one has to calculate the effective Z and A using the proportion
by weight method2.
The Bethe-Bloch formula [23, 24] gives the energy loss due to ionization as
−(cid:18) dE
dx(cid:19)ion
= 4πNo
Z
A
r2
e mc2"ln 2mv2γ2
I
! − 1 −
δ
2#
(2)
where I = 10 Z eV and γ is the Lorentz factor in usual relativistic notation. At very high
energy, one needs to take into account the density effect as the medium becomes polarized.
This is taken care of by the following term inside the square bracket of Eq. (2):
where ωp is the plasma frequency of the medium.
δ/2 = ln(¯hωp/I) + lnβγ − 1/2,
(3)
A rough expression for the critical energy, Ec, can be found by the ratio of the expres-
sions for radiation and ionization losses. The log term in Eq. (1) is roughly 4 and the
square bracket term in equation (2) is approximately 11 if we include polarization effects.
The ratio is then
Ec ≈
605
Z
MeV.
(4)
An alternate definition of the critical energy is that of Rossi [25], according to which,
the critical energy (Ec) is the energy at which the ionization loss per radiation length is
equal to the electron energy. Critical energies determined from these two definitions are
compared in a subsequent section.
The radiation length (Xo) is given by
1
Xo
=
4NoZ 2r2
e
137A
ln (cid:18) 183
Z 1/3(cid:19)
(5)
so that the approximate expression for radiation loss (Eq.
E/Xo or
< E >= Eoe−x/Xo.
(1)) becomes −(dE/dx) =
(6)
Eq. (6) serves as a definition of the radiation length; Xo is the thickness of material needed
to reduce the mean energy of an electron to 1/e of its original through bremsstrahlung.
2.2 Longitudinal Profile
Heitler [26] developed a simplified model of an electromagnetic shower according to which
the initial electron, with energy Eo, radiates a photon, of energy Eo/2, in the 1st radiation
length.
In the 2nd radiation length, the photon creates an electron-positron pair and
the previous electron emits another photon; each particle now has energy Eo/4. This
process continues until the critical energy Ec, defined in the previous section, is reached,
at which point the shower is defined to be at maximum. At any radiation length, t, the
2see for example CONS110 section of [13].
5
number of particles (electron, positron and photon) is N (t) = 2t = exp (t ln 2) and the
energy per particle is E(t) = Eo/2t. Thus in this model the shower maximum occurs
at tmax = ln (Eo/Ec)/ln 2 and the total number of particles at the maximum is Nmax =
Eo/Ec. We can also calculate the integrated track length of the charged particles as
L = (2/3 ln 2)(Eo/Ec).
Although this model is over-simplified in assuming equality of all the primary cross-
sections and in assuming that the shower is artificially cut off, it nevertheless provides
a good description of the qualitative features of an electromagnetic cascade. Namely, it
predicts that the total track length of the charged particles scales linearly, while the posi-
tion of the shower maximum scales logarithmically with the initial energy of the shower.
These are important consistency checks for the electromagnetic showers generated from
any Monte Carlo.
An analytic method for realistic shower development was created by Carlson and Op-
penheimer [27] and later extended by Landau and Rumer [28]. Analytic shower equations
are solved in two approximations, namely approximation A and approximation B [25].
One neglects the Compton effect and ionization loss in approximation A. One includes a
constant energy loss per radiation length for all electrons and positrons in approximation
B, which is therefore expected to be a better model of the data. The derivation of the
integral spectra or number of particles (electrons and positrons) at different shower depth3
t in approximation A and B can be found in references [25, 29].
Greisen first parametrized the longitudinal profile of a photon-induced electromagnetic
air shower [30]. The Greisen parametrization (GP) is based on analytic shower theory in
approximation A. The difference between the more realistic approximation B and approx-
imation A is that of a slowly increasing function of the age parameter (s). The number of
particles in a given energy range increases when s < 1, reaches a maximum when s = 1
and then declines when s > 1.
Hillas [31] later modified Greisen's parametrization to fit Monte Carlo simulation of
0.1-1 TeV photon-induced showers. This modified GP was later used by Fenyves et. al.
[32] and is given for a single electron or photon of energy Eo as:
N (π)(Eo, E, t) =
0.31 A(E)
√y
exp[t1(1 − 1.5 ln s1)]
(7)
where t1 is the modified depth defined to be t1 = t + aπ,γ(E) and s1 is the modified age
parameter defined to be s1 = 3t1/(t1 + 2y). Superscript π denotes the total electrons and
positrons following the convention in the literature. The parameter y called lethargy is
defined to be y = ln(Eo/Ec), where Ec is the critical energy. One finds the parameters
A(E), aπ(E) for electron-induced and aγ(E) for photon-induced showers by fitting Monte
Carlo simulations.
2.3 Lateral Spread
The transverse development of the shower is well described by a quantity called Moliere
radius (RM). It is determined by the average angular deflection per radiation length at the
3Shower depth t = depth/Xo is always given in terms of the radiation length (Xo).
6
critical energy (Ec) due to multiple Coulomb scattering. The average deflection is given
by [33]:
E (cid:19)2
< δθ2 >= (cid:18) Es
δt
(8)
where Es = p4π/α mec2 = 21.21 MeV is the scale energy.
A numerical estimate of RM is obtained by considering the fraction of energy that
escapes transverse to the shower axis [34, 35]:
(9)
U (r)
E
= R ∞
0 R ∞
R ∞
0 R ∞
r E(r, t)drdt
0 E(r, t)drdt
where energy (E) is expressed as a function of shower depth (t) and radial distance r from
the shower axis. By definition, ninety percent of the shower energy is contained inside a
cylinder of radius RM centered on the shower axis. I.e., r = RM when U (r)/E = 0.1 in
Eq. (9). Moliere radius is independent of the energy of the shower and depends only on
the tracking medium in general.
In Rossi's definitions, the Moliere radius is related to the critical energy (Ec) and
radiation length (Xo) of the material [34, 35] through the equation
RM =
XoEs
Ec
,
(10)
which follows from Eq. (8).
3 Shower Simulations
The target medium in our GEANT simulations, unless otherwise stated, is defined to be
an ice cube of side 1 kilometer. Given the molecular composition, GEANT calculates the
effective atomic number, Z = 7.2 and an effective mass A = 14.3 for the compound ice.
Other parameters like radiation length, absorption length and cross-sections are calculated
automatically, once A and Z are specified.
GEANT gives all the details of particle tracking information like interaction points,
total energy, energy lost in interaction,
interaction time and so on. We used double
precision for all the variables in our output data files to minimize roundoff errors. Unless
stated otherwise, all particles are tracked down to total energy of 0.611 MeV, which is lower
than the energy at which particles are still relativistic and emit Cherenkov radiation.
3.1 Radiation Length
The radiation length given by Eq. (5) depends on the atomic and mass numbers of the
material. However employing the definition given by Eq. (6), we can fit an exponential to
the average energy of the injected electron at increasing depths. The average is taken from
the data of a number of showers generated by the Monte Carlo. Our ability to recover the
input value of Xo will serve as an internal consistency check to ensure that we are tracking
all the particles of interest, along with their energies.
7
)
V
e
G
(
y
g
r
e
n
E
1000
800
600
400
200
0
Fit
Data
0 20 40 60 80 100 120 140 160 180 200
Distance (cm)
Figure 1: Average energy of a 1 TeV electron injected in ice vs. distance in cm. Monte Carlo
data points are obtained from 500 electron tracks and the errorbars correspond to standard error
or s/√N , where s is the standard deviation and N is the number of tracks (500). The solid line
is the exponential fit as in Eq. (6) with Eo = 1 TeV. This fit gives a radiation length of 42.2 ± 4.3
cm. A 100 MeV energy threshold was used in all 500 simulations.
We generated 500 electron tracks each with 1 TeV primary energy (Ermo) incident
on the ice target and calculated energy loss due to bremsstrahlung. We tracked all the
electrons down to energy 100 MeV, well above the value of the critical energy so that
bremsstrahlung is dominant. We show the energy of the injected electron, averaged over
the 500 tracks, as a function of distance in Fig. 1. The errorbars correspond to standard
error or s/√N , where s is the standard deviation and N is the number of tracks (500).
A least squares fit of Eq. (6) to the Monte Carlo data keeping fixed Eo = 1 TeV yields a
radiation length of 42.2 ± 4.3 cm. The confidence level (CL) for the fit is 95.7%.
Given the molecular composition, GEANT also calculates the medium's radiation
length from the standard formula. For ice, GEANT calculates Xo = 38.8 cm, which
is roughly in agreement with the value we extract by tracking bremsstrahlung photons.
3.2 Moliere radius
To calculate the Moliere radius, we construct an imaginary cylinder centered on the shower
axis (up to the physical length of the shower). We add the energies, U , of all the tracks
that leave the cylinder without re-entering. By varying the radius of the cylinder we obtain
RM, which is the radius of the cylinder when the fraction U/Eo is equal to 0.1.
The Moliere radii (RM) for lead and iron are found to be 1.6 cm and 2.1 cm, respectively
(see Fig. 2). The value for lead is in agreement with experimental value of 1.6 cm [34].
We did not find the corresponding value for iron in the literature. A similar analysis for
ice results in Rice
M = 13 ± 1 cm. We quote the standard error, defined as s/√N , where s is
8
1
1
Eo = 1 GeV
Ice
Iron
Lead
Eo = 10 GeV
Ice
Iron
Lead
o
E
/
U
0.1
o
E
/
U
0.1
0.01
1
10
Radius (cm)
100
0.01
1
10
Radius (cm)
100
Figure 2: Moliere radius corresponds to the transverse development of the shower. When the
fraction U/Eo (the ratio of total energy inside an imaginary tube along the showers axis to the
initial energy of the shower) is 0.1 (the horizontal straight line), the corresponding radius is by
definition, the Moliere radius for that material. It depends on the material and not on the energy
of the shower. Here we compare showers of energies 1 GeV (left plot) and 10 GeV (right plot) for
each material; the Moliere radius is observed to be independent of energy in this range.
the standard deviation and N is the number of showers. It should be noted that the error
bars are correlated from bin to bin.
3.3 Energy Loss in the Medium
The signal from a Cherenkov type detector is proportional to the track length, which
is itself proportional to the energy deposition of the shower particles in the medium.
Therefore, we study the ionization loss per unit length in this section.
To determine dE/dx due to ionization from Monte Carlo, we generated 500 separate
5 GeV electron tracks in ice and kept a record of the rate at which energy was lost due
to ionization. The dots in Fig. 3 (left plot) shows dE/dx in ice for the 500 tracks at
different Monte Carlo steps. We also calculated dE/dx ionization loss in carbon, which
has roughly the same atomic weight as ice, as a consistency check. The result is shown in
Fig. 3 (right plot).
The average dE/dx curve from GEANT simulation matches the analytic approxima-
tion, Eq. (2) as shown in Fig. 4. The average ionization loss in ice (left plot) in the
relativistic region is approximately 2.4 MeVg−1cm−2 and the average minimum ioniza-
tion loss is approximately 1.9 MeVg−1cm−2. We have also calculated the average dE/dx
ionization loss in carbon (right plot) as a consistency check.
9
Figure 3: Electron stopping power or dE/dx ionization loss in ice (left plot) and in carbon (right
plot). The dots correspond to energy lost per unit length at each Monte Carlo step of 500 electron
tracks each with 5 GeV primary energy.
10
10
Ice
GEANT average
Bethe-Bloch
Carbon
GEANT average
Bethe-Bloch
)
2
^
m
c
/
/
g
V
e
M
(
n
o
i
t
i
a
z
n
o
i
)
x
d
E
d
(
/
)
2
^
m
c
/
/
g
V
e
M
(
n
o
i
t
i
a
z
n
o
i
)
x
d
E
d
(
/
1
1
10
Electron energy (MeV)
100
1000
1
1
10
Electron energy (MeV)
100
1000
Figure 4: Average electron stopping power or dE/dx ionization loss in ice (left plot) and in
carbon (right plot) calculated from GEANT data in Fig. 3. The error bars correspond to standard
error or s/√N , where s is the standard deviation and N is the number of tracks (500 in this case).
The solid lines are analytic curves calculated from Eq. (2). The agreement between Monte Carlo
and the analytic formula is reasonably good in the relativistic rise region.
10
3.4 Critical Energy
There are several methods to calculate the critical energy (Ec). The simplest is of course,
the rough estimate in Eq. (4). One can also obtain a value for Ec in the Rossi definition
from Eq. (10) after supplying the value of the Moliere radius (RM) and the radiation
length (Xo) found earlier. Yet another method is to find the energy at which the dE/dx
curves due to ionization and radiation losses intersect, which is the Monte Carlo equivalent
of the analytic expression in Eq. (4).
First, we find the critical energy from the Rossi definition using Eq. (10). The Moliere
radius for ice was found to be Rice
M = 13 ± 1 cm and the radiation length, Xo = 42.2 ± 4.3
cm. The critical energy is then calculated from Eq. (10) to be Eice
c = 68.8 ± 8.8 MeV.
Rossi [25] quotes 65 MeV for the critical energy of water, in agreement with our calculated
value. As a consistency check, we also calculated the critical energy for lead using the
Moliere radii found from GEANT simulation (section 3.2). The result is Elead
c = 7.4 MeV,
where we have used the radiation length of lead to be 0.56 cm. This is in agreement with
the experimental result [36].
Second, we calculate the energy loss due to radiation according to the formula in Eq.
(6) using the radiation length Xo = 42.2 ± 4.3 cm we found before. We then find the
energy at which this line crosses the ionization loss curve (see Fig. 5). The value for
critical energy (the energy at the crossing point of these two curves) found in this method
is 90 ± 9 MeV, where the error includes both the standard error in the dE/dx points and
the uncertainty in the radiation loss due to the error in Xo. The approximate formula
(Eq. (4)) gives Ec = 84 MeV, close to what we found from simulation. The critical energy
calculated in this way gives a higher value than Rossi definition, as expected.
3.5 Track Lengths
If we consider all processes to be elastic except ionization, then an estimate of the upper
bound for the total track length is given by,
L = Eo/(cid:18) dE
dx(cid:19)min
ion
(11)
where (dE/dx)min
ion is the minimum ionization energy loss per unit length. This value is
about 1.9 MeVg−1cm−2 for ice, (see Fig. 4) which yields a maximum total track length
of ∼ 1350 radiation lengths or 570 meters in ice for a 100 GeV shower. The actual total
track length is less than this value, since the energy loss due to ionization increases in the
relativistic region.
We plot different track lengths in Fig. 6 for a 100 GeV shower (averaged over 25
showers) versus the kinetic energy threshold used in the Monte Carlo to generate them.
The total track length for electrons and positrons together is denoted by total absolute
track length, the sum of electron and positron track-lengths projected along the shower
axis is denoted by total projected (e + p) track length and the difference between the
electron and positron track lengths projected along the shower axis is denoted by total
projected (e − p) track length. As we increase the threshold, so the simulation neglects
11
)
2
^
m
c
/
/
g
V
e
M
(
x
d
E
d
/
5.5
5
4.5
4
3.5
3
2.5
2
1.5
1
0.5
0
avg. ionization loss
Bethe-Bloch
avg. radiation loss
20
40
60
80
100 120 140 160 180 200
Energy (MeV)
Figure 5: Radiation and ionization energy loss vs. electron energy plot from GEANT Monte
Carlo simulation. The diagonal straight lines correspond to energy loss due to radiation with
radiation length Xo = 42.2 ± 4.3 cm, where the lowest curve corresponds to Xo = 42.2 + 4.3 cm
and the highest to Xo = 42.2 − 4.3 cm. The average ionization energy loss data are plotted as
obtained before (see Fig. 4). The error bars show the standard errors on the dE/dx points. The
crossing point of the two curves gives one definition of the critical energy.
track-lengths from low energy particles, we expect these total calculated track-lengths to
decrease.
We have also compared our results to those obtained using the ZHS Monte Carlo in
Fig. 6. Although the qualitative behavior of this scaling is the same for both Monte
Carlos, ZHS track lengths are consistently higher than GEANT track lengths by about
50%. In particular, GEANT produces an absolute track length of 400 meter for 0.1 MeV
kinetic energy threshold or 0.611 MeV total energy threshold. ZHS on the other hand
produces 650 meters of absolute track length for the same threshold, which is significantly
above our estimated maximum of 570 meters, based on the GEANT generated 100 GeV
data shown in Fig. 6 and application of Eq. (11).
The total track lengths increase with shower energy as more particles are created. The
track lengths are expected to increase linearly with energy for a given a energy threshold
used in the Monte Carlo. This scaling is shown in Fig. 7 below. Straight line fits to the
absolute track length and the projected (e − p) track length are also plotted. The slopes
of those straight lines are 3.2 and 0.5 respectively.
3.6 Particle yield
To generate a longitudinal profile of the shower, one counts the number of particles crossing
planes perpendicular to the shower axis inside the ice. We have calculated profiles for both
the total number of particles (e + p) and for the excess charge (e − p) in the shower.
12
1000
100
10
)
r
e
t
e
m
(
h
t
g
n
e
l
k
c
a
r
T
GEANT absolute
GEANT projected (e+p)
GEANT projected (e-p)
ZHS absolute
ZHS projected (e+p)
ZHS projected (e-p)
1
0.1
1
Kinetic energy threshold (MeV)
10
Figure 6: Total track lengths as functions of the kinetic energy threshold used in the Monte
Carlos. Here we show the total absolute track-lengths, the total (e + p) track-lengths projected
along the shower axis and the total (e − p) track-lengths projected along the shower axis. The
analysis is done for a 100 GeV shower (averaged over 25 showers) using the GEANT and ZHS
Monte Carlo simulations.
10000
)
r
e
t
e
m
(
h
t
g
n
e
l
k
c
a
r
T
1000
100
absolute
projected (e+p)
projected (e-p)
10
100
Shower Energy (GeV)
1000
Figure 7: Total absolute track length, total projected (e + p) track length and total projected
(e − p) track length as functions of shower energy from GEANT simulations. The linear scaling of
the track lengths with the shower energy is clearly observed. We used a kinetic energy threshold
of 1 MeV (see Fig. 6) and averaged over 25 showers in each case.
13
)
p
+
e
(
s
e
c
i
t
r
a
p
l
f
o
r
e
b
m
u
N
100
90
80
70
60
50
40
30
20
10
0
0
2
4
EGS4
GEANT
ZHS
8
6
14
Depth (radiation length)
10
12
16
18
20
Figure 8:
The longitudinal profiles of a 30 GeV shower (averaged over 50 showers) in iron
simulated by GEANT. The longitudinal profile is calculated by counting the number of electrons
and positrons crossing planes on the shower axis every half radiation length with total energy
greater than 1.5 MeV. We also plot the longitudinal profiles of the same shower from EGS4 (from
PDG 2000 [37]) and by ZHS Monte Carlos.
As a consistency check of GEANT, we have simulated 30 GeV electron-induced showers
in iron and compared the profiles with the modified Greisen parametrization, Eq. (7).
The longitudinal profiles were obtained by adding the number of particles with total
energy greater than 1.5 MeV, crossing planes spaced one-half radiation length apart, and
perpendicular to the shower axis (see Fig. 8). The number of particles agrees reasonably
well with EGS4, an electromagnetic shower code developed at SLAC which simulates the
same shower [37]. The total number of particles from ZHS simulation is also shown in the
plot.
We have fitted the GEANT generated electromagnetic showers in ice by the modified
GP in Eq. (7). We used the value for critical energy (Ec) in ice to be 67.7 MeV. The
fitting parameters and the confidence level (CL) of the fits for showers of energy 100 GeV
(averaged over 100 showers), 500 GeV (averaged over 50 showers) and 1 TeV (averaged
over 20 showers) each with 5 MeV threshold energy are given in Table 1 for both electron
and photon primaries. The fits to longitudinal profiles are plotted in Fig. 9.
14
)
p
+
e
(
f
o
r
e
b
m
u
N
1000
100
10
1
0.1
1 TeV
500 GeV
100 GeV
electron-induced
0
2
4
6
8
10
Depth (t)
12
14
16
18
20
)
p
+
e
(
f
o
r
e
b
m
u
N
1000
100
10
1
0.1
1 TeV
500 GeV
100 GeV
photon-induced
0
2
4
6
8
10
Depth (t)
12
14
16
18
20
Figure 9:
Longitudinal profiles of electron-induced (left) and photon-induced (right) showers
in ice of primary energy 1 TeV, 500 GeV and 100 GeV, averaged over 20, 50 and 100 showers
respectively. The threshold energy used is 5 MeV. The error-bars correspond to standard error.
The smooth curves are Greisen parametrization fits as given in Eq. (7). The fitting parameters
and the confidence levels (CL) for the fits are given in Table 1. The critical energy for ice used in
the fits is 68.8 MeV.
Table 1: Greisen parameters for GEANT
showers in ice with 5 MeV threshold and
Ec = 68.8 MeV.
Primary Eo (GeV) A(E)
0.50
0.50
0.50
0.52
0.51
0.52
aπ, γ CL(%)
0.33
0.26
0.76
0.99
1.14
1.01
100
500
1000
100
500
1000
γ
e−
96.3
81.9
71.8
95.6
95.8
88.4
The parameter A(E) gives over-all normalization for the number of particles. We see
from Table 1 that the number of particles at the shower maximum for a photon-induced
shower (A(E) = 0.50) is slightly less than that of an electron-induced shower (A(E) =
0.52). The other parameter aπ, γ takes care of the shift of the whole shower between a
photon-induced (aγ) and an electron-induced (aπ) shower. This shift is approximately 0.7
radiation length.
We have also fitted GEANT generated 100 GeV electron and photon induced showers
(each averaged over 50 showers) in ice with 0.611 MeV threshold energy by GP given in
Eq. (7). The fitting parameters and the CL for the fits are given in Table 2. The fits with
longitudinal profiles are plotted in Fig. 10. The critical energy in ice for the fits is 68.8
MeV as before.
15
120
100
80
60
40
20
)
p
+
e
(
f
o
r
e
b
m
u
N
0
0
2
4
6
8
electron induced
photon induced
10
12
Depth (t)
14
16
18
20
Figure 10: Greisen parametrization fits (smooth curves) to the longitudinal profiles of 100 GeV
electron and photon induced showers (each averaged over 50 showers) with 0.611 MeV threshold
energy from GEANT. The fitting parameters are given in Table 2. The critical energy for ice used
in the fits is 68.8 MeV.
Table 2: Greisen parameters for GEANT
showers in ice with 0.611 MeV threshold
and Ec = 68.8 MeV.
Primary Eo (GeV) A(E)
0.65
0.66
aπ, γ CL(%)
0.16
0.66
85.6
89.1
γ
e−
100
100
A three parameter fit to the GEANT generated 100 GeV electron-induced shower in
ice leaving the critical energy (Ec) as a free parameter along with A(E) and aπ gives
Ec = 58.3 ± 3.1 MeV. This is within error bars of the critical energy (Ec = 68.8 ± 8.8
MeV) found from Moliere radius calculation. The CL for this fit is 97.7%.
A comparison of averages over 50 showers of 100 GeV each from GEANT to the same
from ZHS Monte Carlo (see Fig. 11) shows about a 25 − 35% discrepancy for the total
number of particles at the shower max. The percentage discrepancy between the two
simulations remains the same at higher energies.
maximum, about twice Askaryan's original rough estimate [1].
The excess charge in the shower (N (e − p)/N (e + p)) is about 18% at the shower
Fig. 12 shows a longitudinal depth distribution of a much different kind. Unlike
particles crossing planes transverse to the shower axis, here we look at the distribution
of the complete count of excess electrons present in bins of one radiation length along the
shower axis. The figure shows the distribution of excess electrons (dN (e − p)/dt) in a
100 GeV shower (averaged over 50 showers) with 0.611 MeV total energy threshold. The
distribution is broken down into different energy bins, showing that half of the excess
16
180
160
140
120
100
80
60
40
20
l
s
e
c
i
t
r
a
p
f
o
r
e
b
m
u
N
0
0
2
4
6
8
GEANT (e+p)
ZHS (e+p)
GEANT (e-p)
ZHS (e-p)
10
12
Depth (t)
14
16
18
20
Figure 11: Comparison between shower profiles of a 100 GeV shower (averaged over 50 showers
in each case) from GEANT and from ZHS Monte Carlos. The error bars correspond to standard
error, or s/√N , where s is the standard deviation and N is the number of showers. The total
energy threshold used is 0.611 MeV in both cases. The difference in particle yield between the two
Monte Carlos is 25 − 35% and remains the same at higher energy.
electrons have energy 5 MeV or lower. This plot shows clearly the important point that
the bulk of the particles at shower maximum have low energy and contribute significantly
to the Cherenkov emission.
3.7
1-D Model Study
The comparisons in Figs. 8 & 11 prompt us to check the sensitivity of shower depth and
particle number at maximum to the cross-sections assumed in a simulation, since ZHS
and GEANT codes differ slightly in cross-sections for some processes (a few percent). To
investigate this question in a setting that is completely under our control, we have written a
one dimensional Monte Carlo that elaborates on Heitler's simple model to make it realistic.
We include Compton scattering, positron annihilation, ionization loss, Moeller scattering,
where interaction lengths are chosen from exponential distributions. The default values
for the cross-sections are taken from GEANT. We examine the effect of changing cross-
sections of various processes. The total number of particles (e + p and γ) from the 1-D
model scales linearly with shower energy and the position of the shower maximum scales
logarithmically with energy as expected.
The effect of changing individual cross-sections (increasing and decreasing by 25%) are
listed in Table 3 & 4. We also list the total number of electrons and positrons (N (e + p))
at the maximum depth (te+p
max) and the total number of photons (N (γ)) at the maximum
depth (tγ
max) for the default values of the cross-sections (denoted by σo).
17
5000
4500
4000
3500
3000
2500
2000
1500
1000
500
t
d
/
)
p
-
e
(
N
d
E > 50 MeV
5 < E < 50 MeV
E < 5 MeV
0
0
2
4
6
8
10
12
Depth (t)
14
16
18
20
Figure 12: Depth distribution of excess charge (dN (e − p)/dt) broken down into 3 different
energy bins: E < 5 MeV, 5 < E < 50 MeV and E > 50 MeV. This figure shows that half of the
excess electrons are created with energy below 5 MeV. We used a 100 GeV shower (averaged over
50 showers) with 0.611 MeV threshold to generate this plot.
Table 3: Results of increasing individual cross-section by 25% in the
1-D Monte Carlo. The default setting (GEANT) is denoted by σo.
The cross-sections are changed one at a time from their default values.
Parameter
te+p
max
N (e + p)
tγ
max
N (γ)
δ
5.6
269
6.4
1396
σo
6.9
266
6.9
1418
Brem. Pair Comp.
6.4
231
6.7
1209
6.9
281
6.4
1292
7.2
236
7.0
1570
6.7
265
6.4
1416
6.2
307
5.9
1495
dE/dx
Anni.
Table 4: Results of decreasing individual cross-section by 25% in the
1-D Monte Carlo. The default setting (GEANT) is denoted by σo.
The cross-sections are changed one at a time from their default values.
Parameter
te+p
max
N (e + p)
tγ
max
N (γ)
δ
7.4
298
6.9
1544
σo
6.9
266
6.9
1418
Brem. Pair Comp.
7.4
298
6.9
1544
7.4
249
7.4
1516
6.4
310
6.7
1173
6.1
271
6.4
1399
7.7
212
8.2
1259
dE/dx
Anni.
The individual cross-sections are changed one at a time from their default values. The
changes in the values of tmax and N are noticeable, but the percentage changes are much
smaller than the percentage changes in the cross-sections. For example, comparing the ef-
fect of increasing the bremsstrahlung cross section, listed in the column under "Brem." in
Table 3, the depth and number at maximum change by at most 4% and 15% respectively,
18
compared to the default values, listed under "σo, for a 25% change in the cross section.
It seems that the shower features are not especially sensitive to precise values of an in-
dividual cross section. We have not made an exhaustive study of the effects of changing
combinations of cross sections in all possible ways, but we have seen no indication that
effects would be large without pathological distortions of the cross-sections.
4 Shower Fluctuation
Any determination of shower energy is intrinsically uncertain because of the fluctuations
from shower to shower for the same injected energy. The fluctuations must be quantified
compactly so an efficient and realistic energy uncertainty from this source can be assigned
to detected showers. Fortunately, for UHE showers the huge numbers of particles and
fractionally smaller fluctuations make this less problematic than for small showers. We
summarize an approach to the fluctuation question in this section.
Conservation of total energy plays an important role in shower development. The shape
of a shower strongly depends on the position and energy of the first hard bremsstrahlung -
the later the first hard bremsstrahlung event, the deeper the shower maximum is. However,
for any shower, the primary energy (i.e., the total energy of the shower) qualitatively
dictates the energy loss profile and the particle yield with depth.
It is common practice to describe the mean longitudinal profile of an electromagnetic
shower by a Gamma distribution which is similar in shape to the Greisen parametrization.
The Gamma distribution normalized to unity is given by
f (t; a, b) = b
(bt)a−1e−bt
Γ(a)
.
(12)
However, Grindhammer et. al. [38] have shown that the fluctuations of the parameters
a and b from an average profile do not necessarily follow the individual shower fluctuations.
It is more reasonable to fit individual profiles by the Gamma distribution. One can then
fit a 2-dimensional Gaussian distribution to the parameter set {a, b} thus obtained and
study shower fluctuation by studying these parameters. The correlation between these
parameters a and b can be expressed as:
ρ = Covariance (a, b)/σaσb,
(13)
where σa and σb are standard deviations of the parameters a and b respectively and
Covariance (a, b) is the covariance matrix. The correlation ρ is roughly independent of the
energy of the shower.
4.1 Fit to Simulations
We have fitted the Gamma distribution Eq.
(12) to 50 GEANT generated individual
shower profiles in ice4. The energy of each shower is 100 GeV with 0.611 MeV threshold
energy. The parameters a and b are extracted from the Gamma function fits. Two such
fits to individual shower profiles are plotted in Fig. 13.
19
Figure 13:
Individual shower profiles of 100 GeV showers from GEANT with 0.611 MeV threshold
energy. The solid curves are Gamma function fits to the profiles. The parameters a and b in Eq.
(12) are obtained from the fits.
Figure 14: Scatter plot (left plot) of the parameters a and b from the Gamma distribution fit to
50 individual shower profiles each with energy 100 GeV and 0.611 MeV threshold. The contours
are from a 2-dimensional Gaussian distribution with mean (a, b) and (σa, σb) obtained from the
data set a, b generated as described in the text. Shower fluctuations (right plot) due to variation
of the parameters a and b within a standard deviation. The dark solid curve is the profile with
mean values of a and b. All particle numbers are normalized to 1.
20
Fig. 14a shows the scatter plot of the parameters a and b obtained by fitting 50
individual showers. We also show the contours of the 2-dimensional Gaussian distribution
fit to the set of parameters {a, b}. This plot is similar in shape to the one obtained by
Grindhammer using GEANT [38]. The correlation coefficient ρ = 0.72 we found is close
to 0.73 as found by Grindhammer et. al.
Fig. 14b shows shower-to-shower fluctuations as we vary the parameters a and b within
a standard deviation (σa and σb) of their mean values (< a > and < b >). The dark solid
curve corresponds to the shower with mean values of the parameters a and b. As can be
seen, the shower maximum can vary as much as ∼ 50%. As remarked above, the variations
become statistically less significant as the energy rises to the multi-PeV range, where UHE
neutrino detectors such as RICE are sensitive. The uncertainties from fluctuations, though
requiring study, are not expected to be a major part of the uncertainties in event locations
and energies in any case.
5 Electromagnetic Pulse Theory
Overview
Given a current density, one calculates the electric field by straightforward application
of Maxwell's equations. The most useful form of the electric field is the complex vector
~Eω(~x), depending on three-dimensional coordinate ~x and angular frequency ω. Before
getting into the details of the radiation field calculation, we review its general features
and several subtle points involving coherence, and near and far field limits.
The energy radiated per unit frequency interval per unit solid angle is proportional to
Eω2. In an asymptotically far field limit R → ∞, the radiation field Eω ∝ 1/R, where R is
the distance from the shower to the field point. Because the quantity Eω has dimensions
of mass in particle physics units and depends linearly on the track length L (reviewed
below), one expects the frequency dependence Eω ∝ Lω/R on dimensional grounds. This
linear dependence on ω breaks down at high frequencies, when wavelengths are smaller
than the typical dimensions of the shower and create a coherence cutoff. The purpose of
this section is to develop the concepts and formalism needed to address this and other
issues.
The asymptotically far field is not a feature of Cherenkov physics as commonly pre-
sented in texts or applied in particle physics detectors. In particle physics applications,
the track length is comparable to the distance to the observation point R. Textbook
treatments take the limit that the track length is infinite, [39, 40, 41] in which case the
electric field Eω ∼ √ω/√ρ, where ρ is the cylindrical radial coordinate. In this situation,
the radiation intensity is proportional to ω, a familar feature of laboratory Cherenkov ra-
diation. The cylindrical symmetry dictates this special frequency and radial dependence.
The transition from cylindrical configuration to spherical configuration is associated with
the terms "Fresnel and Fraunhoffer" zones. (Conditions of "far fields" R ≫ λ, ρ ≫ λ
are separate and assumed throughout; criteria separating the zones are given below.) The
4particle number (e + p) normalized to 1.
21
conditions of the RICE experiment are such that spherical symmetry is a good approxi-
mation for the most distant events, R ∼ 1 km, while Fresnel zone effects start to become
noticeable for the nearest likely events, R ∼ 100 m. Consequently both are reviewed.
There are two methods to calculate the electric field from the charged particles in a
shower. One method calculates the electric field from each charged track and adds them
by superposition. The other method parametrizes the shower's current and calculates elec-
tric fields analytically. Direct track-by-track calculation and superposition are presently
limited to the Fraunhoffer zone. The result is simple, and depends only on the angle to
the observation point. The analytic method works in either zone, and gives a compact
way to take a few parameters characterizing the shower and calculate the field. Of course
one must have a good parameterization for this to be a good approximation.
By using both the analytic and numerical methods, with their different strengths, we
are able to characterize the coherence structure of showers in considerably more detail
than previous work. The coherence extends to much higher frequencies than previously
thought. This is a very important point, confirmed by two independent methods, and
clears up a misconception that coherence should cut off at about 1 GHz.
5.1 Fraunhoffer Limit
The Fraunhoffer limit is appropriate to most of the RICE sensitive volume. It forms the
basis for our numerical, track-by-track computation of the field produced by a shower at
antenna sites remote from the shower.
In this subsection, we derive the expression for
the field produced by an individual track, which forms the basis for the calculation of the
full field calculated from a whole shower. In the next subsection, we outline the method
of parametrizing the shower as an effective current and calculating the field directly from
that current.
The power at time t, position ~x, radiated per unit solid angle by a moving electric
charge is given, in Gaussian units, by [39]
dP (t)
dΩ
=
c
4πR ~E(~x, t)ret2,
(14)
evaluated at the retarded time t = t′ + R(t′)/c.
(For now we calculate effects in vac-
uum; shortly we will supply the factors for the effects of a medium.) We use the Fourier
transformed variables
R ~Eω =
R ~E(t) =
1
√2π Z ∞
√2π Z ∞
1
−∞
−∞
R ~E(t)eiωtdt ;
R ~Eωe−iωtdω .
The energy radiated per unit frequency interval per solid angle is then given by
d2I
dωdΩ
=
c
4πREω2.
(15)
(16)
(17)
We will require the frequency dependence of the fields, so we work with the Fourier trans-
formed fields below.
22
Figure 15: Geometry for calculating electromagnetic fields from a single track segment. (r1, t1)
and (r2, t2) are the starting and ending position and time of the segment along which the particle
moves with velocity ~v.
The expression for the radiation field from a point source is conventionally defined as
the electric field term linear in the acceleration
~β:
~E(~x, t) =
q
c
~β}
(1 − ~β · n)3R
n × {(n − ~β) ×
ret
(18)
where ~β is the velocity of the particle, n is the direction of the observer and R is the
distance from the track to the observation point (see Fig. 15). The factor 1/R that
accompanies the ~β factor is the other trademark of the radiation field. As is the case for
the term that comes from the boosted Coulomb field, which has no explicit acceleration
dependence, Eq. (18) is singular at the Cherenkov angle in a medium with real index of
refraction greater than 1.
Combining Eq. (15) and Eq. (18) we have
~Eω(~x) = s q2
8π2c Z ∞
−∞
eiω(t′+R(t′)/c)
~β}
(1 − ~β · n)2R
n × {(n − ~β) ×
dt′.
(19)
At distances large compared to the range of motion of the source, n is approximately
constant and
R(t′) ≈ ~x − n · ~r(t′).
(20)
This is the Fraunhoffer approximation: the error in the phase ω~x − ~x′/c must be kept
small compared to 2π. Conditions for use of this approximation are discussed in the next
23
subsection5 (see Eq. (29)).
After integrating Eq. (19) by parts, and using the boundary conditions to set the end
point contributions to zero, one finds [39]
R ~Eω(~x) ≈ −iωs q2
8π2c
eiωR/cZ ∞
−∞
eiω(t−n·~r/c)[n × (n × ~β)]dt,
(21)
where R ≡ ~x. We will later apply this track-by-track expession to segments over which
~β is constant with time from t1 < t < t2. On each segment ~r = ~r1 + c~β(t − t1). We insert
the result into Eq. (21) and obtain
R ~Eω(~x) =
1
√2π (cid:18) q
c(cid:19) eikReiω(t1−n·~r1/c) ~β⊥
(eiωδt(1−n·~β) − 1)
1 − n · ~β
(22)
where n × (n × ~β) = −~β⊥, and δt = t2 − t1.
In a medium we replace c → c/n = c/√ǫµ, where n = n(ω) is the refractive index
of the material and ǫ = ǫ(ω), the dielectric constant. We also replace ~E → ~D = ǫ ~E and
~β → ~v n
c = n~β everywhere, to get
√2π (cid:18) µrq
R ~Eω(~x) =
1
c2 (cid:19) eikReiω(t1− n
c n·~r1)~v⊥
(eiωδt(1−n·~βn) − 1)
1 − n · ~βn
(23)
where µr is the relative permeability and k = nω/c.
The particle velocity ~v in the medium can be greater than that of light. The apparent
singularity in Eq. (23) at 1 − n · ~βn = 0 defines the Cherenkov angle θc as cos θc ≡ 1/nβ.
However there is no singularity, as seen by expanding the exponent. Close to the Cherenkov
angle, Eq. (23) reduces to the form
R ~Eω(~x) =
iω
√2π (cid:18) µrq
c2 (cid:19) eikReiω(t1− n
c n·~r1)~v⊥δt.
(24)
Eqs. (23 & 24), used by ZHS without detailed derivation [9], are coded into the simulation
to produce the field values for the tracks. Eq. (24) is explicitly linear in the track length
~vδt, a feature we mentioned earlier.
Eqs. (23 & 24) are incorporated into a track-by-track Monte Carlo simulation. The
code uses the exact formula Eq. (23) unless the conditions are very close to the Cherenkov
angle, posing a 0/0 numerical problem, in which case Eq. (24) is used.
Numerical summation of the electric field from all the tracks weighted by the proper
charge automatically incorporates the features of coherence. As illustrated shortly, coher-
ence produces a signal that peaks at the Cherenkov angle (θc), with a width away from
the Cherenkov angle that shrinks with increasing frequency [9]. The coherence features
are sufficiently intricate that we have devoted separate subsections to the topic.
5Analysis of the effects of keeping the next order in the expansion of the phase shows that significant
deviations from the Fraunhoffer result appear at distances where the Fresnel zone sets in [42], as discussed
from a general point of view below [43].
24
Figure 16:
Shower parametrization is done by assuming the longitudinal excess charge devel-
opment to be a Gaussian n(z ′) moving along the shower axis (z-direction) with speed v. n is the
direction of the observer, making an angle θ with the shower axis. ρ is the transverse distance to
the observer.
5.2 Parametrization Method
Charges of opposite sign which radiate coherently will give electric fields that cancel. It
is the excess charge (e − p) that determines the net field. Following Buniy and Ralston
(BR) [43], one can then parameterize the excess (net) charge development in a shower
and calculate the electric field due to the prescribed relativistic current. This is a flexible
and compact approach that takes a few essential parameters from the shower and allows
inspection of both the Fresnel and Fraunhoffer zones. We outline this method to keep our
presentation self-contained.
The vector potential ~A in the Lorentz gauge (adapted to ǫ(ω) and µ = 1) is given by
~Aω(~x) =
4π
c√2π Z d3x′ eik~x−~x′
~x − ~x′ Z dt′eiωt′ ~J(t′~x′)
(25)
The BR method parametrizes the excess charge development in the shower by a current
along the shower axis (z) as
~J(t′, ~x′) = ~v n(z′)f (z′ − vt′, ~ρ′),
(26)
where ρ = px2 + y2 is the radial distance from the shower axis, and n(z′) is the excess
charge (e − p) distribution, approximated by a Gaussian near the shower maximum (see
Fig. 16), as
n(z) =
e−z2/2a2
.
(27)
nmax√2π
This is analogous to Greisen and Rossi's parameterization [44], where nmax is the excess
charge at the shower maximum, except that the excess (not total) charge (e−p) is modeled
in the BR case. The parameter a corresponds to the "longitudinal spread" of the shower
25
near maximum and can be extracted by fitting a Gaussian to simulations of the excess
charge profile of a shower.
The region of the shower over which the fields arrive nearly in phase at a distance R(t)
is called the coherence zone [43], and is given by
∆zcoh = s R
k sin2 θ
,
(28)
where k = 2π/λ is the wave number and θ is the angle between R and the shower axis
(z). The value of the dimensionless ratio η,
η = (cid:18) a
∆zcoh(cid:19)2
=
ka2
R
sin2 θ,
(29)
determines how one should calculate the field. The ranges η < 1 and η > 1 correspond to
the Fraunhoffer and Fresnel zones, respectively. In the case of η ≥ 1, the series expansion
of the phase, Eq. (20) fails, and the exact phase must be kept. As noted earlier, both the
Fraunhoffer and Fresnel zones are far-field problems in the sense that kR ≫ 1 is assumed
in both cases.
For ω ∼ GHz the Fraunhoffer approximation6 is appropriate for R > 100 m distances
typical of the RICE experiment. However, for a test beam experiment, where the interplay
between different length scales is important, the Buniy-Ralston method to calculate the
field in the Fresnel zone is required.
5.3 Radiated Power and The Form Factor
We return to the ansatz Eq. (26), which describes a current with transverse distribution
independent of the longitudinal evolution. This is a reasonable assumption after the first
few radiation lengths of the shower development. We make the specific ansatz
~J(t′, ~x′) = ezq v n(z′) g(~ρ′) δ(z′ − vt′).
(30)
Note that the transverse extent of showers is small compared to the longitudinal extent.
The phase and the denominator in Eq. (25) can be safely expanded in cylindrical coordi-
nates as
~x − ~x′ = q(z − z′)2 + ρ2 − 2~ρ · ~ρ′ + ρ′2
≈ R(z, z′)(cid:18)1 −
ρρ′
R2!
R2 cos φ′(cid:19) + O R′2
and
6Shown in detail earlier (Eqs. (23 & 24)).
1
~x − ~x′ ≈
1
R
26
(31)
(32)
where R(z, z′) = p(z − z′)2 + ρ2 and R′ = pz′2 + ρ′2. This expansion of the transverse
variables avoids the Fraunhoffer approximation, and leads to a "factored" expression for
the field:
R ~Eω(~x) ≈ I F (ω).
(33)
where I is an integral characteristic of the longitudinal shower history, and F (ω) is called
the form factor. The integral I can be evaluated in the saddle-point approximation, as
detailed elsewhere [43], or evaluated by Monte Carlo. This feature of factorization to
express the field using a form factor works in both the Fraunhoffer and Fresnel zones.
In the Fraunhoffer zone, there is a further, and very remarkable, simplification. Keep-
ing the term linear in z′, and dropping terms of order R′2/R2 in the phase expansion, the
electric field at the Cherenkov angle (θc) is
R ~Eω(~x) = q 2√2π ieikR(cid:20)ez(cid:18) ω
c2 −
ǫv (cid:19) − eρ
×Z dz′n(z′)Z dρ′ρ′g(ρ′)Z dφ′e−i nω
k cos θc
k sin θc
ǫv
(cid:21)
(34)
c ρ′ sin θc cos φ′
It is remarkable that all terms in the phase exponent depending on z′ have vanished. This
means that radiation from the shower is coherent over the entire length of the shower
in the Fraunhoffer approximation:
the "coherence zone" is limited only by the track
length7. In obtaining these results, the Cherenkov condition cos θc = c/nv = 1/nβ and
the approximations: cos θc = z/R, sin θc = ρ/R (see Fig. 16) and ρ′/R ≪ 1 were used.
F (ω), which is given by
Inspecting the transverse integral, we see that it alone contributes to the form factor
⊥),
⊥g(~x′
F (ω) = Z d2x′
= Z ∞
= 2πZ ∞
⊥ e−i~k⊥·~x′
dρ′ρ′g(ρ′)Z 2π
dρ′ρ′g(ρ′) Jo(cid:18) nω
0
0
0
ρ′ sin θc(cid:19) .
c
dφ′ e−i nω
c ρ′ sin θc cos φ′
,
(35)
Here Jo is the Bessel function of order unity. Just as in particle physics usage, F (ω) is the
Fourier transform of the (transverse) excess charge distribution.
The preceding analysis leads to a convenient expression for the electric field at the
Cherenkov angle in the Fraunhoffer approximation, which apart from some normalization
factors is given by
R ~Eω(~x) = q 2√2π I(a)
ω sin θc
c2
F (ω).
(36)
The radiated power at the Cherenkov angle (θc) depends on the form factor only. As noted
earlier, the same form factor can be used in both the Fraunhoffer and Fresnel limits, so
that laboratory information about the form factor can be used directly. The form factor
is the central topic of subsequent sections discussing the coherence.
7The importance of the coherence zone is well known in the Landau-Pomeranchuk-Migdal [14, 15, 17]
effect, where the lengthening coherence zone suppresses bremsstrahlung. The Cherenkov condition causes
the coherence zone to expand to equal the entire track length.
27
6 Electric Field Pulse Calculation
In this section we carry out the computational outline just developed. The vector nature
of superposition is taken into account. The analytic construction of the form factor, de-
veloped earlier, is adapted to the numerical calculation, so that two independent methods
can be compared. We observe using the simulation data that coherence extends far above
the frequency regime anticipated from simple estimates using characteristic scales such as
the Moliere radius.
6.1 Vector Superposition
To calculate the electric pulse from a GEANT generated shower, we summed the con-
tributions to the electric field in Eqs.
(23 & 24) from all charged track segments us-
ing full 3-dimensional geometry. We denote the observation point by a unit vector,
n = (sin θ cos φ, sin θ sin φ, cos θ). The starting time t1 of each track segment is ob-
tained from GEANT output. The time interval δt for each track-segment is calculated for
a particle of total energy E (from GEANT output) traveling at constant speed β through-
out the track segment. Slowing of low energy particles is thus taken care of approximately.
The electrical field has an azimuthal symmetry which will be investigated later.
The electric field is a complex vector quantity. As such there are two questions of
coherence, one to do with phase and one to do with the vector nature of the field. The
vector nature of the field depends on the term n × (n × ~v) = −~v⊥ in Eqs. (23 & 24).
This construction picks out the component of the velocity which is perpendicular to the
direction of the observer. Since the track segments vary in direction, we add the electric
field contributions as vectors which allow any cancellation that may occur. The electric
field amplitude is then proportional to the track length transverse to the direction to
the observer (~v⊥δt). On the Cherenkov cone, one then expects that the field generated
by the entire shower is proportional to the track length projected along the shower axis
times the sine of the Cherenkov angle. The projected track length thus accounts for the
vector nature of the electric field for tracks that go in different directions, and serves as
an important diagnostic of the whole procedure.
However, the projected track length says nothing in itself about the conditions for
phase coherence, which is the topic of the next subsection.
6.2 Phase Coherence
The numerically generated phases of the fields, track-by-track, are determined by the
complex exponentials in Eq. (23). We call the phase angles ω(t1 − n
c n · ~r1) and ωδt(1 −
n · ~βn) the translational phase (TP) and the Cherenkov phase (CP) respectively. The
translational phase is kinematic, a consequence of translational invariance, and depends
on the beginning of each track segment. The Cherenkov phase vanishes at the Cherenkov
angle, which a point of stationary phase and dominates the emission. Two or more track
segments will contribute in phase if the observer lies on the Cherenkov cone, and the
beginnings of the tracks do not destructively interfere.
28
Cherenkov Phase (CP)
Angle = Cherenkov
500 MHz
1 GHz
5 GHz
10 GHz
Translational Phase (TP)
Angle = Cherenkov
500 MHz
1 GHz
5 GHz
10 GHz
1
0.1
0.01
n
o
i
t
u
b
i
r
t
s
D
i
1
0.1
n
o
i
t
u
b
i
r
t
s
D
i
0.01
0.001
0.0001
0.001
-4
-3
-2
-1
0
1
Phase Angle (Radian)
2
3
4
1e-05
-4
-3
-2
-1
0
1
2
3
4
Phase Angle in Radian
Figure 17: Distributions of the translational phase (TP) and the Cherenkov phase (CP) at the
Cherenkov angle (θ = θc) for a single shower of energy 100 GeV and threshold of 0.611 MeV.
The TP (left plot) shows strong coherence (sharp peak) at low frequencies. A slight positive
enhancement of the TP can be interpreted as particle positions slightly lagging the light cone
because β < 1. The CP (right plot), in comparison, remains strongly peaked at all frequencies,
which underscores the argument that the phase (CP) varies a little over the whole track.
We studied the distribution of the TP and CP at the Cherenkov angle (θ ≈ 55.8◦) and
at an angle (θ = 40◦) off the Cherenkov angle ( Figs. 17 & 18). We used a single 100 GeV
shower with 0.611 MeV total energy threshold to make the phase distribution plots. We
studied the distribution at frequencies: 10 GHz, 5 GHz, 1 GHz and 500 MHz.
The distribution of the TP (Fig. 17) shows a strong peak at low frequency, indicating
coherent phase emission from track segments at the Cherenkov angle. At high frequencies,
the distribution tends to become flat, with random phases indicating a loss of coherence.
A net positive phase indicates the particle positions slightly lagging behind the light cone
due to their speed β slightly less than 1. The flat distribution of the TP in Fig. 18, on
the other hand, clearly indicates random phases coming from track segments at angles off
the Cherenkov cone.
The distribution of the CP remains qualitatively the same both at the Cherenkov
angle (Fig. 17) and at angles off the Cherenkov angle (Fig. 18). It is also apparent that
frequency dependence of the CP is very weak.
The preceding discussion makes clear that phase coherence is dominated by the TP.
Finally, the distribution of TP and CP variables are substantially uncorrelated (see Figs.
19 & 20). These features indicate a valid factorization of the electric field, leading to an
independent motivation for the Ansatz Eq. (30) and consequent appearance of the form
factor. We discuss this factorization in the following subsection.
29
0.1
Translational Phase (TP)
Angle = 40 degrees
500 MHz
1 GHz
5 GHz
10 GHz
n
o
i
t
u
b
i
r
t
s
D
i
0.01
-4
-3
-2
Cherenkov Phase (CP)
Angle = 40 degrees
500 MHz
1 GHz
5 GHz
10 GHz
1
0.1
n
o
i
t
u
b
i
r
t
s
D
i
0.01
0.001
0.0001
-1
0
1
Phase Angle (Radian)
2
3
4
1e-05
-4
-3
-2
-1
0
1
2
3
4
Phase Angle (Radian)
Figure 18: Distributions of the TP and the CP at an angle (θ = 40◦) off the Cherenkov cone for
a single shower of energy 100 GeV with total energy threshold of 0.611 MeV. The flat distributions
of the TP (left plot) at all frequencies indicates the randomness of phases coming from randomly
located track segments. The CP distribution (right plot), is qualitatively the same as the case on
Cherenkov angle. Taken together, these features indicate that phase coherence is dominated by
the TP.
Figure 19: Scatter plots of the translational phase (TP) and the Cherenkov phase (CP) at the
Cherenkov angles (θc) and at frequencies: 1 GHz (left plot) and 5 GHz (right plot). A correlation
between the two variables would appear as a slanted line, or similar feature. The plots demonstrate
that the translational and Cherenkov phases are substantially uncorrelated.
30
Figure 20: Same as Fig. 19 at angle (θ = 40◦).
6.3 Factorization of the Electric Field, and the Discrete Form Factor
The electric field from all the shower particles can be factorized because the TP and CP
phases are uncorrelated, as shown above. We write the field in factored form as
Etot
ω ∝ Xj
eiφTP
j (eiφCP
j − 1) ∼ Xj
eiφTP
j Xj
(eiφCP
j − 1) ,
(37)
where the sums are over all track segments. The factorization takes a rather simple form
at the Cherenkov angle. The electric field close to the Cherenkov angle in Eq. (24) can
be written approximately for a single track as
R ~Eω(~x) =
iω
√2π (cid:18) q
c2(cid:19) ~v⊥δt eikR eiωz1( 1
v − n
c cos θ) e−i nω
c n·~r1⊥.
(38)
The approximation t1 ≃ z1/v has been assumed in writing Eq. (38), which is well sup-
ported by the TP phase plot at the Cherenkov angle shown above. The phase angle
ωz1(1/v− n cos θ/c) vanishes for the Cherenkov condition (cos θc = c/nv) and we have full
coherence along the z-direction (shower axis) as observed earlier.
We now write the factorization in Eq. (37) for the total electric field at the Cherenkov
angle from all the charged particles in the shower as
R ~Etot
ω (~x) =
iω
√2π (cid:18) q
c2(cid:19) eikRXj
sj(~v⊥)jδtj e−i nω
c n·(~r1⊥)j
(39)
where sj = ±1 for positrons and electrons respectively. The total electric field in Eq. (39)
is thus proportional to the total track length (Pj sj(~v⊥)jδtj) transverse to the direction
of the observer at any frequency ω. This track length is the projected (e − p) track length
times sin θc, as described earlier.
The coherent electric field emission at different frequencies now depends on the Monte
Carlo "discrete form factor" F (ω)MC, given by
F (ω)MC = Xj
sje−i nω
c n·(~r1⊥)j
= Xj
31
sje−i nω
c xj
1 sin θc.
(40)
1
10000
1 TeV
500 GeV
100 GeV
N
/
F
0.1
N
/
F
x
q
e
r
f
1000
1 TeV
500 GeV
100 GeV
0.01
0
10000
20000
30000
Frequency (MHz)
40000
50000
100
0
10000
20000
30000
Frequency (MHz)
40000
50000
Figure 21: Absolute value of the form factor (F (ω)) (left plot) and the frequency spectrum
(ωF (ω)) (right plot) plots for 1 TeV, 500 GeV and 100 GeV showers. The form factor has been
calculated using Eq. (40), where we naively sum over all the shower particles along the z-axis
(shower axis). The curves are normalized by dividing by the particle count N .
Note that the observation point is in the x − z plane.
6.4 The Discrete Form Factor and the Frequency Spectrum
We calculated the discrete form factor in Eq.
(40) for 1 TeV, 500 GeV and 100 GeV
showers. A factor of ω times the form factor is proportional to the electric field amplitude.
First, we naively carried out the sum in Eq. (40) for all the shower particles along
the whole shower axis (z) as suggested by the Cherenkov condition in the Eq. (38). The
absolute value of the form factor and the frequency spectrum are plotted in Fig. 21. It
shows an extended region of coherence, with F ormf actor ∼ N/ω, where N is the number
of shower particles. This naive estimate is somewhat misleading because the Cherenkov
condition is not satisfied for the whole shower axis. The CP also becomes flat at very high
frequencies, as does the TP at the Cherenkov angle (see Fig. 17a). A further complication
is statistical fluctuations, which become large where the form factor is small.
For a better understanding of the form factor and the frequency spectrum, we now
turn to the analytic method.
6.5 The Analytic Form Factor and the Frequency Spectrum
The analytic form factor as defined in Eq. (35) is the Fourier transform of the snapshot
of the charge distribution. It is a good approximation to assume that most of the electric
field contribution comes from the particles at the shower maximum [43].
To calculate the analytic form factor, we then determined the transverse distribution
of the particles within half a radiation length on both sides of the shower maximum for a
32
0.3
0.25
0.2
0.15
0.1
0.05
/
p
d
N
d
0
0
2
4
Fit
Data
16
18
20
6
8
14
Radial distance (p) in cm
10
12
Figure 22: Radial distributions dN/dρ of the excess (e − p) electrons within half a radiation
length on both sides of the shower maximum for a 100 GeV shower (averaged over 50 showers).
The solid curve is a Pade(1,3) fit to the excess charge distribution.
100 GeV shower (averaged over 50 showers). We found both the distribution of the total
particles dN (e + p)/dρ and the distribution of the excess electrons dN (e − p)/dρ. Here,
ρ is the cylindrical radial distance from the shower axis. The excess charge distribution
(dN (e − p)/dρ) is plotted in Fig. 22.
We observed that the radial distribution falls approximately like 1/ρ2 for large ρ,
and peaks at a remarkably small value of 0.5 cm. One way (by no means complete) to
understand these results recalls the multiple scattering of an electron by a static Coulomb
field. For a screened Coulomb field, the modified Rutherford scattering cross section is
given by [39]:
dσs
dQ
βc !2
= 8π Ze2
Q
(Q2 + Q2
s)2
(41)
where Q is the momentum transfer and Z is the atomic number of the scatterer; Qs =
(Z 1/3/192)mc is the momentum transfer associated with the screening radius; m is the
mass of the electron. For elastic scattering in the ultra-relativistic limit, Q2 = 2p2(1 −
cos θ) ≈ p2θ2, where p is the electron momentum. The cross-section formula in Eq. (41)
now takes the form:
dσs
dθ
βc !2
= 8π Ze2
p2θ
(p2θ2 + Q2
s)2 .
(42)
The average deflection angle, < θ >≈ ρ/∆z. We can get a crude estimate of the
transverse distribution dN/dρ from Eq. (42) as:
dN
dρ
= 8π Ze2
βc !2 n∆z3
p2 !
ρ
(ρ2 + ρ2
o)2
(43)
33
where n is the number density related to N by the formula N = nσ∆z. The peak of this
distribution is given by
Qs∆z
p
ρo =
.
(44)
We evaluate this with the average energy of the particles taken to be approximately
equal to the critical energy (Ec) at the shower maximum. The peak of the distribution is
then ρo = 0.5 cm for ice (Z = 7.2, ∆z = 39 cm and p ∼ Ec = 70 MeV). The excellent
agreement may be fortuitous, but gives some confidence that the Monte Carlo excess
charge distribution, which also includes numerous atomic collision, pair production and
Compton scattering processes, has a simple physical origin.
We made a fit to the excess charge distribution with a Pade (1,3) approximation of the
form dN/dρ = f (ρ) = n(ρ − a)/(1 + bρ + cρ2 + dρ3) (see Fig. 22). The fitting parameters
are: n = 2.15, a = 0.01, b = 4.73, c = 6.96, d = 0.33. The fit is good up to ρ ∼ 20 cm.
The choice of the Pade fit was made to preserve the ρ1 geometric zero in dN/dρ at the
origin, and the 1/ρ2 asymptotic behavior.
Finally, we calculated the analytic form factor using Eq. (35) as:
F (ω) = 2πZ ∞
0
dρ f (ρ) Jo(cid:18) nω
c
ρ sin θc(cid:19) .
(45)
For the high frequency end it was necessary to use a convergence procedure to modulate
the Bessel transform. We compare the analytic prediction to the discrete form factor of Eq.
(40) at the shower maximum for 100 GeV, 500 GeV and 1 TeV showers. The comparison
is shown in Fig. 23. For reference we also plotted the analytic and the discrete frequency
spectra, which are ω times the absolute value of the form factor. The agreement between
the two methods is good up to 10 GHz for all energies and up to 50 GHz for 1 TeV.
6.6 Direct Calculation: Monte Carlo Field Spectrum
The frequency spectrum of the electric field calculated using Eqs.
(23 & 24) at the
Cherenkov angle and an angle off the Cherenkov angle are plotted in Fig. 24a. We
calculated the spectrum for 1 TeV, 500 GeV and 100 GeV showers (each averaged over
many showers). The electric field amplitude rises linearly at low frequency: this is the lin-
ear dependence on ω due to dimensional analysis. The coherent behavior at the Cherenkov
angle and the incoherent behavior at a widely separated angle (40◦) are clear. The fre-
quency spectrum of the electric field calculated from a 100 GeV GEANT generated shower
is compared to the same from the ZHS code in Fig. 24b.
From the previous analysis, the frequency dependence of the electric field amplitude is
expected to be ωF (ω). We compare ωF (ω) to Eω generated by the Monte Carlo in Fig.
25. The electric field amplitude at the Cherenkov angle is proportional to the projected
(e− p) track length × sin θc as shown in Eq. (39). We used the total projected (e− p) track
length value of 70 meter (as we found before) for a 100 GeV shower as a pre-factor to
normalize the analytic form factor. It is clear from the plot that the analytic form factor
and simple dimensional considerations explain the electric field amplitude rather well up
through 5-15 GHz range. The coherence persists up to frequencies as high as 50 GHz. We
34
1
0.1
N
/
F
0.01
0.001
Analytic
1 TeV
500 GeV
100 GeV
10000
N
/
F
x
q
e
r
f
1000
Analytic
1 TeV
500 GeV
100 GeV
0.0001
0
10000
20000
30000
Frequency (MHz)
40000
50000
100
0
10000
20000
30000
Frequency (MHz)
40000
50000
Figure 23: Absolute value of the form factor F (ω) (left) and the frequency spectrum ωF (ω)
(right) plotted versus frequency for 1 TeV, 500 GeV and 100 GeV showers. The discrete form
factor has been calculated using Eq. (40), where we sum over all the particles within a radiation
length at the shower maximum. The analytic form factor has been calculated from the Fourier
transform to a fit to the excess charge distribution (see Fig. 22). Both curves are normalized by
dividing by the particle count N Note that the analytic curve tends to be lower than the numerical
calculation at high frequency.
notice that the agreement between the direct Monte Carlo calculation of the frequency
spectrum and the analytic frequency spectrum gets better with increasing shower energy.
We believe that the analysis captures the important physics of the processes, and validates
the results of the Monte Carlo.
6.7 Related Issues
Electrons and positrons undergo Coulomb scattering while traversing through the medium.
The particle track is therefore not a straight line from the point of its creation to the point
where it falls below threshold. The track contains many kinks due to elastic Coulomb
scattering. GEANT gives a detailed output of particle tracks which contain these kinks as
stated earlier; we calculate the electric field from step-tracks. The Monte Carlo developed
by ZHS, on the other hand used the straight tracks from the start points to the end points,
with timing correction, to calculate field [9] to simplify their calculation.
We studied the effect of these kinks and also the effect of taking all tracks along the
shower axis (see Fig. 26). We found little difference between the cases when we included
(the usual case) and not included the kinks. This is a result of the extended coherence
zone along the shower axis at the Cherenkov angle. When we calculated the electric field
taking only the components parallel to the shower axis, we found that the field increases
almost linearly with frequency, as expected from a single charge.
For practical purposes, particles are removed from the Monte Carlo simulation once
35
0.016
0.014
0.012
0.01
0.008
0.006
0.004
0.002
V
e
G
0
0
1
/
)
z
H
M
V
u
(
/
o
E
/
E
x
R
100 GeV-C
500 GeV-C
1 TeV-C
100 GeV-40
500 GeV-40
1 TeV-40
0.02
0.015
0.01
0.005
)
z
H
M
V
u
(
/
E
x
R
GEANT
ZHS
Eo = 100 GeV
0
0
2000
4000
6000
Frequency (MHz)
8000
10000
0
0
1000 2000 3000 4000 5000 6000 7000 8000 9000 10000
Frequency (MHz)
Figure 24: Direct Monte Carlo calculation of the frequency spectrum of the electric field mag-
nitude for 1 TeV, 500 GeV and 100 GeV showers (left). The calculation is done using Eqs. (23
& 24) at the Cherenkov angle and at 40◦ angle. The coherent behavior at the Cherenkov angle
and the incoherent behavior at the 40◦ viewing angle are clear. The frequency spectrum at the
Cherenkov angle for a 100 GeV shower is also compared to the same calculated using ZHS code
(right). All thresholds are 0.611 MeV and the averages are done with 50 showers in case of 100
GeV, 20 showers in case of 500 GeV and 10 showers in case of 1 TeV.
V
e
G
0
0
1
/
/
)
z
H
M
V
u
(
o
E
/
E
x
R
0.03
0.025
0.02
0.015
0.01
0.005
0
0
Analytic
100 GeV
500 GeV
1 TeV
10000
20000
30000
Frequency (MHz)
40000
50000
Figure 25: Comparison between the frequency spectra of the electric field magnitude at the
Cherenkov angle from direct Monte Carlo calculation and from analytic calculation. The analytic
calculation is done using Eq. (40) with the total projected track length of 70 m for a 100 GeV
shower as found before.
36
)
z
H
M
V
u
(
/
E
x
R
0.03
0.025
0.02
0.015
0.01
0.005
0
parallel
with kinks
no kinks
2000
4000
6000
Frequency (MHz)
8000
10000
Figure 26:
Frequency spectrum of the electric field magnitude at the Cherenkov angle for a
100 GeV shower (averaged over 50 showers). Our study shows that the electric field calculated
using particle tracks with and without kinks make little difference. The electric field magnitude,
on the other hand increases almost linearly with frequency like a single charge when we took the
component of the tracks parallel to the shower axis only.
they fall below certain threshold energy (0.611 MeV in our case). While physical particles
do not suddenly stop moving, simulated particles in the Monte Carlo, have potential
emission of stopping radiation. Even though high energy particles are almost parallel to
the shower axis, the low energy particles close to threshold are isotropic. The stopping
radiation (which is forward) from all these isotropic particles in general could cause a
problem if they are aligned with the observer on the Cherenkov cone.
Our study shows (see Fig. 27) that there is indeed an isotropic component in the
shower below energy 0.85 MeV. However, they do not contribute to the radiation at the
Cherenkov angle (θc ≈ 55.8◦). Particles with energy 1 MeV, for example, Cherenkov
radiates at an angle 49◦.
6.8 Angular Pulse Distribution
The angular distribution of the electric field amplitude peaks at the Cherenkov angle
(θc ≈ 55.8◦). Fig. 28 shows the 1 GHz pulse from a 100 GeV shower (averaged over
50 showers) with 0.611 MeV total energy threshold at 4 different azimuthal angles φ =
90◦, 180◦, 270◦ and 360◦. The pulse shows very little dependence on φ which corresponds
to the approximately symmetric distribution of particles about the shower axis as stated
earlier. We have also plotted the electric field amplitude from the ZHS Monte Carlo for a
100 GeV shower (averaged over 50 showers) at 1 GHz frequency in Fig. 28 for comparison.
Fig. 29 shows the angular pulse distribution for a 100 GeV shower (averaged over
50 showers) at 1 GHz, 750 MHz, 500 MHz and 250 MHz frequencies. The Gaussian half
37
Figure 27: Plot of the cosine of the polar angle (θ) of each track-segment with respect to the
shower-axis versus its total energy. Although high energy particles are very much aligned with
the shower axis, there is an isotropic low energy component which moves with the shower. The
isotropic component of the shower does not contribute to the Cherenkov pulse, which validates the
removal of the particles when they fall below a certain low-energy threshold, say 0.611 MeV.
0.012
0.01
0.008
0.006
0.004
0.002
)
z
H
M
V
u
(
/
E
x
R
0
20
30
GEANT 90
GEANT 180
GEANT 270
GEANT 360
ZHS
40
50
Observation Angle (Degrees)
60
70
80
90
Figure 28: Angular pulse distribution of a 100 GeV shower (averaged over 50 showers) with 0.611
MeV total energy threshold from GEANT. The pulse is calculated at 1 GHz frequency and at 4
different azimuthal angles (φ). The Cherenkov peak at the observation angle θ = θc shows very
little dependence on φ. The same pulse from the ZHS Monte Carlo is also plotted for comparison.
38
0.008
0.007
0.006
0.005
0.004
0.003
0.002
0.001
)
z
H
M
V
u
(
/
E
x
R
0
20
30
1 GHz
750 MHz
500 MHz
250 MHz
40
50
Observation Angle (Degrees)
60
70
80
90
Figure 29: Angular pulse distribution of a 100 GeV shower (averaged over 50 showers) with
0.611 MeV total energy threshold from GEANT. The pulse is calculated at frequencies: 1 GHz,
750 MHz, 500 MHz and 250 MHz. The plot shows inverse scaling of Gaussian half width of the
Cherenkov pulse with frequency.
width of the pulses are 2.0◦, 3.7◦ and 7.3◦ at 1 GHz, 500 MHz and 250 MHz frequencies re-
spectively. This corresponds to approximate inverse scaling of pulse width with frequency
which is analogous to a single-slit diffraction pattern as pointed out in reference [9].
7 Summary of Results and Conclusions
We have analyzed 100 GeV - 1 TeV electromagnetic showers and the radio frequency radi-
ation they produce in great detail. These studies are a necessary ingredient for designing
an experiment for radio detection of UHE cosmic ray induced showers in radio- transpar-
ent media. In particular, experiments to detect radio emission from showers induced by
high energy cosmic ray neutrinos interacting in the surface of the moon [45] and in the
South Polar ice-cap [46] are underway and have reported preliminary results. Coherent
radio emission from electromagnetic showers has recently been demonstrated in the labo-
ratory [12]. The technique is gaining recognition as a powerful tool for particle detection.
The thorough dissection and understanding of all the intricacies of the showers and their
relationships to the final radio pulse produced is our goal achieved in this paper.
Energy information for each stage of every track is readily available from the GEANT
shower code. This information is essential for tracking the energy distribution in the
shower and identifying the sources of radio emission. Among our results, is the direct
determination of the radiation length in ice from an exponential fit to bremsstrahlung
radiation energy loss as a function of depth in Sec. 3.1. When this information is combined
with the direct extraction of the ionization loss, Sec. 3.3, we determined the critical energy
39
from the data shown in Fig. 5. The value obtained is nicely consistent with that found by
using the Moliere radius extracted from the data for the radial energy flow, Sec. 3.2, in
combination with the radiation length. The consistency between the critical energy values
indicates that our application of the code and analysis of the data gives a correct physical
picture of the interplay among the competing processes in the shower as it develops.
Because the coherent radio emission of interest depends upon the charge excess in the
shower, we need the energy profile of the contributions to the charge excess. This again
requires the GEANT track-by-track energy information, and we show the total charge
imbalance broken down into energy ranges in Fig. 12. A large fraction, more than 50%, of
the imbalance comes from the energy range below 5 MeV. Though the track lengths are
small, the large number of particles leads to a significant contribution to the total track
length of the shower.
The total and projected track lengths for both the total and excess charge populations
are determined and shown to be proportional to total shower energy in Sec. 3.5. The total
track length is comfortably below a rough estimate of the upper bound for the total track
length. Our detailed study of the longitudinal profiles in iron shows overall agreement but
differences in detail among the GEANT, the EGS4 and the ZHS simulations in Fig. 8.
The profiles in ice for 100 GeV, 500 GeV and 1 TeV electron and photon induced showers
in ice are well described by a modified Greisen parametrization with critical energy value
extracted from the simulation data, as described above, and two fit parameters. The
confidence levels of the fits are typically 80% - 90%, as summarized in Tables 1 & 2.
The GEANT profiles in ice are qualitatively similar to those from the ZHS code, but
lie typically 25% - 35% lower. To determine whether small differences in cross section
values used in the different simulations could account for this difference, we developed a
1-dimensional shower code with the full set of cross sections for the relevant processes. As
we report in Sec. 3.7, the depth at maximum and the number of particles at maximum are
rather insensitive to changes in the cross sections. We therefore believe that the differences
in cross-sections are not the source of profile and total track length differences between our
simulation and that of ZHS. Unfortunately, one important check that we have not been
able to make is the GEANT vs. ZHS ionization energy loss (dE/dx). The ZHS code does
not admit a readout of energy loss by shower particles. Therefore, we were unable to make
from ZHS code plots similar to Figs. 3 & 4 we made with GEANT. If the ionization loss
in ZHS code is much lower than in GEANT, it might account for the difference between
showers produced by the two Monte Carlos.
We develop the framework to calculate the electric field in the 100 MHz to multi
GHz frequency range in Sec. 5. The track-by-track field calculation, applicable in the
Fraunhoffer zone, and the calculation treating the shower as a continuous current density,
applicable in the Fraunhoffer and and Fresnel zones, complement each other. We present
both for this reason. The direct, numerical calculation of the total field from the vector
sum of the fields from individual tracks allows a detailed study of the dependence on to-
tal track length. Moreover, we elucidate the effects of random direction changes due to
collisions and the pattern of phase relationships among the contributions from all tracks.
Complementing these intensive numerical calculations, we present analytic work that em-
ploys the form factor of the effective current density to calculate the field in both Fresnel
40
and Fraunhoffer regimes. Given our realistic model for the current density, which we fit
to the transverse charge distribution from the simulation, we show the remarkable result
that the radiation from the shower is coherent over the whole shower at the Cherenkov
angle in the Fraunhoffer limit.
Our study of the track-by-track phase coherence in Sec. 6 reveals that the phase
associated with the initial time and position of a track (TP) and the phase associated
with the "diffractive emission" (CP) from the track as a whole are uncorrelated. This
result supports our model of the current, and consequently of the field, in Sec. 5. Figs.
17 & 18 show the strong phase peaking of the "diffractive" phase, called Cherenkov phase
(CP) in the discussion, on and off the Cherenkov angle as defined by the shower axis.
The phase associated with the initial coordinates of the track, called the translation phase
(TP), is coherent at the Cherenkov angle but completely random away from that angle.
We noticed some time ago [47] that the frequency spectrum of the electric field at
the Cherenkov angle, calculated directly from the track data and the Fraunhoffer zone
formulas developed in Sec. 5., flattens out at frequencies above 2 GHz, as shown in Fig.
24.
It is clear from the figure that the ZHS simulation shows the same behavior but
they did not examine the high frequency region further [9]. We employ the form factor,
which correctly accounts for the transverse spread of the shower and its effect on the
field to analyze the frequency spectrum. We evaluate its Fourier transform, F (ω), two
independent ways and find agreement. We then show that the observed spectrum of the
field is faithfully represented up to 5 GHz for 100 GeV showers and up to 15 GHz for 500
GeV and 1 TeV showers. Up to 15 GHz, we now have a clear picture of the behavior of
the frequency spectrum, which continues to show coherent behavior.
The electric field in the Fraunhoffer zone as a function of the observation angle from
the shower axis peaks at the Cherenkov angle. The width of the peak shrinks inversely
with frequency and the height (field strength) rises linearly with shower energy. These
features confirm the ZHS results and the off-Cherenkov angular dependence is also similar.
The height of the peak at the Cherenkov angle for a given energy and frequency is larger
in their case by about 25-35%. This is perhaps not a surprising difference between two
independent shower simulations and field calculations. We have performed an extensive
set of tests and cross-checks to validate our results.
7.1 Conclusions
Our study quantifies coherent Cherenkov radiation from high energy showers and shows
that coherence persists from 100 MHz to tens of GHz. The existence of Coherent Cherenkov
radiation goes back to Askaryan, and has been studied for decades, while the persistence
of coherent emission at the Cherenkov angle at multi-GHz frequencies is new. The persis-
tence of coherence is established by our new, highly detailed study of the actual shower
currents and corresponding phase distributions in CP and TP introduced in Sec. 6. The
recognition that the TP is coherent over the whole length of the shower at the Cherenkov
angle is a new insight into the connection between shower particles and the fields they
produce. We have analyzed the electromagnetic shower characteristics in great detail,
including the energy distributions and energy losses, so that a complete set of shower
41
parameters can be extracted from the simulation data to confirm and cross check the con-
sistency of our picture. A grand summary of pertinent parameters and comparisons with
the ZHS and other sources where available is presented in Table 5.
GEANT
Table 5: Comparisons between 100 GeV showers in ice from GEANT
and from ZHS Monte Carlos
Parameter
Total Energy Threshold (MeV)
Total absolute track length (meter)
Total projected (e + p) track length (meter)
Total projected (e − p) track length (meter)
Position of the shower max. (radiation length)
Number of particles (e + p) at shower max.
Excess electrons (e − p) at shower max.
Fractional charge excess at the shower max
Cherenkov peak at 1 GHz (Volts/MHz)
0.611
399 ± 5
374 ± 4
70 ± 8
6.5
111 ± 7
20 ± 2
∼ 18%
7.5 × 10−9
ZHS
0.611
642
519
131
7.0
155 ± 10
37 ± 3
∼ 24%
1.0 × 10−8
The simulation and field calculation developed are directly applicable to energies up
to 1 TeV. Higher energy showers can be "bootstrapped" by evolving the shower particles
into this regime with a corresponding multiplication in particle number and emitted field
intensity. We intend to parametrize and extrapolate the data to the multi-PeV data
needed for the UHE showers relevant to the RICE experiment. With the GEANT base, it
will also be interesting to investigate similar shower and field calculation for the hadronic
component of neutrino-induced showers.
Acknowledgement
Help and advice from Enrique Zas at various stages of this work were invaluable. Discus-
sions and working sessions with Jaime Alvarez-Muniz were essential to our understanding
of the workings of the ZHS code. We thank George Frichter and Florian Hardt for early
diagnostic works using the ZHS code. Comments and questions from Frances Halzen on
this topic over many years have been most useful. Thanks to Tim Bolton for helping with
initial set up of the GEANT Monte Carlo code. This work is supported in part by the
NSF, the DOE, the University of Kansas General Research Fund, the RESEARCH COR-
PORATION and the facilities of the Kansas Institute for Theoretical and Computational
Science.
References
[1] G. A. Askar'yan. Sov. Phys., JETP14:441, 1962.
[2] M. A. Markov and I. M. Zheleznykh. Nucl. Inst. Meth., A248:242, 1986.
[3] A. L. Provorov and I. M. Zheleznykh. Astropart. Phys., 4:55, 1995.
42
[4] J. P. Ralston and D. W. McKay, in Arkansas Gamma Ray and Neutrino Workshop:
1989, edited by G. B. Yodh, D. C. Wold, and W. R. Kropp, Nuc. Phys. B (Proc.
Suppl.) 14A 356 (1990); reprinted in Proceedings of the Bartol Workshop on Cosmic
Rays and Astrophysics at the South Pole (AIP, NY 1990) Proceedings #198.
[5] S. W. Stecker and M. H. Solomon. Space Sci. Rev., 75:341, 1996.
[6] F. Halzen and E. Zas. astrophys. J., 488:669, 1997.
[7] J. Rachen, R. Protheroe, and K. Mannheim. astro-ph/99080301, 1999.
[8] G. Frichter, D. W. McKay, and J. P. Ralston. Phys. Rev., D53:1684, 1996.
[9] E. Zas, F. Halzen, and T. Stanev. Phys. Rev., D45:362, 1992.
[10] P. B. Price. Astropart. Phys., 5:43, 1996.
[11] RICE Collaboration. Proceedings of the ICRC 1997. astro-ph/9709223.
[12] D. Saltzberg et al. Phys. Rev. Lett., 86:2802, 2001.
[13] CERN Publication. GEANT-Detector Description and Simulation Tool, 1994.
[14] L. D. Landau and I. J. Pomeranchuk. Dokl. Akad. Nauk., 92:535, 1953.
[15] L. D. Landau and I. J. Pomeranchuk. Dokl. Akad. Nauk., 92:735, 1953.
[16] H. Messel and D. F. Crawford. Pergamon Press, Oxford, 1970.
[17] A. B. Migdal. Phys. Rev., 103:1811, 1956.
[18] A. Misaki. Phys. Rev., D40:3086, 1989.
[19] J. P. Ralston and D. W. McKay, in High Energy Gamma Ray Astronomy (Ann Arbor
1990), APS Conference Proceedings No. 220 (AIP, NY, 1991) edited by J. Matthews.
[20] J. Alvarez-Muniz and E. Zas. Phys. Lett., B411:218, 1997.
[21] for air shower applications of radio-detection, see: H. R. Allan, Progress in Elementary
Particles and Cosmic Ray Physics, North-Holland, 1971. J. L. Rosner and J. F.
Wilkerson, hep-ex/9702008. J. L. Rosner, hep-ex/9508011. J. L. Rosner, in First
International Workshop on Radio Detection of High Energy Particles, RADHEP 2000
(Los Angeles), AIP Conference Proceedings No. 579 (2001) edited by D. Saltzberg
and P. Gorham.
[22] H. Bethe and W. Heitler. Proc. Roy. Soc., A146:83, 1934.
[23] H. Bethe. Z. Phys., 5:325, 1930.
[24] F. Bloch. Ann. Phys., 16:285, 1933.
43
[25] B. Rossi. High Energy Particles. Prentice Hall, 1952.
[26] W. Heitler. The Quantum Theory of Radiation. Clarendon, 1954.
[27] J. F. Carlson and J. R. Oppenheimer. Phys. Rev., 51:220, 1937.
[28] L. D. Landau and G. Rumer. Proc. Roy. Soc., A166:213, 1938.
[29] T. K. Gaisser. Cosmic Rays and Particle Physics. Cambridge University Press, 1990.
[30] K. Greisen. Prog. Cosmic Ray Physics, 3:1, 1956.
[31] A. M. Hillas. J. Phys., G8:1461, 1982.
[32] E. Fenyves, S. Balog, N. Davis, D. Suson, and T. Stanev. Phys. Rev., D37:649, 1988.
[33] J. Nishimura. Handbuch der Physik, 1967.
[34] W. R. Nelson et al. Phys. Rev., 149:201, 1966.
[35] G. Bathow et al. Nucl. Phys., B20:592, 1970.
[36] O. I. Dovzhenko and A. A. Pomanskii. Sov. Phys. JETP, 18:187, 1964.
[37] Particle Data Group. The European Physical Journal, C15, 2000. The details of EGS4
code system can be found in the report by W.R. Nelson, H. Hirayama and D.W.O.
Rogers, SLAC-265, 1985.
[38] G. Grindhammer et al., at SSC Workshop on Calorimetry for the Supercollider
(Tuscaloosa, 1989). SLAC-PUB-5072.
[39] J. D. Jackson. Classical Electrodynamics. Wiley, 1975.
[40] Fracis E. Low. Classical Field Theory. John Wiley & Sons, Inc., 1997.
[41] L. D. Landau and E. M. Lifshitz. The Classical Theory of Fields. Pergamon Press,
1971.
[42] Florian Hardt. Unpublished, 1998.
[43] R. V. Buniy and J. P. Ralston. astro-ph/0003408, 2000.
[44] B. Rossi and K. Greisen. Revs. Mod. Phys., 13:240, 1941.
[45] P. W. Gorham et. al., in First International Workshop on Radio Detection of High
Energy Particles, RADHEP 2000 (Los Angeles), AIP Conference Proceedings No.
579 (2001) edited by D. Saltzberg and P. Gorham.
[46] D. Besson, in First International Workshop on Radio Detection of High Energy Par-
ticles, RADHEP 2000 (Los Angeles), AIP Conference Proceedings No. 579 (2001)
edited by D. Saltzberg and P. Gorham.
44
[47] S. Razzaque. Talk given at Phenomenology for the Nu Century, PHENO 2000 (Madi-
son) and at First International Workshop on Radio Detection of High Energy Parti-
cles, RADHEP 2000 (Los Angeles).
45
|
0811.0704 | 1 | 0811 | 2008-11-05T11:59:12 | Trigonometric Parallaxes of Massive Star Forming Regions: IV. G35.20-0.74 and G35.20-1.74 | [
"astro-ph"
] | We report trigonometric parallaxes for the high-mass star forming regions G35.20-0.74 and G35.20-1.74, corresponding to distances of 2.19 (+0.24 -0.20) kpc and 3.27 (+0.56 -0.42) kpc, respectively. The distances to both sources are close to their near kinematic distances and place them in the Carina-Sagittarius spiral arm. Combining the distances and proper motions with observed radial velocities gives the locations and full space motions of the star forming regions. Assuming a standard model of the Galaxy, G35.20-0.74 and G35.20-1.74 have peculiar motions of ~13 km/s and ~16 km/s counter to Galactic rotation and ~9 km/s toward the North Galactic Pole. | astro-ph | astro-ph |
Trigonometric Parallaxes of Massive Star Forming Regions: IV.
G35.20−0.74 and G35.20−1.74
B. Zhang1,5, X. W. Zheng1, M. J. Reid2, K. M. Menten3, Y. Xu3,6, L. Moscadelli4, A. Brunthaler3
Version: 2008 Oct 29
ABSTRACT
for
the high-mass star
We report trigonometric parallaxes
forming regions
G35.20−0.74 and G35.20−1.74, corresponding to distances of 2.19+0.24
−0.20 kpc and
3.27+0.56
−0.42 kpc, respectively. The distances to both sources are close to their near kine-
matic distances and place them in the Carina-Sagittarius spiral arm. Combining the
distances and proper motions with observed radial velocities gives the locations and full
space motions of the star forming regions. Assuming a standard model of the Galaxy,
G35.20−0.74 and G35.20−1.74 have peculiar motions of ≈ 13 km s−1 and ≈ 16 km s−1
counter to Galactic rotation and ≈ 9 km s−1 toward the North Galactic Pole.
Subject headings:
-- distances -- individual (G35.20−0.74, G35.20−1.74)
techniques: interferometric -- astrometry -- spiral arm: Sagittarius
1.
Introduction
G35.20−0.74 (IRAS 18556+0136) and G35.20−1.74 (IRAS 18592+0108) are high-mass star
forming regions. They both are home to strong methanol (CH3OH) masers at 12 GHz and were
selected as targets for our large project with the NRAO 1 Very Long Baseline Array (VLBA) to
study the spiral structure and kinematics of the Milky Way. Details of this project are discussed
in Reid et al. (2009), hereafter called Paper I.
1Department of Astronomy, Nanjing University, Nanjing 210093, China
2Harvard-Smithsonian Center for Astrophysics, 60 Garden Street, Cambridge, MA 02138, USA
3Max-Plank-Institut fur Radioastronomie, Auf dem Hugel 69, 53121 Bonn, Germany
4INAF, Osservatorio Astrofisico di Arcetri, Largo E. Fermi 5, 50125 Firenze, Italy
5Shanghai Astronomical Observatory, Chinese Academy of Sciences, Shanghai 200030, China
6Purple Mountain Observatory, Chinese Academy of Sciences, Nanjing 210008, China
1The National Radio Astronomy Observatory is a facility of the National Science Foundation operated under
cooperative agreement by Associated Universities, Inc
-- 2 --
The kinematic distance of G35.20−0.74 is about 3.3 kpc, based a CO line velocity of 35 km s−1
(Solomon et al. 1987). This star forming region is home to a massive protostar with a jet driven
outflow, rarely observed towards massive young stellar objects. Sub-arcsecond resolution VLA
observations of the region by Gibb et al. (2003) at 3.5 and 6.0 cm show that three concentrations of
radio continuum emission break up into 11 individual sources all lying along the outflow. As for most
high-mass star forming regions, G35.20−0.74 contains water and OH masers (Forster & Caswell
1989), methanol masers (Caswell et al. 1995), and many thermal lines from molecules like CH3CN
(Kalenskii et al. 2000), HCO+ and HCN (Gibb et al. 2003) have been detected. This region was
also observed recently at mid-infrared wavelengths by De Buizer et al. (2006). These observations
reveal an extended source with a cometary shape with the masers concentrated around the head.
The G35.20−1.74 star forming region also contains water, OH and methanol masers, and
a large cometary UCHII region. The UCHII region, known as W48A, lies about 20′′ from the
water and OH masers. Mid-infrared images (De Buizer et al. 2004) show a bright point source
at the location of the water masers as well as the UCHII region to the south-east. Kinematic
distances to G35.20−1.74, from various molecular lines associated with W48A (Genzel & Downes
1977; Palagi et al. 1993; Churwell et al. 1990; Braz & Epchtein 1983; Vallee & MacLeod 1990),
range between 3.0 and 3.4 kpc.
Starting at the Sun, a ray toward Galactic longitude 35◦ first crosses the Carina -- Sagittarius
spiral arm, then the Crux -- Scutum arm, before recrossing the Carina -- Sagittarius arm and then
crossing the Perseus arm. Thus distances are crucial to establish to which arm G35.20−0.74 and
G35.20−1.74 belong. Given the large kinematic anomalies measured for some massive star forming
regions, for example for W3OH (Xu et al. 2006a; Hachisuka et al. 2006), it is important to measure
distances to these and other sources by direct methods such as trigonometric parallax.
2. Observations and Calibration Procedures
On 2005 October 30, 2006 April 7 and October 7, and 2007 April 16, we observed methanol
masers at 12 GHz toward G35.20−0.74 and G35.20−1.74 with 8-hour tracks on the VLBA. Since,
for these sources, the declination parallax signature is considerably smaller than for right ascension,
we scheduled the observations so as to maximize the right ascension parallax offsets. We observed
several extragalactic radio sources as background references which provide independent checks
for parallax solutions. Table 1 lists the positions, intensities, source separations, observed radial
velocities and beam sizes.
-- 3 --
Table 1: Positions and Brightnesses
Source
R.A. (J2000)
Dec. (J2000)
Tb
(h m s)
′
(◦
′′)
G35.20−0.74...
J1855+0215...
J1855+0251...
J1907+0127...
18 58 13.0517 +01 40 35.674
18 55 00.1130 +02 15 41.100
18 55 35.4365 +02 51 19.562
19 07 11.9963 +01 27 08.963
G35.20−1.74...
J1903+0145...
J1904+0110...
J1907+0127...
J1855+0251...
19 01 45.5364 +01 13 32.545
19 03 53.0632 +01 45 26.306
19 04 26.3978 +01 10 36.696
19 07 11.9963 +01 27 08.963
18 55 35.4365 +02 51 19.562
(Jy/b)
0.7 − 0.9
0.03
0.20
0.10
3.5 − 9.5
0.18
0.33
0.30
0.50
θsep P.A.
(◦)
(◦)
...
...
1.0 −54
1.4 −29
2.3 +96
...
...
0.8 +45
0.7 +94
1.4 +80
2.2 −43
VLSR
(km s−1)
Beam
(mas)
31
...
...
...
43
...
...
...
...
1.2
...
...
...
1.2
...
...
...
...
Note. -- The fourth and fifth columns give the peak brightnesses (Tb) at 12 GHz and their separations
(θsep) and position angles (P.A.) east of north between maser and reference sources. The sixth column gives
the Local Standard of Rest (LSR) velocities of the masers. The last column gives the FWHM size of the
Gaussian restoring beam. Calibrators J1855+0251 (Fomalont et al. 2003) and J1907+0127 (Petrov et al.
2005) are from the VLBA Calibrator Survey, the other calibrators are from Xu et al. (2006b).
Our general observing setups and calibration procedures are described in Paper I, and here
we discuss only aspects of the observations that are specific to G35.20−0.74 and G35.20−1.74. We
observed two ICRF sources, J1800+3848 and J1800+7828 (Ma et al. 1998), near the beginning,
middle and end of the observations in order to monitor delay and electronic phase differences
among the observing bands. The right and left-circularly polarized emission from the 12 GHz
methanol masers toward G35.20−0.74 and G35.20−1.74 were observed in 4 MHz bands centered
at LSR velocities of 31 km s−1 and 43 km s−1, respectively.
The strongest maser spots at VLSR of 28.3 km s−1 and 41.5 km s−1 served as the phase-
references for G35.20−0.74 and G35.20−1.74, respectively. The point-source response functions
(dirty beams) had FWHMs of 1.9 by 0.8 mas at a PA of −8◦ east of north. After experimenting
with different restoring beams, we adopted a circular restoring beam with a 1.2 mas FWHM for
both maser sources and background sources. Figures 1 and 2 show radio continuum (from VLA
archive data) and velocity-integrated methanol maser emission from these star-forming regions.
-- 4 --
Fig. 1. -- Images of 4.9 GHz radio continuum emission associated with the star-forming region
G35.20−0.74, generated from archival VLA data (program AH241), and the velocity-integrated
maser emission (inset). The position of the methanol masers is designated with a plus (+) sign.
Contour levels are linearly spaced at 0.45 mJy for the continuum emission; they start at 0.06 Jy
beam−1 km s−1 and increase by factors of 2 for the maser emission.
-- 5 --
Fig. 2. -- Images of 8.4 GHz radio continuum emission associated with the star-forming region
G35.20−1.74, generated from archival VLA data (program AK477), and the velocity-integrated
maser emission (inset). The position of the methanol masers are shown as a plus (+) sign. Contour
levels start at 66 mJy beam−1 and increase by factors of 2 for the continuum emission; they start
at 0.27 Jy beam−1 km s−1 and increase by factors of 2 for the maser emission.
3. Parallax and Proper Motion Results
We fitted elliptical Gaussian brightness distributions to strong maser spots and the extragalac-
tic radio sources for all four epochs. The change in position of each maser spot relative to each
background radio source was modeled by the parallax sinusoid in both coordinates, determined by
one parameter (the parallax), and a secular proper motion in each coordinate.
3.1. G35.20−0.74
Fig. 3 shows maser reference channel images for G35.20−0.74 from the first and last epochs.
Imaging sources near zero declination is generally problematic for radio interferometers, and we
suspect that the low-level symmetric structures seen in the G35.20−0.74 images are caused by
-- 6 --
small calibration errors. However, since the parallax measurement comes almost exclusively from
the East-West data, this should not be significant problem. Keeping this in mind, one can see
that the emission appears dominated by a single compact component, and there is no significant
variation over the 1.5 year time span of our observations. One of the three background sources
(J1855+0215) proved to be very weak, and we used only J1855+0251 and J1907+0127 in the
parallax fits. Fig. 4 provides the images of background radio sources used for the fitting of parallax
and proper motion for G35.20−0.74.
Fig. 3. -- Images of reference maser source at VLSR = 28.3 km s−1 in G35.20−0.74 at the first and
last epoch. Observation dates are indicated in the upper left corner of each panel. The restoring
beam is in the lower left corner. Contour levels are spaced linearly at 90 mJy beam−1 .
-- 7 --
Fig. 4. -- Images of extragalactic radio sources used for the parallax measurements of G35.20−0.74
from the first epoch observation on 2005 Oct 30. Source names are in the upper left corner of each
panel. Contour levels are spaced linearly at 7.0 mJy beam−1 for J1855+0251 and 10.0 mJy beam−1
for J1907+0127.
-- 8 --
Fig. 5. -- Parallax and proper motion data and best-fitting models for G35.20−0.74. Filled and
opened markers are for maser spots at VLSR of 27.9 km s−1 and 27.5 km s−1, respectively. Plotted are
positions of both maser spots relative to the two extragalactic radio sources: J1855+0251 (triangles)
and J1907+0127 (squares). Left Panel: Positions with superposed curves representing the modeled
maser track on the sky with the first and last epochs labeled. The expected positions from the
parallax and proper motion fit are indicated (crosses). Middle Panel: Eastward (solid lines) and
northward (dashed lines) offsets and best-fitting models versus time. Data for the eastward and
northward positions are offset vertically and small time shifts have been added to the data for
clarity. Right Panel: Same as the middle panel, except the best-fitting proper motions have been
removed, allowing all data to be overlaid and the effects of only the parallax seen. The solid
(dashed) line shows the combined fitted eastward (northward) parallax curve.
-- 9 --
Table 2: G35.20−0.74 Parallax & Proper Motion Fits
Maser VLSR Background
Parallax
µx
µy
(km s−1)
27.9 ...
27.9 ...
27.5 ...
27.5 ...
27.9 ...
27.5 ...
Source
J1855+0251
J1907+0127
J1855+0251
J1907+0127
(mas)
(mas y−1)
(mas y−1)
0.411 ± 0.057 −0.07 ± 0.10 −4.50 ± 0.33
0.426 ± 0.037 −0.05 ± 0.06 −4.47 ± 0.10
0.506 ± 0.083 −0.37 ± 0.14 −2.83 ± 0.27
0.522 ± 0.094 −0.35 ± 0.16 −2.80 ± 0.25
combined
0.456 ± 0.045 −0.07 ± 0.08 −4.47 ± 0.16
−0.29 ± 0.08 −2.79 ± 0.16
Note. -- Combined fit used a single parallax parameter for both maser spots relative to the background
sources, but a different proper motion was allowed for the two maser spots to allow for internal maser
motions. Uncertainties are formal errors adjusted to give χ2 per degree of freedom of unity, except for the
"combined" parallax uncertainty which has been increased to allow for the difference between the parallaxes
of the two maser spots.
In order to measure the parallax and proper motion of G35.20−0.74, we fitted the two brightest
maser spots, at VLSR = 27.9 and 27.5 km s−1, and the two background radio sources, J1855+0251
and J1907+0127, for all four epochs. In Fig. 5, we plot the positions of these two maser spots relative
to the two background radio sources, with superposed curves representing the model maser tracks
across the sky. The fitting results, listed in Table 2, reveal that the individually measured parallaxes
and proper motions for two maser spots in G35.20−0.74 are somewhat different. However, the
results using the two background sources are reasonably consistent. This suggests that time variable
maser structure in one or both spots, and not atmospheric calibration, may limit the parallax
accuracy. Thus, we have increased the parallax uncertainty for the "combined" solution to half
the average difference between the fits for the two maser spots. The combined parallax of the two
astrometric maser spots is 0.456 ± 0.045 mas, corresponding to a distance of 2.19+0.24
−0.20 kpc.
The average proper motion of the two maser spots is −0.18 ± 0.06 mas y−1 toward the East
and −3.63 ± 0.11 mas y−1 toward the North. The difference in the northward proper motions of
the two maser spots of ±1.7 mas y−1 (±18 km s−1) is unusually large compared to other sources
given in Papers I -- V (Reid et al. 2009; Moscadelli et al. 2009; Xu et al. 2009; Zhang et al. 2009;
Brunthaler et al. 2009). While such an internal motion is not impossible, we suspect that the
difference may be a result of the poor north-south beam of the interferometer for this near zero
Declination source. At the distance implied by the parallax measurement, the average proper
motions correspond to −1.9 km s−1 and −37.7 km s−1 eastward and northward, respectively.
Completing the 3-D space velocity, the average LSR velocity of the spots is 27.7 km s−1, which
corresponds to a heliocentric radial velocity of 10.7 km s−1.
-- 10 --
3.2. G35.20−1.74
Figure 6 displays images of the maser reference channel for G35.20−1.74 from the first and
last epoch. The emission appears dominated by a single compact component, and there is no
significant variation over the 1.5 years spanned by our observations. In Fig. 7, we show images of
the four background radio sources at the first epoch. We fitted two-dimensional Gaussian brightness
distributions to the four background radio sources and the two brightest maser spots at VLSR =
41.9 km s−1 and 41.5 km s−1. In Fig. 8, we plot the positions of these maser spots relative to the
two background radio sources with superposed curves representing the model maser tracks across
the sky.
Our parallax estimates, given in Table 3, for three of the four background sources are consistent
for an individual maser spot. The parallax estimate using J1903+0145 differs from the others by
about 2-sigma, but gives the smallest post-fit residuals. Given the limited number of degrees of
freedom for the fits, it is unclear whether the data for this background source is worse than for the
other three background sources and has a fortuitously good fit, or if it is the best data and the
data for the other three sources agree fortuitously well. Rather than make this decision, we use the
data from all four background sources.
The difference in parallax between the two maser spots is roughly 0.10 mas. The scatter
among the 4 measures for each spot suggests an uncertainty of about 0.03 mas, the joint error in
the difference would be about 0.04 mas. So the discrepancy is just over 2-sigma. Changes in maser
structure might be a possible cause. The 41.9 and 41.5 km s−1 masers varied by 3 and 6 Jy over
the 4 epochs. However, with our limited data we cannot really evaluate which maser gives a better
parallax. So we have used both and conservatively expanded the parallax uncertainty to half the
difference between the individual values.
We adopt the "combined" parallax solution, which uses both maser spots and the four back-
ground sources, which gives a parallax for G35.20−1.74 of 0.306 ± 0.045 mas. As done for source
G35.20−0.74, we do not use the formal parallax uncertainty of ±0.015 mas, but instead conser-
vatively adopt a parallax uncertainty of ±0.045 mas, which allows for the larger than expected
difference in parallax estimates from the two maser spots. The parallax for G35.20−1.74 corre-
sponds to a distance of 3.27+0.56
−0.42 kpc.
The proper motion estimates for the two maser spots and for all four background sources appear
consistent within their joint uncertainties and the average proper motion is −0.71 ± 0.05 mas y−1
toward the East and −3.61±0.17 mas y−1 toward the North. For the measured parallax, the proper
motion corresponds to −11.0 km s−1 and −56.0 km s−1 eastward and northward, respectively. The
average LSR velocity of the spots is 41.7 km s−1, which corresponds to a heliocentric radial velocity
of 24.8 km s−1.
-- 11 --
Fig. 6. -- Images of reference maser source at VLSR = 41.5 km s−1 in G35.20−1.74 at the first and
last epochs. The epochs are indicated in the upper left corner of each panel. The restoring beam is
in the lower left corner. Contour levels are spaced linearly at 1.2 and 1.5 Jy/beam, respectively.
-- 12 --
Fig.
7. -- Images of background continuum sources used for the parallax measurements of
G35.20−1.74 at the first epoch. Source names are in the upper left corner and restoring beams
are in the lower left corner of each panel. Contour levels are spaced linearly at increments of 8.5
mJy beam−1 for J1855+0251, 9.5 mJy beam−1 for J1903+0145, and 15 mJy beam−1 for J1904+0110
and J1907+0127.
-- 13 --
Fig. 8. -- Parallax and proper motion data and fits for G35.20−1.74. Filled and opened markers
are for maser spot at VLSR of 41.9 km s−1 and 41.5 km s−1, respectively. Plotted are position
measurements of both maser spots relative to the four background sources: J1855+0251 (circles),
J1903+0145 (triangles), J1904+0110 (squares), J1907+0127 (hexagons). Left Panel: Positions on
the sky with the first and last epochs labeled. Data for the maser spot are offset horizontally for
clarity. The expected positions from the parallax and proper motion fit are indicated (crosses).
Middle Panel: Eastward (solid lines) and northward (dashed lines) positions and best fitting models
versus time. Data for the eastward and northward positions are offset vertically and small time
shifts have been added to the data for clarity. Right Panel: Same as the middle panel, except the
best fit proper motions have been removed, allowing all data to be overlaid and the effects of only
the parallax seen. The solid (dashed) line shows the combined fitted eastward (northward) parallax
curve.
-- 14 --
Table 3: G35.20−1.74 Parallax & Proper Motion Fits
Maser VLSR Background
Parallax
µx
µy
(km s−1)
41.9 ...
41.9 ...
41.9 ...
41.9 ...
41.5 ...
41.5 ...
41.5 ...
41.5 ...
41.9 ...
41.5 ...
Note. -- see table 2 note.
Source
J1855+0251
J1903+0145
J1904+0110
J1907+0127
J1855+0251
J1903+0145
J1904+0110
J1907+0127
(mas)
(mas y−1)
(mas y−1)
0.272 ± 0.048 −0.76 ± 0.08 −3.83 ± 0.66
0.220 ± 0.021 −0.67 ± 0.04 −3.71 ± 0.44
0.287 ± 0.042 −0.76 ± 0.07 −3.66 ± 0.36
0.274 ± 0.067 −0.75 ± 0.12 −3.85 ± 0.58
0.368 ± 0.047 −0.78 ± 0.08 −3.62 ± 0.08
0.319 ± 0.021 −0.70 ± 0.03 −3.50 ± 0.21
0.385 ± 0.043 −0.78 ± 0.07 −3.44 ± 0.31
0.363 ± 0.066 −0.76 ± 0.12 −3.63 ± 0.11
Combined
0.306 ± 0.045 −0.74 ± 0.07 −3.73 ± 0.24
−0.69 ± 0.07 −3.50 ± 0.24
4. Galactic Locations & 3-dimensional Motions
In order to study the 3-dimensional motion of the maser sources in the Galaxy, we convert
their radial and proper motions from the equatorial heliocentric reference frame in which they are
measured into a Galactic reference frame. A convenient frame is one rotating with a constant speed
at the position of the maser source: ie, a "local standard of rest" for the location of maser. The
methods used to convert to this frame will be documented in Reid et al. (in preparation, Paper VI).
We use the IAU recommended values of R0 = 8.5 kpc and Θ0 = 220 km s−1, and the Hipparcos
solar motion values U = 10.0 ± 0.40, V = 5.2 ± 0.60, and W = 7.2 ± 0.40 km s−1(Dehnen & Binney
1998). For this Galactic model, G35.20−0.74 has a peculiar motion of −13 km s−1 in the direction of
Galactic rotation, 0 km s−1 toward the Galactic Center and −8 km s−1 toward the North Galactic
Pole. For G35.20−1.74, we find peculiar motion component of −16 km s−1 in the direction of
Galactic rotation, 1 km s−1 toward the Galactic Center and −9 km s−1 toward the North Galactic
Pole. Thus, both sources are rotating slower than the Galaxy spins, that is, slower than for a
circular orbit for a flat rotation curve for the Galaxy. Neither source has a significant peculiar
velocity component toward the Galactic center and both are moving toward the NGP at about
−8 km s−1.
Research on the structure of the galaxy in Nanjing University is supported by the National
Science Foundation of China (NSFC) under grants 1062130, 10673024, 10703010 and 10733030,
and NBRPC (973 Program) under grant 2007CB815403. Andreas Brunthaler was supported by
the DFG Priority Programme 1177.
-- 15 --
REFERENCES
Braz, M.A., Epchtein, N., 1983, A&AS, 54, 167
Brunthaler, A. et al. (2009), this volume, (Paper V)
Caswell, J.L., Haynes, R.F. 1983, Aust. J. Phys., 36, 417
Caswell, J. L., Vaile, R. A., Ellingsen, S. P., Whiteoak, J. B. & Norris, R. P., 1995, MNRAS, 272,
96
Churchwell, E., Walmsley, C.M., Cesaroni, R., 1990, 159, A&AS, 83, 119
De Buizer, J. M., Radomski, J. T., Telesco, C. M. & Pinal, R. K., 2004, A&AS, Vol. 36, 1619
De Buizer, J. M, James M. , 2006 ApJ, 642L, 57
Dehnen, W., & Binney, J. J., 1998, MNRAS, 298, 387
Fomalont, E., Petrov, L., McMillan, D. S., Gordon, D., & Ma. C., 2003, AJ, 126, 2562
Forster, J. R., Caswell, J. L., 1989, A&A, 213, 339
Genzel, R., Downes, D., 1977, A&AS, 30, 145
Gibb, A. G., Hoare, M. G., Little, L. T., Wright, M. C. H., 2003 MNRAS, 339, 1011
Hachisuka, K. et al. 2006, ApJ, 645, 337
Kalenskii, S. V., Promislov, V. G., Alakoz, A., Winnberg, A. V. & Johansson, L. E. B., 2000 A&A,
354, 1036
Palagi, F., Cesaroni, R., Comoretto, G., Felli, M., Natale, V., 1993, A&AS, 101, 153
Petrov, L., Kovalev, Yu. Y., Fomalont, E., & Gordon, D., 2005, AJ, 129, 1163
Ma C., Arias E. F., Eubanks T. M., et al., 1998, AJ, 116, 516
Moscadelli, L. et al. 2008, this volume, (Paper II)
Reid, M. J. et al. (2009) this volume, (Paper I)
Reid, M. J. (in preparation), Paper VI
Solomon, P. M., Rivolo, A. R., Barrett, J., Yahil, A., 1987 ApJ, 319, 730
Vallee, J.P., MacLeod, J.M., 1990, ApJ, 358, 183
Xu Y., Reid M. J., Zheng X. W. & Menten, K. M. 2006a, Science, 311, 54
-- 16 --
Xu Y., Reid M. J., Menten K. M., Zheng X. W., 2006, ApJS, 166, 526
Xu, Y. et al. (2009), this volume, (Paper III)
Zhang, B. et al. (2009), this volume, (Paper IV, this paper)
This preprint was prepared with the AAS LATEX macros v5.2.
|
astro-ph/0701829 | 1 | 0701 | 2007-01-29T16:08:30 | A large CO and HCN line survey of Luminous Infrared Galaxies | [
"astro-ph"
] | A large CO, HCN multi-transition survey of 30 Luminous Infrared
Galaxies ($\rm L_{IR}>10^{11} L_{\odot}$) is nearing completion with the James Clerk Maxwell Telescope (JCMT) on Mauna Kea (Hawaii), and the IRAM 30-meter telescope at Pico Veleta (Spain). The CO J=1--0, 2--1, 3--2, 4--3,6--5, $ ^{13}$CO J=2--1, HCN J=1--0, 3--2, 4--3 observations, resulting from $\sim 250$ hours of JCMT, $\sim 100$ hours of 30-m observing time and data from the literature constitute {\it the largest extragalactic molecular line survey to date}, and can be used to address a wide range of issues and eventually allow the construction of reliable Spectral Line Energy Distributions (SLEDs) for the molecular gas in local starbursts. First results suggest that: a) HCN and HCO$^+$ J=1--0 line luminosities can be poor mass estimators of dense molecular gas ($\rm n\geq 10^4 cm^{-3}$) unless their excitation is accounted for, b) CO cooling of such gas in ULIRGs may be comparable to that of the CII line at $\rm 158 \mu m$, and c) low excitation of the {\it global} molecular gas reservoir remains possible in such systems. In such cases the expected low CO $\rm J+1\to J$ line luminosities for $\rm J+1\geq 4$ can lead to a strong bias against their detection from ULIRGs at high redshifts. | astro-ph | astro-ph | Astrophysics and Space Science manuscript No.
(will be inserted by the editor)
A large CO and HCN line survey of Luminous Infrared Galaxies
Padelis P. Papadopoulos · Thomas R. Greve · Paul van der Werf · Stefanie Muehle ·
Kate Isaak · Yu Gao
7
0
0
2
n
a
J
9
2
1
v
9
2
8
1
0
7
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
Received: date / Accepted: date
Abstract A large CO, HCN multi-transition survey of 30
Luminous Infrared Galaxies (LIR > 1011 L⊙) is nearing com-
pletion with the James Clerk Maxwell Telescope (JCMT)
on Mauna Kea (Hawaii), and the IRAM 30-meter telescope
at Pico Veleta (Spain). The CO J=1 -- 0, 2 -- 1, 3 -- 2, 4 -- 3,6 -- 5,
13CO J=2 -- 1, HCN J=1 -- 0, 3 -- 2, 4 -- 3 observations, resulting
from ∼ 250 hours of JCMT, ∼ 100 hours of 30-m observ-
ing time and data from the literature constitute the largest
extragalactic molecular line survey to date, and can be used
to address a wide range of issues and eventually allow the
construction of reliable Spectral Line Energy Distributions
(SLEDs) for the molecular gas in local starbursts. First re-
sults suggest that: a) HCN and HCO+ J=1 -- 0 line luminosi-
ties can be poor mass estimators of dense molecular gas
(n ≥ 104 cm−3) unless their excitation is accounted for, b)
CO cooling of such gas in ULIRGs may be comparable to
that of the CII line at 158 m m, and c) low excitation of the
global molecular gas reservoir remains possible in such sys-
tems. In such cases the expected low CO J+1 → J line lumi-
nosities for J + 1 ≥ 4 can lead to a strong bias against their
detection from ULIRGs at high redshifts.
Padelis P. Papadopoulos
Institut fur Astronomie, ETH Zurich, Zurich, Switzerland
Tel.: +41 (0)44 633 3826
Fax: +41 (0)44 633 1238
E-mail: [email protected]
Thomas R. Greve
California Institute of Technology, Pasadena, CA 91125, USA
Paul van der Werf
Sterrewacht Leiden, 2300 RA Leiden, The Netherlands
Stefanie Muehle
Department of Astronomy, Univ. of Toronto, 60 St. George Street,
Toronto, ON M5S 3H8, Canada
Kate Isaak
Physics and Astronomy, Univ. of Wales, Cardiff CF24 3YB, UK
Yu Gao
Purple Mountain Observatory, Chinese Academy of Sciences, Nanjing
210008, People's Republic of China
Keywords Galaxies: starbursts · Galaxies: ISM · Galaxies:
active · ISM: molecules
1 Introduction
The importance of Luminous Infrared Galaxies (hereafter
LIRGs) as the sites of the most extreme local star forma-
tion events ([9]) makes them the best available templates for
similar events in the distant Universe. Moreover their com-
pact CO-emitting regions ([1]) makes them ideal for multi-
line studies of their global molecular gas reservoir up to
very high frequencies (where the high-excitation molecular
lines lie) since the resulting narrow beams of today's sin-
gle dish radio telescopes can still measure their total line
fluxes with single pointings. Our sensitive CO J=2 -- 1, 3 -- 2,
4 -- 3, 6 -- 5, 13CO J=2 -- 1 and HCN J=3 -- 2 and 4 -- 3 line ob-
servations with the JCMT and the IRAM 30-m telescope
for ∼ 30 LIRGs, combined with available CO and HCN
J=1 -- 0 data from the literature, will allow for excellent new
constraints on the state of their global molecular gas reser-
voirs. More specifically the high-J CO and the three HCN
transitions can probe the excitation and mass of the dense
(n ≥ 104 cm−3) gas, considered as the immediate fuel of
their prodigious star formation ([10], [3]). The completion
of our CO J=4 -- 3 and J=6 -- 5 observations and follow-up ob-
servations of still higher rotational CO, 13CO and HCN tran-
sitions with Herschel will ultimately allow the deduction of
robust SLEDs for the star-forming molecular gas of LIRGs
in the local Universe. These can then be used to uncover any
universal aspects of the star-formation/molecular gas inter-
play in galaxies and find the best rest-frame mm/sub-mm
lines for detecting and imaging the numerous LIRGs that
will be found at high redshifts in future deep mm/sub-mm
continuum surveys ([5]).
2
Arp 220
CO J=3−2
5x(HCN J=4−3)
Arp 193
CO J=3−2
100x(HCN J=4−3)
Fig. 1 HCN J=4 -- 3 (scaled by the numbers in the upper right of each panel) and CO J=3 -- 2 spectra of four LIRGs. The HCN (4 -- 3)/(1 -- 0) brightness
temperature ratio r43(HCN) is found to vary from r43(HCN) ≤ 0.1 (Arp 193) to r43(HCN) ∼ 0.8 (Arp 220).
Table 1 HCN vs HCO+ line ratios in Arp 220 and NGC 6240
Galaxy
HCN (4−3)
(1−0)
HCN (3−2)
(1−0)
HCO+ (4−3)
(1−0)
HCO+ (3−2)
(1−0)
Arp 220
NGC 6240
0.8 ± 0.2
0.6 ± 0.2
1.0 ± 0.3
1.0 ± 0.3
0.33 ± 0.10
0.21 ± 0.06
0.27 ± 0.10
0.24 ± 0.08
2 HCN versus HCO+ lines as dense gas mass tracers
The rotational transitions of the HCN and HCO+ molecules
with their high dipole moments are currently the best (i.e.
most luminous) dense gas tracers in LIRGs. An HCN J=1 -- 0
line survey of such galaxies ([3]) has even revealed a poten-
tially constant star formation efficiency ((cid:181) LIR/Mdense(H2))
for molecular gas at densities n ≥ 104 cm−3, attributed to
such a gas phase being the direct fuel of star formation in
molecular clouds ([13]). Recent studies have inserted some
doubt as to whether HCN lines are good tracers of such a gas
phase and suggested those of HCO+ instead ([2]). Our sur-
vey reveals a large range of HCN line excitation in LIRGs
with otherwise similar IR, CO and HCN J=1 -- 0 luminosi-
ties (Figure 1). In the case of Arp 193 the observed very
low HCN (4 -- 3)/(1 -- 0) line ratio (its HCN 4 -- 3 is >100 times
weaker than its CO J=3 -- 2 line!) is compatible with the com-
plete absence of a massive and dense molecular gas phase.
On the other hand for the two archetypal ULIRGs Arp 220
and NGC 6240 the HCN ratios are significantly larger than
the corresponding ones (in rotational level) of HCO+ (Table
1). This, along with the fact that HCO+ rotational transi-
tions also have ∼ 5 − 7 times lower critical densities than
those of HCN, signifies that the latter is tracing a denser gas
phase and may thus may remain as the most suitable gas
mass tracer in galaxies once the excitation of its rotational
lines is properly accounted for.
3 The dense molecular gas in Mrk 231
The prototypical ULIRG/QSO Mrk 231 is the first object
in our sample for which the full set of line observations
Fig. 2 CO, CI, CII cooling contributions in the ULIRG/QSO Mrk 231
has been completed, allowing for a unique view into the
global excitation conditions of its molecular gas. Analysis
of the relative strengths of the CO J=1 -- 0, 2 -- 1, 3 -- 2, 4 -- 3,
6 -- 5 and HCN J=1 -- 0, 4 -- 3 lines (and 13CO J=2 -- 1 from the
literature,[4]) finds the low-J CO lines dominated by dif-
fuse (n ∼ few × 102 cm−3), and warm (Tk ∼ 90 − 100 K)
gas, while the high-excitation CO J=6 -- 5, 4 -- 3 and the HCN
transitions trace a denser (∼ (1 − 3) × 104 cm−3) phase with
somewhat lower temperatures (Tk ∼ 50 − 70 K). The latter
dominates the total molecular gas mass in Mrk 231, and its
total CO line cooling is similar to that of the CII line at
158 m m (Figure 2). This suggests a different thermal balance
to that of lower IR-luminosity galaxies, and may explain the
relative weakness of the CII cooling line in such systems as
a result of the dominance of dense Photon Dominated Re-
gions (PDRs) for most of their molecular gas ([6]).
3
Table 2 CO line ratios of a "warm" and a "cold" ULIRG
Galaxy
CO (2−1)
(1−0)
CO (3−2)
(1−0)
CO (4−3)
(1−0)
12CO
13CO
17208-0014
05189-2524
1.02 ± 0.22
0.32 ± 0.08
0.72 ± 0.17
0.24 ± 0.07 < 0.32
0.74 ± 0.23 ≥ 35
6 ± 2
4 A "cold" and a "warm" ULIRG
Molecular gas excitation at levels far below those found in
Mrk 231 or other such galaxies used as templates for high
redshift starbursts (e.g. Arp 220), have been uncovered by
our ongoing survey, further underlying the need for local CO
SLEDs to be established for a large sample of local LIRGs.
The two ULIRGs (LIR > 1012 L⊙) IRAS 17208-0014 and
IRAS 05189-2524 have been found to have very different
CO line excitation (Figure 4) despite similar IR continuum
and CO J=1 -- 0 line luminosities and similar S60 m m/S100 m m
ratios. For IRAS 17208-0014 the CO brightness temperature
ratios measured are typical for the high-excitation star form-
ing molecular gas found in many such galaxies, but those
found for the "cold" IRAS 05189-2524 are surprisingly typ-
ical of the quiescent molecular clouds found in the Milky
Way disk (Table 2). One-phase radiative transfer models us-
ing the Large Velocity Gradient (LVG) approximation and
constrained by the line ratios of IRAS 17208-0014 are com-
patible with densities of n ∼ 103 cm−3 and mostly warm
Tk ∼ (70 − 100)K. The high 12CO/13CO J=2 -- 1 ratio mea-
sured for this merger system (Table 2) is responsible for
the low CO J=1 -- 0 optical depths (t 10 ≤ 0.5) of this gas
phase, a result of its high temperatures and large average ve-
locity gradients dV/dR ∼ 30 km s−1 pc−1 (for an abundance
of [CO/H2] = 10−4). The latter corresponds to a parame-
ter Kvir = (dV/dR)obs
∼ 30 − 50, typical of an unbound gas
(dV/dR)virial
phase ([7]). It is also worth noting that the high (4 -- 3)/(1 -- 0)
CO ratio (comparable to that of CO(3 -- 2)/(1 -- 0)) cannot be
accounted by a single gas phase model and probably signi-
fies the emergence of a second warmer and denser phase. In
several LIRGs (e.g. Mrk 231) such a phase has been found
to dominate CO J + 1 → J, J + 1 ≥ 4 and HCN line emis-
sion and containing most of their molecular gas mass, and is
most likely the direct fuel of their prodigious star formation.
The CO emission in IRAS 05189-2524 on the other hand
is compatible with gas of n ∼ 100 cm−3, Tk ∼ 15 K, and
Kvir ∼ 1, typical of the quiescent and mostly self-gravitating
molecular clouds found in the Galactic disk immersed in a
FUV radiation field of G◦ ∼ 1 (Habing units). In such con-
ditions high-J CO lines are expected to be very faint with
CO(J+1 −J)/(1 −0) ≤ 0.01 for J+1 ≥ 4 for the bulk of the
molecular gas. This is much lower than that expected from
CO SLEDs of other ULIRGs such as Mrk 231 or Arp 220
or indeed most high redshift objects ([8]), with only the ex-
tremely red object HR 10 coming close (Figure 3).
Fig. 3 CO line luminosities of various high-redshift objects compared
to the Mrk 231 CO dense gas SLED template (bars). Luminosities are
normalized to those of Mrk 231 at either CO J=3 -- 2 or J=2 -- 1. Line
fluxes for the various objects are taken from the literature ([11], [8]).
3.1 A comparison with high-z starbursts
The high-J CO and the HCN lines detected in Mrk 231 pro-
vide good constraints on radiative transfer models which
then allow a good CO SLED to be constructed. Its compari-
son with objects at high redshifts shows that despite the high
excitation of its dense molecular gas its CO SLED stands
as rather average (Figure 3). However, strong lensing (af-
fecting several objects in Figure 3) can selectively amplify
the highly excited CO line emission of compact starburst re-
gions at the expence of a much more extended low-excitation
non-star forming molecular gas phase (e.g. [12]). Further-
more, the high-J CO observations usually conducted after
spectroscopic redshifts of high-z objects become available
may select mostly from a subset of starbursts where the (star-
forming)/(non star-forming) H2 gas mass ratio (expected to
vary strongly in the several merger systems among LIRGs)
is particularly large. The limited sensitivity of today's cm
and mm/sub-mm single dish telescopes and interferometers
will further accentuate the aforementioned biases, which may
thus produce CO SLEDs at high redshifts which are biased
towards the high excitation regime, and are thus unrepresen-
tative of the full excitation range in star-forming galaxies.
In that respect CO (and HCN) SLEDs of local ULIRGs,
established from observationally well-sampled J-ladders of
rotational transitions, are indispensible as benchmarks unaf-
fected by the aforementioned biases.
4
Fig. 4 CO J=4 -- 3, 3 -- 2 and 2 -- 1 spectral lines (plotted with declining line thickness) for the "cold" IRAS 05189-2524 (left panel) and the "warm"
IRAS 17208-0014 (right panel) galaxies.
Acknowledgements We would like to thank the telescope operators
and staff of the JCMT and IRAM 30-m telescope for their assistance
during the extensive observations that made this project possible. Spe-
cial thanks to Jim Hoge, Per Friberg and Iain Coulson at the JCMT for
expert advice and assistance with the observations during many years.
References
1. Downes D., & Solomon P. M. 1998, ApJ, 507, 615
2. Graci´a-Carpio J., Garci´a-Burillo S. & Colina L. 2006, ApJ, 640,
L135
3. Gao Y., & Solomon P. M. 2004, ApJ, 606, 271
4. Glenn J., & Hunter T. R. 2001, ApJS, 135, 177
5. Hughes D. H. in The Neutral ISM in Starburst Galaxies, ASP Con-
ference Series, Vol. 320, 2004, pg. 317
6. Kaufman M. J., Wolfire M. G., Hollenbach D. J., & Luhman M. J.
1999, ApJ, 527, 795
7. Papadopoulos P. P., & Seaquist E. R. 1999, ApJ, 516, 114
8. Riechers, D. A. et. al. 2006, ApJ, 650, 604
9. Sanders D. B., & Ishida C. M. 2004, in The Neutral ISM in Starburst
Galaxies, ASP Conference Series, Vol. 320, 2004, pg. 230
10. Solomon P. M., Downes D., & Radford S. J. E. 1992, ApJ, 387,
L55
11. Solomon P. M., & Vanden Bout P. A. 2005, ARA&A, 43, 677
12. Weiss, A., Walter, F., & Scoville, N. Z. 2004, in The Neutral ISM
in Starburst Galaxies ASP Conference Series Vol. 320, p. 142
13. Wu J., Evans II N. J., Gao Y., Solomon P. M., Shirley Y. L., &
Vanden Bout P. A. 2005, ApJ, 635, L173
Thus starburst galaxies at high resdshifts with similar
CO SLEDs to that of IRAS 05189-2524 would remain unde-
tected by typical high-J CO observations with today's radio
telescopes, even after scaling their CO line fluxes by their
∼ 10 times higher IR luminosities. One of the goals of the
CO and HCN multi-transition survey of LIRGs is to find
how often such low-excitation molecular SLEDs occur in
starbursts, and seek out correlations with other galaxy char-
acteristics (e.g. merger status, starburst age, dust tempera-
tures). This may in turn shed some light into how otherwise
vigorously star-forming galaxies can harbor large amounts
of low-excitation molecular gas.
5 Conclusions
We report preliminary results from our ongoing CO and HCN
multi-transition of 30 local LIRGs, the largest such survey
to date. These can be summarized as follows,
-- A large range of HCN line excitation precludes the sim-
ple use of the HCN or HCO+ J=1 -- 0 line luminosity as a
dense gas mass tracer.
-- In the two cases of Arp 220 and NGC 6240, the HCN
versus the HCO+ line excitation suggests the former trac-
ing the densest gas phase.
-- In the ULIRG/QSO Mrk 231 the luminous CO J=4 -- 3
and J=6 -- 5 line emission emanates from a different gas
phase than the one dominating the low-J CO transitions.
In that phase total CO line cooling is comparable to that
of the CII line at 158 m m. If confirmed for more such
galaxies this raises the possibility of a very different ther-
mal balance than that seen in lower IR-luminosity sys-
tems where the CII line is the dominant coolant.
-- The discovery of very low CO excitation in the ULIRG
IRAS 05189-2524 raises the possibility of a large exci-
tation bias against detecting similar objects at high red-
shifts via their high-J CO line emission.
|
0809.4012 | 1 | 0809 | 2008-09-24T16:30:05 | Properties of the Youngest Protostars in Perseus, Serpens, and Ophiuchus | [
"astro-ph"
] | We present an unbiased census of deeply embedded protostars in Perseus, Serpens, and Ophiuchus, assembled by combining large-scale 1.1 mm Bolocam continuum and Spitzer Legacy surveys. We identify protostellar candidates based on their mid-infrared properties, correlate their positions with 1.1 mm core positions, and construct well-sampled SEDs using our extensive wavelength coverage (lam=1.25-1100 micron). Source classification based on the bolometric temperature yields a total of 39 Class 0 and 89 Class I sources in the three cloud sample. We compare to protostellar evolutionary models using the bolometric temperature-luminosity diagram, finding a population of low luminosity Class I sources that are inconsistent with constant or monotonically decreasing mass accretion rates. This result argues strongly for episodic accretion during the Class I phase, with more than 50% of sources in a ``sub-Shu'' (dM/dt < 1e-6 Msun/yr) accretion state. Average spectra are compared to protostellar radiative transfer models, which match the observed spectra fairly well in Stage 0, but predict too much near-IR and too little mid-IR flux in Stage I. Finally, the relative number of Class 0 and Class I sources are used to estimate the lifetime of the Class 0 phase; the three cloud average yields a Class 0 lifetime of 1.7e5 yr, ruling out an extremely rapid early accretion phase. Correcting photometry for extinction results in a somewhat shorter lifetime (1.1e5 yr). In Ophiuchus, however, we find very few Class 0 sources (N(Class0)/N(ClassI)=0.1-0.2), similar to previous studies of that cloud. The observations suggest a consistent picture of nearly constant average accretion rate through the entire embedded phase, with accretion becoming episodic by at least the Class I stage, and possibly earlier. | astro-ph | astro-ph | Draft version June 12, 2018
Preprint typeset using LATEX style emulateapj v. 08/22/09
8
0
0
2
p
e
S
4
2
]
h
p
-
o
r
t
s
a
[
1
v
2
1
0
4
.
9
0
8
0
:
v
i
X
r
a
PROPERTIES OF THE YOUNGEST PROTOSTARS IN PERSEUS, SERPENS, AND OPHIUCHUS
Melissa L. Enoch1, Neal J. Evans II2, Anneila I. Sargent3, and Jason Glenn4
Draft version June 12, 2018
ABSTRACT
We present an unbiased census of deeply embedded protostars in Perseus, Serpens, and Ophi-
uchus, assembled by combining large-scale 1.1 mm Bolocam continuum and Spitzer Legacy surveys.
We identify protostellar candidates based on their mid-infrared properties, correlate their positions
with 1.1 mm core positions from Enoch et al. (2006), Young et al. (2006), and Enoch et al. (2007),
and construct well-sampled SEDs using our extensive wavelength coverage (λ = 1.25 − 1100 µm).
Source classification based on the bolometric temperature yields a total of 39 Class 0 and 89 Class I
sources in the three cloud sample. We compare to protostellar evolutionary models using the bolo-
metric temperature-luminosity diagram, finding a population of low luminosity Class I sources that
are inconsistent with constant or monotonically decreasing mass accretion rates. This result argues
strongly for episodic accretion during the Class I phase, with more than 50% of sources in a "sub-Shu"
(dM/dt < 10−6 M⊙ yr−1) accretion state. Average spectra are compared to protostellar radiative
transfer models, which match the observed spectra fairly well in Stage 0, but predict too much near-IR
and too little mid-IR flux in Stage I. Finally, the relative number of Class 0 and Class I sources are
used to estimate the lifetime of the Class 0 phase; the three cloud average yields a Class 0 lifetime
of 1.7 ± 0.3 × 105 yr, ruling out an extremely rapid early accretion phase. Correcting photometry
for extinction results in a somewhat shorter lifetime (1.1 × 105 yr). In Ophiuchus, however, we find
very few Class 0 sources (NClass 0/NClass I ∼ 0.1 − 0.2), similar to previous studies of that cloud. The
observations suggest a consistent picture of nearly constant average accretion rate through the entire
embedded phase, with accretion becoming episodic by at least the Class I stage, and possibly earlier.
Subject headings: stars: formation - ISM: clouds - ISM: individual (Perseus, Serpens, Ophiuchus)
– submillimeter – infrared: ISM
1. INTRODUCTION
The problem of how low mass stars like the sun form
has been studied extensively over the last few decades.
Compared to more evolved protostars and pre-main se-
quence objects, however, the earliest stages of the star
formation process, from the formation of dense cores
through the main mass accretion phase, are relatively
poorly understood. This lack of information is due in
large part to the difficulty of observing young, deeply
embedded sources, which are shrouded within dense pro-
tostellar envelopes and only observable via reprocessed
emission at mid-infrared to millimeter wavelengths. Fur-
thermore, most previous observations of deeply embed-
ded objects have naturally focused on a small number
of very bright or well known sources, due to the sensi-
tivity and resolution limitations of long-wavelength sur-
veys. Understanding the formation of typical stars re-
quires complete samples of young objects, over molecular
cloud scales.
Currently, the details of the early evolution of proto-
stellar sources are extremely uncertain, including mass
accretion rates during the Class 0 and Class I phases.
In addition, measurements of the timescales associated
Electronic address: MLE: [email protected]
1 Department of Astronomy, Univ. of California, Berkeley, CA,
94720
2 The University of Texas at Austin, Astronomy Department, 1
University Station C1400, Austin, TX, 78712-0259
3 Division of Physics, Mathematics & Astronomy, California In-
stitute of Technology, Pasadena, CA 91125
4 Center for Astrophysics and Space Astronomy, 389-UCB, Uni-
versity of Colorado, Boulder, CO 80309
with the earliest stages vary considerably, ranging from
105 to 107 yr for prestellar cores (Ward-Thompson et al.
2007) and from 104 to a few ×105 yr for Class 0
(Andr´e & Montmerle 1994; Visser et al. 2002). In fact,
the association of Class 0 and Class I with distinct
evolutionary stages is still a matter of debate (e.g.,
Jayawardhana et al. 2001).
Large surveys at mid-
infrared to millimeter wavelengths, where the SEDs of
embedded sources peak, are essential for understanding
how protostars evolve through their earliest stages. In
addition to providing complete samples of young sources,
large surveys are also important for characterizing vari-
ations in the star formation process with environment.
With a few notable recent exceptions (Jørgensen et al.
2007; Hatchell et al. 2007a), previous samples of very
young protostars have typically been compiled from
many different surveys, and suffered from systematics,
unquantified environmental effects, and small number
statistics. We recently completed large-scale 1.1 mm
continuum surveys of Perseus, Serpens, and Ophiuchus
with Bolocam at the Caltech Submillimeter Observatory
(CSO). Maps have a resolution of 31′′ and cover 7.5 deg2
in Perseus (140 pc2 at our adopted cloud distance of
d = 250 pc), 10.8 deg2 in Ophiuchus (50 pc2 at d =
125 pc), and 1.5 deg2 in Serpens (30 pc2 at d = 260 pc)
(Enoch et al. 2006; Young et al. 2006; Enoch et al. 2007,
hereafter Papers I, II, and III, respectively). These Bolo-
cam surveys complement large Spitzer Space Telescope
IRAC and MIPS maps of the same clouds from the
"From Molecular Cores to Planet-forming Disks" Spitzer
Legacy program ("Cores to Disks" or c2d; Evans et al.
2
2003).
Millimeter emission traces the dust in dense starless
cores and protostellar envelopes, and provides a measure
of core and envelope properties, including sizes, masses,
and spatial distribution. Spitzer IRAC and MIPS ob-
servations are complementary in that they provide in-
formation about the properties of any young protostars
embedded within dense cores. Combining these data en-
ables us to assemble a mass limited, unbiased census of
the youngest star-forming objects in three different envi-
ronments, including prestellar cores, Class 0, and Class I
protostars.
In Paper III we looked at how the global cloud envi-
ronment influences the properties of star-forming cores
(Enoch et al. 2007). In a companion paper to this work
(Enoch et al. 2008), we examine the properties of prestel-
lar and protostellar cores in Perseus, Serpens, and Ophi-
uchus, focusing on the prestellar core mass distribution
and the lifetime of prestellar cores. A similar analysis
comparing the c2d Spitzer data to large SCUBA maps
of Perseus and Ophiuchus has recently been carried out
by Jørgensen et al. (2007, 2008), focusing on the differ-
ence between starless and protostellar cores, as well as
cloud properties such as the star formation efficiency
and how it varies with spatial clustering. We follow
Di Francesco et al. (2007) in defining millimeter cores
containing a compact luminous internal source (i.e., an
embedded protostar) as "protostellar cores", regardless
of whether the final object will be stellar or sub-stellar
in nature. We use "starless cores" to refer to dense cores
without an internal luminosity source, and "prestellar
cores" as starless cores that are likely to be gravitation-
ally bound (see Enoch et al. 2008).
In this work we exploit the combined power of millime-
ter and mid- to far-infrared observations to study the
evolution of young protostars embedded within the pro-
tostellar cores. Extensive wavelength coverage from λ =
1.25− 1100 µm allows us to trace the evolution of proto-
stellar sources in their main mass accretion phase, from
formation through the end of the embedded phase. We
follow Robitaille et al. (2006) and Crapsi et al. (2008)
in using "Stage" (e.g., Stage 0, I, II, III) to refer to
a source's true physical nature, regardless of its ob-
served properties, while the corresponding "Class" refers
to the observational classification, typically based on
the near- to mid-infrared slope αIR (Adams et al. 1987;
Greene et al. 1994) or on the bolometric temperature
(Myers & Ladd 1993).
We assume that Stage 0 and Stage I refer to an evo-
lutionary sequence of embedded protostars with M∗ <
Menv and M∗ > Menv, respectively (e.g. Andr´e 1994),
where M∗ is the protostar mass and Menv the enve-
lope mass. Similarly, we assume Stage II refers to pre-
main sequence stars with very little remaining envelope
(Menv < 0.1M⊙; Crapsi et al. 2008). Our adopted def-
initions of various Classes and Stages are summarized
in Table 1.
Ideally, Class I would directly correspond
to Stage I, etc., but as shown by Robitaille et al. (2006)
and Crapsi et al. (2008), geometric effects can cause, for
example, Stage II sources to be classified as Class I.
In §2 we describe the 1.1 mm and Spitzer infrared (IR)
data, including how protostellar candidates are identi-
fied and their association with 1.1 mm cores. We cal-
culate bolometric luminosities and temperatures, enve-
Enoch et al.
lope masses, and discuss completeness (§3). Protostel-
lar classification methods are compared in §4. Spectral
characteristics of Class 0, I and II sources detected at
1.1 mm are discussed in §5, including selected individual
sources (§5.2). In §6 we explore alternative classification
schemes made possible by our large sample, and com-
pare the average observed spectra to protostellar models
in §7. Sources are placed on a bolometric temperature-
luminosity diagram for comparison to protostellar evolu-
tionary models and to study their luminosity evolution,
mass accretion rates, and envelope evolution (§8). Fi-
nally, in §9 we calculate lifetimes for the Class 0 phase.
2. COMBINING BOLOCAM AND SPITZER C2D DATA
Both Bolocam 1.1 mm and Spitzer maps were designed
to cover down to a visual extinction of AV & 2 mag in
Perseus, AV & 3 mag in Ophiuchus, and AV & 6 mag in
Serpens (Evans et al. 2003). The actual overlap in area
between Bolocam and IRAC maps is shown in Figure 1
of Papers I, II, and III for Perseus, Ophiuchus, and Ser-
pens, respectively. Catalogs listing c2d measured Spitzer
fluxes of all sources in each of the three clouds, as well as
near-infrared fluxes for sources in the 2MASS catalogs,
are available through the Spitzer database (Evans et al.
2007). We utilize wavelength coverage from λ = 1.25 to
1100 µm, using 2MASS (1.25, 1.65, 2.17 µm), IRAC (3.6,
4.5, 5.8, 8.0 µm), MIPS (24, 70, 160 µm), and Bolocam
(1.1 mm) data. Note that 160 µm flux measurements are
not included in the c2d delivery catalogs due to substan-
tial uncertainties and incompleteness, but are included
here when possible. Photometry at 160µm is discussed
in Rebull et al. (2007) and Harvey et al. (2007b), where
160µm fluxes for point sources in Perseus and Serpens
are also given.
Basic data papers describe the processing and analysis
of the Spitzer IRAC and MIPS maps of Perseus, Ser-
pens, and Ophiuchus, as well as presenting general prop-
erties of the sources in each cloud, such as color-color
and color-magnitude diagrams (Jørgensen et al. 2006;
Harvey et al. 2006; Rebull et al. 2007; Harvey et al.
2007b). The young stellar object (YSO) population in
Serpens is discussed in detail by Harvey et al. (2007a).
Here we are most interested in the young protostellar
sources that are most likely to be embedded in the mil-
limeter cores detected with Bolocam. For the following
we will use the term "embedded protostar candidate" in
general to encompass candidate Stage I and younger ob-
jects. Although it will become apparent in §2.1 that our
criteria also pick up a number of Stage II sources, in gen-
eral we focus on sources with evidence for a protostellar
envelope (Stage 0 and I).
Figure 1 shows the result of combining Spitzer and
Bolocam data, for a few examples of embedded protostel-
lar sources in each cloud. Images are three-color Spitzer
maps (8.0, 24, 70 µm), with 1.1 mm contours overlaid.
The Bolocam ID of the associated 1.1 mm core is given
in each panel, as well as the embedded source IDs from
Tables 2–4. We use the 2MASS, IRAC, MIPS, and Bolo-
cam data to construct complete SEDs from 2 to 1100 µm
for each candidate protostellar source, shown in the lower
panels (see also §3). Additional SHARC II 350 µm fluxes
(Wu et al. 2007) and SCUBA 850 µm fluxes (Kirk et al.
2006, open circles) are included when available. Modified
black-body (Sν ∝ νβBν (T )) curves for a temperature of
Protostars in Serpens, Perseus, and Ophiuchus
3
Fig. 1.- Three-color Spitzer images 8 µm (blue), 24 µm (green), and 70 µm (red) of selected embedded protostar candidates in Perseus,
Serpens, and Ophiuchus, with Bolocam 1.1 mm contours. The Bolocam ID of the centered core and embedded protostar IDs from Tables 2–4
given, and contour intervals for the 1.1 mm emission are 3, 6...15, 20, 25, ...σ. Spectral energy distributions (SEDs) of the candidates are
plotted in the lower panels. SEDs include 2MASS, IRAC, and MIPS photometry from the c2d database and 1.1 mm fluxes (asterisks).
Published 350 µm and 850 µm points (open circles) are also shown when available. A modified blackbody spectrum (T = 15 K, β = 1) is
overlaid for reference.
4
Enoch et al.
15 K and β = 1 are shown for reference in Figure 1 (dot-
ted lines).
Protostellar sources in our sample may be isolated
(e.g., Per-Bolo 57) or lie in crowded regions (e.g., Per-
Bolo 49). Approximately 20%–50% of the time, more
than one protostellar source lies within a single 1.1 mm
core (20% of the embedded protostar sample in Ophi-
uchus, 40% in Perseus, and 55% in Serpens). Much of
this difference is certainly due to the lower resolution of
Bolocam (31′′) compared to Spitzer (7′′at 24 µm), caus-
ing nearby Spitzer sources to be blended in the 1.1 mm
map, although the envelopes detected at 1.1 mm are also
physically more extended than the region emitting at
Spitzer wavelengths. Sometimes the SEDs of such mul-
tiple sources look similar to each other (Per-emb 18 and
22 in the lower-left panel of Figure 1), and sometimes
quite different (Per-emb 49 of the same panel).
Although the coverage of the IRAC, MIPS and Bolo-
cam maps overlaps nearly perfectly for our purposes,
there are a few cases in which embedded protostar candi-
dates are outside the boundaries of the MIPS 70 µm map
(three sources in Serpens) or 1.1 mm map (one source in
Perseus). In addition, a number of bright sources in each
cloud are saturated in the 24µm or 70µm c2d maps. In
these cases we substitute IRAS 25µm or 60µm fluxes is
the source does not appear to be blended in a visual in-
spection of the the IRAS maps.
The 160 µm maps are often saturated near bright
sources and in regions of bright extended emission, such
as near bright clusters of sources. Reliable 160 µm fluxes
are especially difficult to determine in crowded regions,
due both to the large beam size (40′′) and to saturation
issues. The lack of 160 µm data is most problematic in
Ophiuchus, where the 160 µm maps are saturated in all
of the dense source regions. Even for isolated sources,
the measured 160 µm flux density, determined from a
point spread function (PSF) fit, may be underestimated
if the source is extended. The effects of these issues on
our analysis are discussed in more detail in the appendix
(§ A).
2.1. Identifying Embedded Protostar Candidates
We form a sample of candidate embedded protostars
from the c2d catalogs; the first cut is based on the source
"class." All sources in the c2d catalogs are assigned a
class parameter based on colors, magnitudes, and stellar
SED fits (see the c2d Delivery Document (Evans et al.
2007) and Harvey et al. (2007a)). Class parameters in-
clude "star", "star+disk", "YSOc" (young stellar object
candidate), "red", "rising", "Galc" (galaxy candidate),
etc. Embedded protostars will generally be a subset of
YSOc sources, but some of the most embedded may also
be assigned to the "red" class if they are not detected in
all IRAC bands.
Thus, we begin by selecting all sources from the c2d
database that are classified as "YSOc" or "red." From
this list, we keep sources that meet all of the following
criteria:
(a) flux density at 24 µm (S24µm) ≥ 3 mJy
(b) S24µm ≥ 5αIR + 8 mJy, where αIR is the near-
to mid-infrared spectral index, determined by a least-
squares fit to photometry between 2 and 24 µm. This
criteria is motivated by a comparison to the carefully
vetted Serpens YSO sample of Harvey et al. (2007a).
(c) νS24µm > νS8µm, i.e., the SED is rising from 8 to
24 µm in νSν space.
(d) S24µm must be of high quality, i.e., signal to noise
(S/N) greater than 7.
(e) S24µm is not a "band-filled" flux. For sources not
originally detected in all Spitzer bands, a flux or up-
per limit is measured at the source position (band-filling,
Harvey et al. 2007a). Because the resolution is lower at
24 µm than at the shorter wavelengths, some IRAC-only
sources have unreliable band-filled fluxes at 24 µm (e.g.,
sources are confused with the PSF wings of a nearby
source).
In addition to sources that meet the above criteria, we
include any 70 µm point sources not classified as galaxy
candidates ("Galc"). Note that these 70µm sources need
not be classified as "red" or "YSOc". In each cloud, a
number of deeply embedded sources that are bright at
70 µm but very weak at 24 µm (e.g., HH211 in Perseus)
are recovered by this last criteria (5 in Perseus, 4 in
Serpens, and 3 in Ophiuchus), as are a few very bright
sources that are saturated at 24 µm (6 in Perseus, 2 in
Serpens, and 6 in Ophiuchus; these are often classified as
"rising").
Figure 2 plots S24µm versus αIR for embedded candi-
dates in Serpens, where "+" symbols indicate the orig-
inal ("YSOc"+"red"+ 70µm sources) sample for Ser-
pens, and boxes indicate our embedded protostar can-
didates after applying criteria (a)–(e). For compar-
ison, diamonds indicate the carefully vetted Serpens
YSO list from Harvey et al. (2007a). The majority of
"+"-only points, which were rejected as true YSOs by
Harvey et al. (2007a), are removed by criteria (a) and
(b) (shown as solid lines). More than half of the vetted
YSOs do not appear in the candidate embedded proto-
star sample (diamond, no box); the majority of these are
rejected by the rising SED criteria (c) and as they are
primarily classified as "star+disk" we do not expected
them to be embedded. There are a few sources in our
sample that are not in the vetted YSO sample (box, no
diamond). A few of these were easily identified as non-
protostellar, and rejected, when examining by eye. We
do identify three embedded protostar candidates in Ser-
pens that are not in the vetted YSO list but seem to be
associated with 1.1 mm emission (see §2.2): one "red",
and two "rising".
We include red sources to ensure that we identify the
most embedded protostars, but we need to reject the
large percentage of these sources that are likely to be
galaxies. As shown in Figure 3, criteria (a) and (b) are ef-
ficient at eliminating extragalactic sources. Here S24µm
is plotted versus αIR for all "Galc" galaxy candidates in
Serpens. As in Figure 2, lines show criteria (a) and (b)
After forming our embedded candidate samples based
on the above criteria, the images and SEDs of each source
are examined by eye to remove any galaxies that are
extended in the near-infrared, and other obviously non-
embedded sources. SEDs include 2MASS, IRAC, MIPS
and 1.1 mm fluxes. In some cases, there is no available
point-source flux at 70 or 160 µm even if there is emis-
sion at the position of the source, generally because the
source is extended at these wavelengths. In these cases a
flux density is measured by hand, if possible, using aper-
ture photometry. Large uncertainties (50% or more) are
Protostars in Serpens, Perseus, and Ophiuchus
5
Fig. 2.- Plot of S24µm versus spectral index αIR for sources in
Serpens, to demonstrate selection criteria for the embedded pro-
tostar candidate samples. Plus symbols ("+") indicate the sam-
ple from which the candidate embedded protostars is drawn, in-
cluding all catalog sources labeled as "YSOc" or "red" as well as
non-"Galc" 70µm sources. Lines show 24 µm flux cuts imposed
by criteria (a) and (b), which remove the majority of spurious or
background sources. Boxes indicate the final embedded protostar
sample, after applying criteria (a)-(d). Smaller blue diamonds show
the carefully vetted YSO sample from Harvey et al. (2006). The
two samples agree fairly well, but because we select for embed-
ded sources we miss many of the more evolved YSOs in the vetted
YSO list (Class II/III). There are a few sources in the embedded
protostar sample not in the Harvey et al. (2006) list, two of which
are likely to be real (see text), while the others are rejected when
examined by eye.
associated with these band-filled flux measurements.
2.2. Association with a 1.1 mm Core
The next step after assembling a sample of candidate
embedded protostars is to determine which are associ-
ated with 1.1 mm emission. The correlation between
candidate protostar and 1.1 mm core positions was done
in Enoch et al. (2008, see §2.3 and Figure 2), following
a similar analysis by Jørgensen et al. (2007); we summa-
rize the results here. We found that the coldest proto-
stellar candidates (Tbol < 300 K, see §3.1) are primarily
located within 1.0 × θ1mm of a millimeter core position,
where θ1mm is the angular FWHM size of the 1.1 mm
core, as expected if they are deeply embedded. Based on
that analysis, a given embedded protostar candidate is
assumed to be associated with a millimeter core if it is
located within 1.0 × θ1mm of the 1.1 mm core centroid
position.
We also found in Enoch et al. (2008), based on anal-
ysis of a spatially random distribution of sources, that
we can expect approximately 5 false associations with
1.1 mm cores in each cloud. Using a more restrictive cri-
teria (e.g., 0.5× θ1mm) would reduce the number of false
Fig. 3.- Plot of S24µm versus spectral index αIR for galaxy
candidates ("Galc") in Serpens. Selection criteria (a) and (b) are
overlaid, as in Figure 2. These selection criteria eliminate most of
the parameter space inhabited by extragalactic sources.
associations, but would likely miss at least a few embed-
ded protostars. For reference, adopting 0.5×θ1mm would
result in 10 fewer embedded protostars in Perseus, 7 in
Serpens, and 16 in Ophiuchus.
If an embedded protostar candidate is found to be as-
sociated with a millimeter core by the above criteria, the
1.1 mm flux is included in the protostellar SED, and is
If a
used to calculate an envelope mass (Menv, §3.2).
candidate is not located within 1.0 × θ1mm of a millime-
ter core position, we calculate a flux or upper limit from
the original 1.1 mm map (although sources with upper
limits at 1.1 mm are not included in the final lists of
embedded protostars given in Tables 2–4).
In a number of cases (10 in Perseus, 3 in Serpens, and
11 in Ophiuchus) there is clearly 1.1 mm flux at the pro-
tostar position that was not identified as a core in the
original 1.1 mm source extraction because it is below 5σ
(where σ is the local rms noise, see Paper I) or is in a con-
fused region of the map. If this emission exceeds 3σ, we
measure a "band-filled" 1.1 mm flux using a small aper-
ture (30′′–40′′). We also re-compute the 1.1 mm flux
density in small apertures for all sources in regions of
blended 1.1 mm emission. When more than one proto-
star candidate is associated with a single compact mil-
limeter core, we divide the total 1.1 mm flux of the core
equally between the protostellar sources.
The approach described above means that the 1.1 mm
flux is not measured in a fixed aperture for every source.
In crowded regions or for sources with close neighbors,
the 1.1 mm flux is measured in a 30′′–40′′diameter aper-
ture centered on the Spitzer source position. These
sources are indicated by a footnote in Tables 2–4. For
other sources, the Bolocam 1.1 mm core flux from Papers
I–III is used. Typically, but not always, the total core
flux is used here (integrated in the largest aperture, from
30′′− 120′′ diameters in steps of 10′′, that is smaller than
the distance to the nearest neighboring source); fluxes
6
Enoch et al.
in 30, 40, 80, and 120′′ apertures can also be found in
Papers I–III.
Approximately 50% of the embedded protostar candi-
dates in each cloud are lacking 1.1 mm emission, even
after re-examining the 1.1 mm maps at each source po-
sition. Most of these sources appear to be Class II ob-
jects, with little or no remaining envelope. Any 1.1 mm
emission from these objects, therefore, is likely below our
detection limit of ∼ 0.1M⊙ (§ 3.3). A few of these may
actually be low luminosity sources that are truly embed-
ded, with low mass envelopes below our detection limit.
For this reason, we only claim completeness to embed-
ded protostars with Menv > 0.1 M⊙. Note that this
corresponds well to the Crapsi et al. (2008) definition of
Stage I.
3. PROPERTIES OF EMBEDDED PROTOSTARS
3.1. Bolometric Luminosity and Temperature
Our extensive wavelength coverage allows us to con-
struct well-sampled SEDs for the embedded protostar
candidates in all three clouds, from which we calculate
a bolometric luminosity (Lbol) and bolometric tempera-
ture (Tbol) for each source. The bolometric luminosity is
calculated by integrating the SED (Sν ) over frequency:
Lbol = 4πd2Z Sνdν.
(1)
The bolometric temperature is defined as the tempera-
ture of a blackbody with the same mean frequency as the
source SED, and is given by
Tbol = 1.25 × 10−11hνi K
(Myers & Ladd 1993), where the mean frequency is
(2)
(3)
hνi = R νSνdν
R Sνdν
.
Two methods for approximating the integrations over
frequency for finitely sampled SEDs (midpoint and pris-
moidal) are discussed in the appendix (§A).
Tables 2–4 list the bolometric temperatures and lu-
minosities derived for all embedded protostars in each
cloud. As we are primarily interested in young, em-
bedded objects, only sources with detectable 1.1 mm
emission, which likely have substantial envelopes, are
included.
Sources are listed by increasing Tbol, and
identified as, e.g., "Per-emb#", as well as by their c2d
name (SSTc2dJ...), which also gives the position. When
sources are saturated at 24 or 70µm, the IRAS 25 or
60µm flux is utilized when not affected by blending. In
these cases a note is made in Tables 2–4. The correction
of saturated fluxes can increase the luminosity by more
than a factor of two; for saturated sources with no reli-
able IRAS flux (also noted in Tables 2–4), the luminosity
will be an underestimate.
All Tbol and Lbol values quoted use the midpoint in-
tegration method. The difference between the values
calculated by the midpoint and prismoidal integration
methods (given in parentheses in Tables 2–4) gives a
more realistic measure of the uncertainties in Tbol and
Lbol than the formal fitting errors (which are typically
10%). Uncertainties can be larger than 50% depending
on whether or not a 160 µm flux is available, and there
is an additional systematic uncertainty of 15–25% from
finite sampling errors (§A).
Systematic errors introduced by missing 160 µm fluxes
are investigated in §A. Without a 160 µm measurement,
Tbol will almost certainly be an overestimate for very cold
sources, which may bias our classification of protostellar
candidates (§4). Ophiuchus will be most affected, as only
four sources in that cloud have reliable 160 µm fluxes.
3.1.1. Correcting for Extinction
One might argue that we should correct our photom-
etry for extinction before calculating Lbol and Tbol. The
effects of extinction are typically ignored for Class 0 and
Class I sources, for which one might expect the fore-
ground extinction to be negligible relative to the ef-
fect of the envelope itself. While these observed val-
ues are most easily compared to the majority of previ-
ous work, Chen et al. (1995), who connected Tbol to the
classes defined by α, did correct the observed flux den-
sities for extinction before computing Tbol. Therefore,
we compute extinction corrected values for comparison
(see Evans et al. 2008 for a more complete discussion of
the effects of dereddening). We use the mean extinction
to all Class II sources in each cloud (AV = 5.92 mag in
Perseus, 9.57 mag in Serpens, an 9.76 mag in Ophiuchus;
Evans et al. 2008) to deredden the photometry of Class 0
and Class I sources.
As dereddening has a greater effect on shorter wave-
length photometry, correcting for extinction tends to in-
crease both Tbol and Lbol. Throughout this paper we
will primarily use observed values, and not extinction
corrected values, but the effects on derived values such
as the Class 0 lifetime (§ 9) will be noted. When used,
the extinction corrected bolometric luminosity and tem-
perature are indicated by T ′bol and L′bol, respectively.
3.2. Envelope Mass
The envelope mass of candidate protostars, Menv, is
calculated from the flux density at 1.1 mm, S1mm:
M =
d2S1mm
B1mm(TD)κ1mm
,
(4)
where d is the cloud distance, κ1mm = 0.0114 cm2 g−1
is the dust opacity per gram of gas at 1.1 mm, and
B1mm is the Planck function at a dust temperature of
TD. The opacity is interpolated from Table 1 column 5
of Ossenkopf & Henning (1994) for dust grains with thin
ice mantles.
We assume a dust temperature of TD = 15 K for proto-
stellar envelopes, consistent with average isothermal dust
temperatures found from radiative transfer models of a
sample of Class 0 and Class I protostars (Shirley et al.
2002; Young et al. 2003). The isothermal dust tempera-
ture is the temperature that, when used in an isothermal
mass equation (e.g., Eq. (4), above) yields the same mass
as does a detailed radiative transfer model that accounts
for temperature gradients. Dust temperatures will be
higher than 15 K close to the protostar, but the major-
ity of the envelope mass is in the outer, cooler, regions of
the envelope. A dust temperature of 10 K would result in
an increase in masses by a factor of 1.9, while TD = 20 K
would decrease masses by a factor of 1.5.
Protostars in Serpens, Perseus, and Ophiuchus
7
Fig. 4.- Near- to mid-IR spectral index (αIR) versus bolometric temperature (Tbol) for the candidate embedded protostar samples in
Perseus, Serpens, and Ophiuchus. Bold symbols indicate sources that are associated with 1.1 mm emission, and thin symbols denote those
with upper limits at 1.1 mm. Standard class divisions for both Tbol and αIR are shown. The two methods agree fairly well for Class II and
"warmer" Class I sources, but very cold (Tbol . 100 K) sources have a large range of αIR values.
Envelope masses of embedded protostars are listed in
Tables 2–4.
If the source is associated with a distinct
core, the Bolocam identification from Papers I–III is
given in the last column. If the source is not associated
with one of the originally identified cores, but rather the
flux has been "band-filled" at 1.1 mm, then no Bolocam
ID is given. In these "band-filled" cases, the measured
1.1 mm flux can be easily re-constructed from the mass.
3.3. Completeness
Because we require a detection at 1.1 mm to be in-
cluded in the final source lists, we are clearly incom-
plete to objects more evolved than Stage I, which do
not have a substantial protostellar envelope. We do de-
tect a few Class II sources at 1.1 mm; these are dis-
cussed in §5.1. The point-source detection limits of our
1.1 mm surveys limit our sensitivity to Class I sources
with Menv & 0.09 M⊙ in Perseus, Menv & 0.07 M⊙ in
Serpens, and Menv & 0.04 M⊙ in Ophiuchus. These are
5σ point source detection limits; for very extended en-
velopes the completeness limits will be higher. For sim-
plicity we take a detection limit of 0.1M⊙ for all clouds.
As the 1.1 mm detection requirement is more restric-
tive than our 24 µm flux criteria, we explore the possibil-
ity that we are missing some low luminosity embedded
sources that are below our 1.1 mm detection threshold.
Taking the 1.1 mm 5σ detection limits for each cloud
(75 mJy in Perseus, 50 mJy in Serpens, and 110 mJy in
Ophiuchus), and assuming the spectrum of a modified
blackbody, Sν = νβBν(TD), with TD = 15 K, we can
estimate the minimum detectable bolometric luminosity
in each cloud. For β = 1, the minimum Lbol is 0.02 L⊙ in
Perseus, 0.01 L⊙ in Serpens, and 0.01 L⊙ in Ophiuchus.
Assuming TD = 20 K lowers these values by approxi-
mately a factor of 2, while taking β = 2 increases them
by a factor of 4. Although these are very rough estimates,
they agree fairly well with the lowest observed bolometric
luminosities for sources with 1.1 mm detections (0.04 L⊙
in Perseus, 0.05 L⊙ in Serpens, and 0.01 L⊙ in Ophi-
uchus). Dunham et al. (2008) and Harvey et al. (2007b)
demonstrate that the Spitzer c2d surveys are complete
to young objects with luminosities as low as 0.05L⊙.
Finally, a comparison of our source list to the ded-
icated search for very low luminosity protostars by
Dunham et al. (2008) confirms that we are not missing
any embedded protostellar sources in Perseus down to
the completeness limits of that survey (Lbol & 0.03 −
0.05L⊙).
4. SOURCE CLASSIFICATION
To study the early evolution of protostars, it is nec-
essary to identify the evolutionary state of the embed-
ded candidates in our sample. This is typically ac-
complished by classifying sources into discrete groups
based on SED characteristics. A number of classifi-
cation methods are employed in the literature; most
often used are the near- to mid-infrared spectral
in-
dex αIR = d log(λFλ)/d log(λ) (Lada 1987), the bolo-
8
Enoch et al.
metric temperature Tbol (Myers & Ladd 1993, see §3.1),
and the ratio of submillimeter to bolometric luminos-
ity Lsubmm/Lbol (Andr´e et al. 1993). Generally, Lsubmm
is taken to be the integrated luminosity at wavelengths
λ ≥ 350 µm.
Protostars in Serpens, Perseus, and Ophiuchus
9
Fig. 5.- Distribution of Tbol as a function of the distance to the nearest 1.1 mm core, for candidate embedded protostars in all three
clouds. Distances are in units of the 1.1 mm core FWHM size (θ1mm), and symbols are as in Figure 4. Sources within 1.0 × θ1mm of a
1.1 mm core position are considered embedded within that core. There is a clear correlation between smaller distances, or more embedded
sources, and lower Tbol values, suggesting that Tbol is a good measure of evolutionary state for deeply embedded sources. For reference,
the total numbers of sources in each quadrant of the plot are given in parentheses.
10
Enoch et al.
Fig. 6.- Distribution of αIR as a function of the distance to the nearest 1.1 mm core for embedded protostar candidates in all three
clouds, similar to Figure 5. While there is some correlation between higher αIR values and smaller distances, it is not as compelling as the
correlation observed for Tbol (Figure 5). For reference, the total numbers of sources in each quadrant are given in parentheses.
Protostars in Serpens, Perseus, and Ophiuchus
11
Any of these methods must come with the caveat that
protostellar mass may affect the classification, which only
detailed modeling can resolve (e.g. Hatchell et al. 2007a).
We typically have only one flux measurement for λ >
160 µm, so an accurate determination of Lsubmm is not
feasible with these data.5 We will focus, therefore, on
Tbol and αIR.
4.1. Comparing Classification Methods
In Figure 4 we compare classifications based on αIR
and Tbol
for the candidate embedded protostars in
Perseus, Serpens, and Ophiuchus. Thick symbols indi-
cate sources associated with 1.1 mm emission, while thin
symbols denote sources with upper limits at 1.1 mm. The
spectral index αIR is determined from a least squares fit
to all detections between λ = 2µm and λ = 24 µm.
The calculation of Tbol is described in §3.1 and §A. We
adopt Tbol class divisions from Chen et al. (1995) (see
Table 1), and αIR divisions from Andr´e & Montmerle
(1994): αIR < 0 (Class I) and αIR > 0 (Class II). No
well-defined αIR criteria exists for Class 0 sources, as
deeply embedded objects were generally not visible in
the mid-IR prior to Spitzer.
The two classification methods agree fairly well for
Class II and "warmer" Class I sources. Sources with
Tbol . 100 K, by contrast, have a wide range of αIR
values, and a few of these "cold" sources even fall into
Class II based on the spectral index. One of the rea-
sons for this large scatter in αIR becomes apparent when
examining the SEDs of some deeply embedded sources,
which show considerable differences at short wavelengths.
Many protostellar SEDs are not monotonically increas-
ing from 3.6 to 24 µm, often falling from 5.8 to 8 µm and
rising again at longer wavelengths (e.g., Per-emb 18, Per-
emb 22, Oph-emb 1; Figure 1). Geometric effects such as
scattered light from an outflow cavity or absorption from
ices in the protostellar envelope are the most likely cause
of these features. In any case, non-monotonic behavior
at short wavelengths will clearly bias the calculation of
αIR. Calculating αIR from a straight line fit between 2
and 24 µm, rather than a least squares fit, still results in
a large range of αIR values at low Tbol.
To determine whether αIR or Tbol is a more accurate
measure of the true evolutionary state, we look at the cor-
relation between both measures and the degree to which
a given source is embedded. Figure 5 shows the distri-
bution of Tbol with respect to the distance from each
protostellar candidate to the nearest 1.1 mm core, in
units of the core FWHM size.
In all three clouds, es-
sentially all sources with Tbol . 200 K are located within
one core FWHM of a 1.1 mm peak. Thus the majority
of "cold" objects, as defined by low Tbol, appear to be
embedded in dense envelopes, and are likely to be at an
early evolutionary stage. Furthermore, there is a clear
correlation between smaller distance and lower Tbol, sug-
gesting that sources with lower Tbol are more embedded
than those with higher Tbol. This correlation remains
even if we exclude the 1.1 mm flux from the calculation
of Tbol, eliminating the possible bias between millimeter
5 It is possible to calculate Lsubmm by assuming a modified
blackbody spectrum and a value for β (e.g. Hatchell et al. 2007a),
but given the assumptions involved we choose not to pursue this
method.
cores and Tbol. Thus, if we affiliate the degree of embed-
dedness with youth, then Tbol appears to correlate well
with evolutionary state.
Figure 6 is similar, for αIR; while there is some corre-
lation between higher αIR values and smaller distances
to the nearest 1.1 mm core, the relationship is not nearly
as clear as for Tbol. In particular, there are a number of
sources with large αIR values that do not seem to be em-
bedded (the distance to the nearest core is much larger
than one FWHM), and there are several deeply embed-
ded sources with low αIR values. Based on Figures 4–6
and a visual examination of sources with Tbol < 70 K,
a number of which are known Class 0 sources, we con-
clude that Tbol is more reliable than αIR for classify-
ing deeply embedded protostars. Hereafter, Tbol will be
used to characterize protostellar sources, and we use the
divisions from Chen et al. (1995) and Table 1 to place
sources into Class 0, Class I, or Class II.
5. CHARACTERISTICS OF CLASS 0, I, AND II SOURCES
The total numbers of Class 0 and Class I sources in
each cloud, as defined by Tbol classifications, are given in
Table 5. Note that in addition to Tbol, we also require
Class I sources to be detected at 1.1 mm. This is not
an unreasonable requirement, as our definition of Stage I
dictates that they should be detected at millimeter wave-
lengths (Menv > 0.1 M⊙). The individual spectra (for
λ = 1.25 − 1100 µm) of all Class 0 and Class I sources
are shown in Figure 7.
Numbers in parentheses in Table 5 are based on clas-
sifications using the prismoidal, rather than midpoint,
method for calculating Tbol; these give some idea of
the uncertainties in the number of Class 0 and Class I
sources. Despite the sometimes substantial difference in
the midpoint and prismoidal Tbol values for individual
sources, there is very little difference in the resulting
number of Class 0 and Class I protostars, with the ex-
ception of Serpens where the number of Class 0 sources
increases from 10 to 14.
Statistics for Class II are not given; as we intended to
select against sources without a protostellar envelope, we
are necessarily incomplete to these objects. For example,
there are 46 sources in our original Serpens candidate
sample with Tbol > 650 K, while Harvey et al. (2007a)
find 132 Class II sources. We do detect a few Class II
sources at 1.1 mm, however (see §5.1).
The average spectra of Class 0, Class I, and Class II
sources in Perseus and Serpens are shown in Figure 8.
To calculate the average spectra, individual SEDs are
weighted by 1/Lbol, so that we are not biased by the
most luminous sources. Error bars in Figure 8 repre-
sent the 1σ error in the mean (σλ/√Nλ), but the dis-
persion in the sample (σλ) is much larger. Source SED
shapes do not fall into discrete bins, but form a con-
tinuous distribution between the averages shown. Error
bars are large at wavelengths where many sources are
not detected (at short wavelengths for Class 0, long wave-
lengths for Class II, and 160µm for all classes), and where
there are significant variations from source to source.
If Ophiuchus sources are included in the average calcu-
lation, the resulting average spectra are skewed toward
having more flux at short wavelengths, have larger dis-
persion, and the similarity of SEDs within each bin is
reduced. This behavior suggests that Tbol may be biased
12
Enoch et al.
Fig. 7.- Spectral energy distributions of all embedded protostars in Perseus, Serpens, and Ophiuchus, including photometry from 2MASS
(1.25, 1.65, 2.17 µm), IRAC (3.6, 4.5, 5.8, 8.0 µm), MIPS (24, 70, 160 µm), and Bolocam (1.1 mm) data. Also included are SCUBA 450
and 850 µm fluxes from the literature (Sandell & Knee 2001; Kirk et al. 2006) and IRAS 25 or 60 µm fluxes where they are used to replace
saturated 24 or 70 µm fluxes (see § 3.1). A sample is shown here; the full figure is available online only.
Protostars in Serpens, Perseus, and Ophiuchus
13
Fig. 8.- Average spectra of sources in Class 0, Class I, and Class II, for the Perseus and Serpens samples. Sources are classified based
on bolometric temperature, with the additional requirement that Class I sources be detected at 1.1 mm (see Table 1). Individual SEDs are
weighted by 1/Lbol in the average calculation. Error bars are the 1σ error in mean, and not the sample dispersion, which is much larger.
The Class II spectrum is not necessarily representative of all Class II objects, as we are very incomplete to sources without protostellar
envelopes.
for many sources in Ophiuchus, likely due to the lack of
information at 160 µm, so we exclude Ophiuchus from
the average spectra.
The progression from Class 0 to Class I to Class II
is consistent with a sequence of physical evolution.
In
particular, the average Class 0 spectrum has the low-
est flux densities from 1.25 − 24 µm, as expected for
deeply embedded sources with massive, extincting en-
velopes, and the highest fluxes at 70 − 1100 µm, where
reprocessed protostellar flux is emitted by the cold en-
velope. By contrast, the average Class II spectrum is
relatively flat, with a much larger percentage of the pro-
tostar flux emerging at shorter wavelengths, as expected
for older sources without much circumstellar material.
One must keep in mind, however, that Tbol is defined
such that spectra which peak at longer wavelengths will
have a lower Tbol.
Large error bars on the average Class 0 spectrum for
λ = 1−3 µm are indicative of the widely varying behavior
of Class 0 objects in the near-infrared, and the presence
of non-monotonic behavior is apparent at λ = 3.6 µm.
The large separation between the Class 0 and Class I
spectra, as well as the continuous range of SED shapes,
suggests that a more continuous means of estimating evo-
lutionary status is preferable to the standard classes.
5.1. Class II Sources with 1 mm Emission
The average Class II spectrum is not necessarily rep-
resentative of all Class II objects, due to severe incom-
pleteness to non-embedded sources. The Class II sources
we do detect are likely relatively young Stage II objects,
before a substantial fraction of disk mass is dispersed or
accreted.
Given our 1.1 mm sensitivity limit of approximately
0.1 M⊙, almost all sources detected at 1.1 mm will be
dense cores or envelopes around relatively young proto-
stars. In general, by the time a protostar has consumed
or dispersed its massive envelope and enters Stage II (or
the T Tauri phase, e.g., Adams et al. 1987), the rem-
nant disk of gas and dust has too little mass to be de-
tected by our millimeter surveys. Crapsi et al. (2008)
define Stage II as sources with a circumstellar mass be-
low 0.1 M⊙. Typical measured masses of Class II disks
are 0.01 − 0.1 M⊙ (e.g., Beckwith & Sargent 1996), al-
though values as large as 1 M⊙ have been measured
(Beckwith et al. 1990).
In a few a cases we do detect 1.1 mm emission around
sources with Class II-type SEDs. These objects have
bolometric temperatures Tbol > 650 K and a flux density
14
Enoch et al.
at 1.1 mm that is lower than the flux densities from 3.6
to 24 µm. With the exception of two "flat spectrum" ob-
jects (−0.3 < αIR < 0.3; Greene et al. 1994), the near-
to mid-infrared spectral indices, αIR, of these objects are
in the range −0.34 to −1.04, confirming their Class II sta-
tus. In some cases, the 1.1 mm emission is unresolved,
consistent with a compact disk. Often, however, these
Class II sources are in a region of confused millimeter
emission, so their physical association with the 1.1 mm
emission is not secure.
Table 6 lists all Class II sources (650 < Tbol < 2800 K)
in each cloud that are detected in our 1.1 mm surveys.
Those in confused regions of 1.1 mm emission are indi-
cated by a "∗" in the "Bolocam ID" column. Sources
with point-like 1.1 mm emission emission centered on
the Spitzer position (3 in Perseus, 1 in Serpens, and 3 in
Ophiuchus) may have massive disks.
5.2. Individual Sources
Here we briefly discuss a few examples of newly-
identified or otherwise interesting embedded protostellar
sources.
5.2.1. IRAS 03292+3039
IRAS 03292+3039 (Per-emb 2) is a little-studied,
Class 0 source associated with the 1.1 mm core Per-
Bolo 66; it was discussed briefly in Paper I. An image
of IRAS 03292+3039, together with the SED, is shown
in Figure 9. Jørgensen et al. (2006) identified this as an
outflow source, noting the large-scale outflow visible in
the 4.5 µm IRAC band.
Bright 1.1 mm emission centered on this object indi-
cates a fairly massive protostellar envelope (2.9 M⊙),
while the powerful outflow (Walawender et al. 2005;
Jørgensen et al. 2006; Hatchell et al. 2007b) and low
bolometric temperature (Tbol = 25 K) are evidence
of an extremely young, energetic embedded protostar.
The spectral energy distribution is similar to well-known
Class 0 protostars in Perseus such as NGC 1333-IRAS 4,
and nearby IRAS 03282+3035. The small fan-shaped
nebulosity visible at 3.6 and 4.5 µm (Figure 9) is most
easily explained by a cone-shaped cavity, carved out of
the dense envelope by an energetic outflow. The orien-
tation of the nebulosity corresponds well to the larger-
scale outflow traced by IRAC emission (Jørgensen et al.
2006), which is indicated by thick gray lines in Figure 9.
The one-sided nebulosity at 3.6 µm is strikingly similar
to the Whitney et al. (2003) model IRAC image of an
early Stage 0 sources viewed at an inclination angle of
30o (their Figure 12a).
5.2.2. Per-Bolo 102
Per-Bolo 102 is a bright 1.1 mm source that was iden-
tified in the Bolocam survey of Perseus (Enoch et al.
2006). It lies within the region of active star-formation
near IC 348, which includes the famous outflow-driving
source HH 211 (McCaughrean et al. 1994). The Spitzer
24 µm image resolves the luminous internal source into a
double object (Per-emb 16 and Per-emb 28). While the
two SEDs are very similar for λ ≥ 8 µm (Figure 9), vari-
ations at shorter wavelengths cause Per-emb 16 (Tbol =
56 K) to fall into Class 0, while Per-emb 28 (Tbol = 72 K)
falls just outside the Class 0/Class I boundary. These
sources are good examples of why a more continuous
evolutionary scheme is preferable to the standard clas-
sifications.
Recently obtained CARMA interferometric observa-
tions at λ = 2.7 mm resolve the millimeter core into
two sources, coincident with the Spitzer source positions
(M. L. Enoch et al., in preparation). The flux ratio at
2.7 mm of the northern (Per-emb 16) to southern (Per-
emb 28) source is at least 2:1, further evidence that Per-
emb 28 is slightly more evolved. Although not as massive
or cold as nearby HH 211 (Per-emb 1) and IC 348-mms
(Per-emb 11), Per-Bolo 102 is an interesting case study.
It may be a separate-envelope binary system, with two
nearly coeval Class 0 or early Class I sources. The sources
are separated by 17′′, or 4200 AU. Binary separations of
this order are consistent with early fragmentation in a
relatively dense cloud ("prompt initial fragmentation",
e.g., Pringle 1989; Looney et al. 2000), in which case the
individual sources would have distinct protostellar en-
velopes.
In a binary formed via gravitational fragmentation, we
would expect the separation to correspond to the local
Jeans length (Jeans 1928):
λJ = (cid:18) πc2
GµpmH n(cid:19)1/2
s
,
(5)
where cs is the local sound speed, and µp = 2.33 and n
are the mean molecular weight and mean particle density,
respectively. A Jeans length of 4200 AU would require a
relatively high density (n ∼ 6× 105 cm−3, assuming cs =
0.2 km s−1). The mean density of the Per-Bolo 102 core,
measured within an aperture of 104 AU, is 4× 105 cm−3,
close to the required value.
5.2.3. Ser-Bolo 33
II
One noteworthy example of a Class
source
associated with 1 mm emission is Ser-Bolo 33
(SSTc2dJ183006.12+004233.8) in Serpens (Figure 9), a
very bright Spitzer source with Lbol = 3.6 L⊙, Tbol =
871 K, and compact 1.1 mm emission centered on the
Spitzer position. The spectral index, αIR = −0.42, also
places this object in Class II. Vieira et al. (2003) included
this source in a sample of Herbig Ae/Be candidates, be-
lieved to be the intermediate mass Class II counterparts
of low-mass T Tauri objects, although the spectral type
is F3. The 1.1 mm mass calculated assuming an optically
thin disk is 0.17 M⊙, approximately 10% of the stellar
mass, M∗ ∼ 1.3 M⊙, estimated from the measured effec-
tive temperature (Teff ∼ 6300 K, Vieira et al. 2003) and
an empirical Teff − M relation (Habets & Heintze 1981).
6. ALTERNATIVE CLASSIFICATIONS
One advantage of such a large sample is that it allows
us to define a more continuous evolutionary sequence
than the standard classes that were appropriate for the
smaller samples previously available. With this in mind,
we divide the protostellar sources in each cloud into
smaller Tbol bins than those of the standard Class 0/I/II
divisions. Average spectra for "early Class 0", "late
Class 0", "early Class I", "late Class I", and Class II
(see Table 1 for definitions) sources in Perseus and Ser-
pens are shown in Figure 10. Error bars represent the 1σ
error in the mean, and average spectra are calculated as
Protostars in Serpens, Perseus, and Ophiuchus
15
Fig. 9.- Three-color Spitzer images and SEDs of individual sources discussed in §5.2.
Images are (8, 24, 70 µm), unless otherwise
noted, with 1.1 mm contours. Left: Three-color (3.6, 24, 70 µm) image of IRAS 03292+3039 (Per-emb 2). Fan-shaped nebulosity is visible
at 3.6 µm, most likely scattered light from a narrow outflow cavity in this deeply embedded Class 0 source. Thick lines indicate the
approximate orientation of the larger-scale outflow (Jørgensen et al. 2006). Center: Per-Bolo 102, a new candidate binary Class 0/I source.
Although the SEDs of the two embedded protostars look very similar at λ ≥ 8 µm, the southern source is brighter at shorter wavelengths,
fainter at 2.7 mm (see text), and has a slightly higher Tbol. Right: Ser-Bolo 33, a bright Class II object (Table 6). While this is clearly a
more evolved source, it is massive enough that we detect 1.1 mm emission from a compact disk.
16
Enoch et al.
Fig. 10.- Average spectra in designated Tbol bins (see Table 1), using sources from Perseus and Serpens. The average is calculated as
described in Figure 8 and the text, and error bars represent the error in the mean. The progression in SED shape for increasing Tbol is as
expected if this represents physical evolutionary sequence.
described in the previous section. The largest error bars
are seen for the shortest wavelengths in the lowest Tbol
bins, where NIR fluxes vary significantly from source to
source, likely depending on outflow opening angle and
viewing geometry. Binning of Tbol is based on general
agreement of SEDs in a given bin, as determined by eye,
and should not be interpreted as strict boundaries.
While there are significant similarities between the
SEDs within each bin, individual source SEDS fill the
continuum between the average spectra, and each aver-
age spectra lies within the 1σ dispersion of neighboring
Tbol bins. The average spectra change as one would ex-
pect if an extincting envelope is gradually accreted or dis-
persed, with the protostar becoming more visible at short
wavelengths. At all wavelengths except 24 and 70 µm,
the flux rises or falls monotonically with increasing Tbol.
For these intermediate wavelengths, the observed flux
may rise initially as hotter dust close to the protostar
is revealed, then fall as the mass of circumstellar mate-
rial drops. Despite the relatively narrow bins, there is
still a rather large change between "late Class 0" and
"early Class I", particularly at λ = 3.6 − 24 µm; this
transition may occur rapidly, or these wavelengths may
be particularly sensitive to geometry.
7. COMPARISON TO MODELS
Even for infinitely well-sampled SEDs, spectrum shape
is not necessarily directly correlated with age or de-
gree of embeddedness. For example, viewing geometry
can have a strong effect on SED shape.
In three di-
mensional radiative transfer models of (Whitney et al.
2003; Robitaille et al. 2006; Crapsi et al. 2008), Stage I
and Stage II sources can have quite similar SEDs when
viewed at the right inclination angle (e.g., when the ob-
server's line-of-sight intersects the outflow opening angle
of Stage I sources). Even Stage 0 sources can appear
much warmer if we happen to be looking directly into
the outflow cavity, although the probability of that oc-
curring is small.
We compare our spectra to the results of protostel-
lar models, which predict protostellar spectra based on
source age, mass, accretion rate, etc., both to gain in-
sight into the evolutionary state of sources and to eval-
uate the effectiveness of such models in matching ob-
served sources. Rather than model each source individu-
ally, we compare the average spectra from Figure 10 with
predicted spectra from Whitney et al. (2003) for "early
Stage 0", "late Stage 0" sources, etc. in Figure 11.
Whitney et al. (2003) begin with a set geometry for
Protostars in Serpens, Perseus, and Ophiuchus
17
Fig. 11.- Comparison of the average observed spectra from Figure 10 with predicted spectra from the protostellar evolution models of
Whitney et al. (2003) (plotted as νSν ). Distinct geometries are chosen to represent five evolutionary stages, and include a combination
of an accreting protostar, a protostellar disk, envelope, and a bipolar outflow. Colors indicate inclinations angles of the protostellar disk,
from 0o (pink) to 90o (dark green). With some exceptions the Stage 0 and Stage II models match the observed spectra fairly well, but the
Stage I models over-predict the near-IR and under-predict the mid-IR flux. In general the models tend to underestimate the millimeter
flux and thus the mass of the protostellar envelope.
18
Enoch et al.
each stage, including some combination of accreting pro-
tostar, flared protostellar disk, infalling envelope, bipolar
outflow, and grain models for each region, then use ra-
diative transfer modeling to predict protostellar spectra.
Colors in Figure 11 correspond to inclination angles of
the protostellar disk from 0o (pink) to 90o (green), where
the outflow is perpendicular to the disk. Model envelope
infall rates decline from 10−4 M⊙yr−1 in early Stage 0 to
10−6 M⊙yr−1 in late Stage I. Similarly, the disk radius
and cavity opening angle increase and the cavity density
decreases as one moves from early Stage 0 to late Stage I.
Both models and average observed spectra (thick black
lines) are scaled to a total luminosity of 1L⊙.
The model early Stage 0 and observed early Class 0
spectra agree fairly well. The observed spectrum lies
above most of the models at λ ≤ 24µm, which could
be explained by gaps in the inner envelope of some
protostars, allowing short wavelength flux to escape
(Jørgensen et al. 2005), or a wider outflow angle than
the model (5o). The Whitney et al. (2003) models pre-
dict that a small fraction (approximately 1/10) of early
Stage 0 sources will have spectra similar to Stage I, at
very low inclination angles (looking down the outflow).
The observed Class II spectrum is also consistent with at
least the low inclination models; as we select for sources
with 1.1 mm emission it is not unexpected that the ob-
served millimeter point is higher than predicted by the
models. For late Stage 0 the agreement is again pretty
good, except that the model under-predicts the millime-
ter flux, and thus the mass of the protostellar envelope,
a feature present in all spectra later than early Stage 0.
By contrast, the agreement between both Stage I mod-
els and the observed average spectra is quite poor. The
models severely over-predict the near-IR flux and under-
predict the mid-IR flux.
In fact, the characteristic
double-peaked profile of the Stage I models is almost
never seen in the observed spectra (Figure 7).
Indebetouw et al. (2006) demonstrate that envelopes
with clumpy density distributions can increase the ob-
served flux at 10 µm, with short wavelength emis-
sion escaping through low opacity regions between the
clumps. This kind of clumpy distribution eliminates
some of the "double-peaked" structure apparent in the
Whitney et al. (2003) models, which assume a smooth
density profile in the envelope and constant density in
the outflow cavity, although the clumpy models also in-
crease the flux at < 3µm. Alternatively, the presence of a
larger, thicker protostellar disk could flatten the Stage I
spectra by absorbing near-IR flux and producing more
mid-IR flux. A foreground cloud could also absorb near-
IR flux, without producing any mid-IR flux.
It seems that we still do not understand how sources
transition from Stage 0 to Stage I, at least in relation
to these models. Whether it is the physical model that
is unrealistic or how that physical geometry translates
to the observed spectra is not clear. It is important to
note that the same authors have developed a more com-
plete grid of models (Robitaille et al. 2006); we compare
here to the earlier models rather than fitting each av-
erage spectrum because we want an idea of the global
agreement with evolutionary stage.
8. PROTOSTELLAR EVOLUTION
8.1. Luminosity Evolution
We examine the evolution of embedded protostars in
more detail using the Lbol−Tbol diagram, the protostellar
equivalent of the H-R diagram (Myers et al. 1998). Here
Tbol is used as a measure of temperature rather than Tef f ,
which is not well-characterized for embedded sources.
As discussed in Myers et al. (1998), newly-formed pro-
tostars should begin at low Lbol and Tbol, increasing in
both Lbol and Tbol as accretion proceeds.
If accretion
abates or is otherwise halted, then Lbol will decrease for
steadily increasing Tbol. Eventually, sources will move
onto the main sequence, at Tbol & 3000 K.
Figure 12 plots Lbol versus Tbol for the candidate em-
bedded protostar samples in Perseus, Serpens, and Ophi-
uchus. Filled and open symbols are used for sources that
are associated with 1.1 mm emission, while "+" sym-
bols indicate sources with upper limits at 1.1 mm.6 All
sources detected at 1.1 mm are further divided accord-
ing to whether they have reliable measured 160 µm fluxes
(squares), are saturated at 160 µm (circles), or are not
detected at 160 µm (triangles). Given the discussion in
§A regarding the effect of missing 160 µm fluxes on the
calculation of Tbol, circles would be expected to move
up and to the right in this diagram for "cold" sources
(Tbol . 100 K), and down and to the left for "warm"
sources (Tbol & 100 K), if 160 µm fluxes were available.
While upper limits at 1.1 mm ("+" symbols) are ex-
pected for more evolved sources with Tbol & 500− 600 K,
colder sources with no 1.1 mm detection may be ei-
ther misclassified (e.g., background galaxies), or very
low mass sources whose 1.1 mm flux is below our de-
tection limit. Given our physical definition of Stage I
(M∗ & Menv; Menv > 0.1M⊙), embedded sources not
detected at 1.1 mm must have stellar masses less than a
few tenths of a solar mass, or have very little remaining
envelope (M∗ ≫ Menv).
Protostellar evolutionary tracks, which predict source
properties as a function of age and mass, can easily be
compared to our data using the Lbol − Tbol diagram.
Model tracks from Myers et al. (1998) (solid line; here-
after M98) and Young & Evans (2005) (dotted lines;
hereafter YE05) are shown in Figure 12. YE05 adopt
the standard inside-out collapse model of Shu (1977),
and assume that no mass is lost in the formation pro-
cess. From top to bottom, the YE05 models are for
sources with masses of 3.0, 1.0, and 0.3 M⊙, assuming
a constant accretion rate of dM/dt = c3
s/G, where cs is
the effective sound speed (Shu 1977). From the initial
singular isothermal sphere, finite masses are achieved by
truncating the outer radius of the envelope. A one di-
mensional radiative transfer model (DUSTY) is used to
calculate observational signatures (Lbol, Tbol, etc.) from
the accretion model.
Unlike YE05, M98 do not assume that that the entire
mass of the original core ends up in the final star, but
rather that a significant fraction of the core mass is lost
in the star formation process. The M98 model shown is
for a source with initial core mass of 1.8 M⊙ and final
stellar mass of 0.3 M⊙. M98 assume an accretion rate
that is initially dM/dt = c3
s/G, but falls off exponentially
6 Note that although we plot sources with 1.1 mm upper limits
here, our final embedded protostar samples (Tables 2–4) includes
only sources with 1.1 mm detections.
Protostars in Serpens, Perseus, and Ophiuchus
19
Fig. 12.- Bolometric luminosity versus bolometric temperature (Lbol − Tbol) diagram for embedded protostar candidates in Perseus,
Serpens, and Ophiuchus. Filled and open symbols indicate that a given source is associated with 1.1 mm emission, while "+" symbols
indicate upper limits at 1.1 mm. Symbol are further divided based on whether there is a reliable 160µm measurement, an upper limit, or
if we are unable to measure a 160µm flux due to saturation. The right axis shows the mass accretion rate, dM/dt, calculated from Lbol
using Eq. 6 and assuming M∗ = 0.5M⊙ and R∗ = 5R⊙. Model evolutionary tracks from Young & Evans (2005) for sources with 3.0, 1.0,
0.3 M⊙ (dotted lines, from top to bottom) and from Myers et al. (1998) for a source with final mass 0.3 M⊙ (solid line) are shown for
comparison. The large population of low-luminosity (Lbol < 0.1L⊙) Class I objects argues strongly for episodic accretion by Stage I.
with time, designed to match the observed luminosity of
pre-main sequence stars at Tbol & 3000 K. Thus, the lu-
minosity is significantly lower than the YE05 tracks at
later times. Both evolutionary models assume an accret-
ing central protostar, a circumstellar accretion disk, an
extended envelope, and a contribution to the luminosity
from gravitational contraction of the protostar. YE05
also include nuclear (Deuterium) burning.
Ignoring for a moment the population of Class I sources
in each cloud with Lbol values well below both mod-
els, the M98 model, for which a large fraction of the
core mass is ejected or dispersed, is more consistent with
the observed protostellar sources. The ratio of the fi-
nal stellar mass to initial core mass for the M98 model
shown (feff = Mcore/Mstar = 0.3/1.8 = 0.17) is smaller
than the values (feff = 0.3 ± 0.1) found by Alves et al.
(2007) and Enoch et al. (2008) by comparing the shape
of the core mass distribution to the initial mass function
(feff = 0.3 ± 0.1 and feff & 0.25, respectively).
The slightly better match to the M98 model may be
irrelevant, however, as neither a constant nor a steadily
decreasing accretion rate is consistent with the observed
protostellar populations in these clouds; many sources lie
below all four model tracks. Our data confirm and exac-
erbate the "luminosity problem" noted by Kenyon et al.
(1990), that Class I protostars in Taurus were observed
to have lower Lbol values than expected based on the av-
erage mass accretion rate required to make a 1M⊙ star.
In particular, the large population of Class I sources
with low Lbol in each cloud is difficult to understand
in relation to most existing protostellar evolutionary
models. A general feature of such models is that the
bolometric luminosity peaks in the Class I stage, a re-
sult that is true for constant accretion rates (YE05),
decreasing rates (M98), and gravo-turbulent models
(Froebrich et al. 2006). In contrast, we find quite similar
mean luminosities for the Class 0 and Class I samples:
2.4 L⊙and 2.2 L⊙, respectively. The standard devia-
tion for both samples is large: 3.5 L⊙ and 4.7 L⊙ for
Class 0 and Class I, respectively. The median luminosity
of Class I sources is a factor of 3 lower than the mean,
0.7 L⊙.
8.2. Episodic Accretion
Given the above discussion and the large observed
spread in Lbol of three orders of magnitude for Class I
sources, much larger than the range in envelope masses,
we conclude that mass accretion during the Class I stage
is episodic. Class I sources with low Lbol can be ex-
plained by periods of relative quiescence when the bolo-
metric luminosity, which is driven primarily by accre-
tion luminosity, drops by at least a factor of 10. Con-
versely, Class I sources with high Lbol values, e.g., those
that form the upper envelope of the distribution and ap-
pear to be consistent with the YE05 models, would cor-
respond to periods where the accretion rate is close to
the "Shu accretion" value. Approximately 20% of the
Class I sources have Lbol < 0.1L⊙ (7/39 in Perseus, 5/25
in Serpens, and 5/25 in Ophiuchus), and could be con-
sidered candidate very low luminosity objects (VeLLOs;
Dunham et al. 2008; Di Francesco et al. 2007).
Episodic accretion is not an unreasonable solution;
evidence for variable mass accretion and ejection is
plentiful,
including that based on modeling FU Orio-
20
Enoch et al.
Fig. 13.- Average spectra of low-luminosity Class I sources and "non-envelope Class I" sources (those that have Tbol < 650 K but no
1.1 mm emission). Average Class 0, I, and II spectra from Figure 8 are shown (dashed lines) for reference. The low-Lbol Class I sources have
lower short wavelength points relative to the 160 µm and 1.1 mm fluxes, as expected for low accretion rates, but are otherwise consistent
with the average Class I spectrum. The average "non-envelope Class I" spectrum is intermediate between the average Class I and Class II
spectra, as expected if these sources represent an transitional stage in which there is little or no remaining envelope.
nis eruptions (e.g., Hartmann & Kenyon 1985), and bow
shocks in Herbig-Haro outflows (e.g., Reipurth & Bally
2001). In traditional models of episodic accretion, infall
from the envelope onto the disk is constant and accre-
tion from the disk onto the protostar is episodic (e.g.,
Kenyon & Hartmann 1995), for example due to gravita-
tional instabilities in the disk (Vorobyov & Basu 2006).
Other scenarios such as the "spasmodic" infall model of
Tassis & Mouschovias (2005), where material is held up
by the magnetic field at the inner edge of the envelope
are also plausible.
Although the observed distribution of Lbol for Class I
sources argues strongly for episodic accretion, that is not
the only possibility. One alternative is that the low-Lbol
Class I sources are simply very low mass objects, and
we are somehow missing their low-mass Class 0 counter-
parts. Based on our sensitivity limits at 70, 160 µm and
1.1 mm, we should be able to detect Class 0 sources with
Lbol & 0.05L⊙. The lower limit to the observed Class 0
luminosity, however, may not be the internal luminosity,
but heating of the envelope by the interstellar radiation
field (ISRF), which can contribute as much as 0.2−0.3L⊙
to Lbol for envelope masses of 1 − 3M⊙ (YE05). For the
same reason, a Class 0 source with a very low mass accre-
tion rate would not necessarily have a very low bolomet-
ric luminosity, making it difficult to determine if episodic
accretion is already present at the Class 0 stage.
In Figure 13, we show the average spectrum of low-
luminosity (Lbol < 0.2L⊙) Class I sources in Perseus
and Serpens. On average, these sources have suppressed
short wavelength points relative the 160 µm and 1.1 mm
fluxes, but are otherwise consistent in shape with the
average Class I spectrum.7 The average low-Lbol spec-
trum is consistent with low accretion luminosities (evi-
dent at short wavelengths) in sources with "normal" en-
velope masses, as expected if the low luminosity objects
are similar to other Class I sources but with lower mass
accretion rates. The characteristics of low luminosity
protostars in Perseus, Serpens, and Ophiuchus are ana-
lyzed in Dunham et al. (2008).
8.3. Mass Accretion Rates
Assuming that the bolometric luminosity in Class 0
and Class I is due entirely to accretion, we can estimate
the accretion rate from Lbol:
,
(6)
M =
dM
dt ∼
2R∗Lbol
GM∗
where R∗ and M∗ are the radius and mass of the em-
bedded protostar. For the following we assume M∗ =
0.5M⊙ and R∗ = 5R⊙, but accretion rates can be eas-
ily scaled for different values of the mass or radius.
A mass of 0.5M⊙ is consistent with the initial mass
7 Note that the average spectra are normalized by 1/Lbol, so
on an absolute scale the low-Lbol spectrum would be substantially
fainter than the Class I spectrum.
Protostars in Serpens, Perseus, and Ophiuchus
21
function (e.g Chabrier 2003) and with average YSO
masses (Mer´ın et al. 2008; Spezzi et al. 2008); a radius
of 3− 5R⊙ is typical for pre-main sequence models of low
mass sources (e.g Palla & Stahler 1991; Robitaille et al.
2006).
For comparison the "Shu accretion rate" of c3
s/G,
which is used in the YE05 models,
is approximately
4 × 10−6 M⊙ yr−1 for cs = 0.2 km s−1. The M98 model
begins with M ∼ 10−6 M⊙ yr−1, falling to 10−9 M⊙
yr−1 by the time it reaches the main sequence. Making
a 1 M⊙ star in 5.4× 105 yr requires an average accretion
rate of approximately 2 × 10−6 M⊙ yr−1.
The mean luminosity of Class I sources corresponds to
M ∼ 1 − 2 × 10−6 M⊙ yr−1, not far from the average
required to make a solar mass star. By contrast, the
luminosity of the low-Lbol (Lbol ∼ 0.1L⊙) Class I sources
implies a mass accretion rate of only 7 × 10−8 M⊙ yr−1,
substantially smaller than the Shu value or the average
required to make a solar mass star. This result again
suggests that either these sources will form very low mass
stars (M∗ . 0.05M⊙)8, or that these sources are in a
If we
suppressed accretion phase of an episodic cycle.
define "sub-Shu" accretion to be M . 10−6 M⊙ yr−1, or
Lbol . 1L⊙, then approximately 55% of Class I sources
are observed to be in such a "sub-Shu" accretion state.
If the episodic accretion picture is correct there must be
a population of sources with accretion rates much higher
than the average; indeed, the highest observed luminosi-
ties (Lbol ∼ 10 − 30 L⊙) imply M ∼ 1 − 2 × 10−5 M⊙
yr−1, at least a factor of five higher than the Shu value
or the average required to make a 1M⊙ star. Approxi-
mately 5% of the Class I sources appear to be in such a
M & 10−5
"super-Shu" accretion rate (Lbol > 10L⊙, or
M⊙ yr−1.
Unfortunately, it is very difficult to determine the duty
cycle of the episodic accretion without detailed star for-
mation and accretion models. The small fraction of
sources observed to have very high accretion rates (5%
with M & 10−5 M⊙ yr−1), suggests that periods of rapid
accretion must be fairly short lived. Evans et al. (2008)
employ a simple accretion model to estimate the time
spent in accretion and quiescent phases, finding that half
the mass of a 0.5M⊙ star could be accreted during 7% of
the Class I lifetime.
8.4. Envelope Mass Evolution
Protostellar evolution models also predict the evolu-
tion of envelope mass, Menv, with Tbol, as shown in Fig-
ure 14. Symbols are similar to Figure 12, with upper
limits at 1.1 mm represented by arrows, and solid lines
indicating the YE05 evolutionary tracks (3.0, 1.0, and
0.3 M⊙ top to bottom). Some of the higher Tbol sources
have high 1.1 mm upper limits because they lie in re-
gions of extended or confused emission. Our determina-
tion that these sources are not actually associated with
the 1.1 mm emission is based on visual examination of
the images and SEDs.
Sources in all three clouds show a weak but consis-
tent trend of decreasing Menv with increasing Tbol, as
8 Although this would require these sources to have a quite low
core-to-star formation efficiency of . 10%, as only a few Class 0
sources have envelope masses less than 0.5M⊙.
expected if the envelope is gradually depleted by accre-
tion onto the protostar. This trend is not a result of in-
cluding the 1.1 mm flux, from which the envelope mass
is derived, in the calculation of Tbol; the same trend is
apparent even if the 1.1 mm point is excluded in the
calculation of Tbol.
Here, the YE05 model tracks fit the observed Menv −
Tbol distribution quite well. Thus, a constant envelope
infall rate reproduces the decrease in envelope mass with
increasing Tbol, although it does not fit the evolution of
Lbol with Tbol (Figure 12).
For standard episodic accretion models, where in-
fall from the envelope is steady and the accretion rate
from the disk to the protostar is the variable quantity
(e.g., Kenyon & Hartmann 1995), the envelope mass will
steadily decrease with increasing Tbol even for variable
accretion. Note that accretion stops when the enve-
lope mass, as defined by the outer radius, has been ex-
hausted. In the context of the YE05 models, the spread
of Menv as a function of Tbol suggests stellar masses in
the range M ∼ 0.3 − 4 M⊙ in Perseus and Serpens, and
M ∼ 0.1 − 1.5 M⊙ in Ophiuchus.
9. LIFETIME OF THE CLASS 0 PHASE
The length of time that sources spend in the Class 0
phase is an important diagnostic of protostellar evolu-
tion and how accretion rates evolve with time.
If the
rate of star formation in these clouds is steady in time
(i.e., not occurring in bursts), and if Class 0–Class I–
Class II represents a true evolutionary sequence, then
we can use the number of objects in consecutive evolu-
tionary phases to estimate the relative lifetimes of those
phases: t1/t2 = N1/N2. As we ultimately calibrate life-
times based on the lifetime of Class II disks (∼ 2×106 yr;
Kenyon et al. 1990; Cieza et al. 2007; Spezzi et al. 2008),
star formation must have been steady in time for at least
the last 2 Myr.
In addition, there can be no signifi-
cant dependence of the evolutionary timescales on source
mass.
It is unlikely that all of these assumptions hold in ev-
ery star forming region. In fact there is some observa-
tional evidence for mass-dependent evolution of dense
cores (Hatchell et al. 2008), which could easily translate
into mass-dependent evolution after protostar formation.
We can, however, hope to mitigate the effects of any
breakdowns in our assumptions by utilizing our large
sample and averaging over three different environments.
In a companion paper (Enoch et al. 2008) we use a
similar argument to derive the lifetime of the prestellar
phase from the ratio of the number of starless cores to the
total number of embedded protostars (Class 0 + Class I).
There we find there that the lifetime of dense (n & 104
cm−3) prestellar cores is approximately equal to the life-
time of the embedded protostellar phase, or 2−5×105 yr,
suggesting that such cores evolve dynamically over a few
free-fall timescales.
The relative number of Class 0 and Class I sources in
each cloud, and the Class 0 lifetime derived from that
ratio, is given in Table 5. Given in parentheses are the
values resulting from using the prismoidal, rather than
midpoint method for determining Tbol; the Class 0 life-
time is based on an average of the two methods, and the
difference between them, added in quadrature with √N
22
Enoch et al.
Fig. 14.- Envelope mass versus bolometric temperature for embedded protostars in Perseus, Serpens, and Ophiuchus. Filled and open
symbols indicate that a given protostar is associated with 1.1 mm emission, while sources with upper limits at 1.1 mm are plotted as
arrows. Solid lines show the predictions of protostellar evolutionary models from Young & Evans (2005) for sources of mass (from top to
bottom) 3.0, 1.0, 0.3 M⊙. The Young & Evans (2005) models describe the evolution of envelope mass with Tbol quite well, for sources of
mass M ∼ 0.3 − 4 M⊙ in Perseus and Serpens, and M ∼ 0.1 − 1.5 M⊙ in Ophiuchus.
statistical uncertainties, yields an uncertainty for the life-
time.
I
sources
There are approximately half as many Class 0
as Class
in both Perseus and Serpens
(NClass 0/NClass I = 0.7 and 0.4, respectively), suggest-
ing that the Class 0 phase lasts roughly half as long as
the Class I phase. We adopt a total embedded phase
lifetime of temb = tCl0 + tClI ∼ 5.4 × 105 yr, derived
based on the relative number of embedded protostars and
Class II sources (Evans et al. 2008). Thus our measured
ratios imply Class 0 lifetimes of tClass 0 ∼ 2.2 × 105 yr in
Perseus, and 1.7× 105 yr in Serpens. In Ophiuchus there
are only 3 or 4 Class 0 sources, resulting in a ratio of
NClass 0/NClass I = 0.1 − 0.2 and tClass 0 ∼ 0.7 × 105 yr.
Taking all three clouds together, we find a lifetime for
the Class 0 phase of 1.72 ± 0.25 × 105 yr. This value is
significantly longer than a number of previous estimates
of tClass 0 ∼ 104 yr, based both on the number of Class 0
sources in Ophiuchus (Andr´e & Montmerle 1994), and
on comparison to evolutionary models (2 − 6 × 104 yr;
Froebrich et al. 2006). A short Class 0 lifetime is gen-
erally interpreted as evidence for a period of very rapid
accretion early on in the evolution of protostars, caus-
ing them to quickly reach Stage I, at which point the
accretion rate decreases significantly. Our results argue
against such a rapid accretion phase. Although accretion
may decrease somewhat in Class I (or become episodic,
see §8), it appears unlikely that the average accretion
rate drops by more than a factor of two from Class 0 to
Class I, based on the relative lifetimes of the two phases.
Similar mean luminosities for the Class 0 and Class I
phases also argue against very high accretion rates in
Class 0.
Our derived Class 0 lifetime is similar to the results of
Visser et al. (2002) for a sample of Lynds dark clouds
For example,
(tClass 0 ∼ 2 × 105 yr), and to the recent findings
of Hatchell et al. (2007a) that the Class 0 lifetime in
Perseus is similar to the Class I lifetime (2.5−6.7×105 yr
with 95% confidence).
Our large, unbiased sample provides a distinct advan-
tage over many other previous studies, which have nec-
essarily relied on small samples or accumulated sources
from a number of different surveys, wavelengths, and
detection methods.
the Visser et al.
(2002) Class 0 sample consists of 7 sources, and the
Andr´e & Montmerle (1994) lifetime for Ophiuchus is
based on one Class 0 object. Froebrich et al. (2006) note
that their 50 Class 0/I sources are selected from a variety
of surveys including NIR imaging of outflows, IRAS data,
submillimeter and millimeter mapping, and radio contin-
uum surveys, causing their source sample to be subject
to strong selection effects. Within each cloud our surveys
are very uniform, providing protostellar samples that are
envelope mass limited and not biased by selection effects.
For this reason, our estimated Class 0 lifetimes should be
more robust than most previous measurements. The re-
cent work by Hatchell et al. (2007a) comparing SCUBA
850 µm maps and Spitzer c2d data of Perseus, with 34
Class 0 sources, is a notable exception.
Classifying sources based on T ′bol, calculated from pho-
tometry corrected for extinction (see § 3.1.1) results in
a somewhat smaller number of Class 0 sources, and a
slightly shorter Class 0 lifetime. Table 7 gives the class
statistics and corresponding Class 0 lifetime derived us-
ing the extinction corrected photometry. The three-cloud
average Class 0 lifetime is shorter than when using ob-
served photometry, by approximately 35% (t′class 0 =
1.1 × 105 yr).
9.1. Cloud to Cloud Differences: Ophiuchus
Protostars in Serpens, Perseus, and Ophiuchus
23
Fig. 15.- Evolution of the bolometric temperature with time, based on estimated lifetimes for four Tbol intervals from Tbol < 50 K to
650 K. Tbol ranges, inferred lifetimes, and uncertainties are taken from Table 8, and plotted as thick gray lines and symbols. The dashed line
connects to the Class II lifetime at t = 2×106 yr and Tbol = 2800 K. We fit a simple function to Tbol versus time: Tbol = 25 K+C(t/105 yr)n;
the best fit is for Tbol ∝ t1.8.
The results in Ophiuchus are strikingly different from
Perseus and Serpens, with nearly 10 times more Class I
than Class 0 sources. Our results for Ophiuchus are sim-
ilar to previously observed ratios of NClass 0/NClass I ∼
1/10 in that cloud (Andr´e & Montmerle 1994). The de-
rived Class 0 lifetime, tClass 0 ∼ 7 × 104 yr, is still sub-
stantially longer than the very short tClass 0 ∼ 104 yr
found by Andr´e & Montmerle (1994), however.
There are two obvious, but conflicting, possible expla-
nations for the smaller NClass 0/NClass I ratio in Ophi-
uchus. First, the star formation rate may be temporally
variable in Ophiuchus. Visser et al. (2002) suggest that
a burst of star formation approximately 105 yr ago is re-
sponsible for the large ratio of Class I to Class 0 sources.
Alternatively, the Class 0 phase may be much shorter in
Ophiuchus than in the other two clouds due to higher ac-
cretion rates at early times. Were this the case, however,
we would expect the mean luminosity to be significantly
higher for the Class 0 sources than for the Class I sources,
which is not observed. While a burst of star formation
seems the more likely explanation, accretion rates, and
thus lifetimes, could conceivably depend on cloud envi-
ronmental factors such as mean density or turbulent ve-
locity.
On the other hand, the observed number of Class 0
sources in Ophiuchus may be biased by the dearth of
160 µm flux measurements in that cloud. As discussed
in §A, bolometric temperatures are likely overestimated
for sources without a 160 µm fluxes, which may cause
up to 50% of Class 0 sources to be classified as Class I.
The lack of 160 µm fluxes is especially problematic in
Ophiuchus, where the majority of sources are either sat-
urated at 160 µm or in regions of poor coverage, and may
be partially responsible for the small number of observed
Class 0 sources.
9.2. Limitations on the Class I Sample
Recall that we only include sources in our Class I sam-
ple if they are detected at 1.1 mm. This criteria excludes
24 sources from the original candidate protostar sample
in Perseus with 70 < Tbol < 650 K, 18 sources in Ser-
pens, and 36 in Ophiuchus. The average spectrum of
these "non-envelope Class I" sources is shown in Fig-
ure 13 (Perseus and Serpens only), and appears to be
intermediate between the Class I and Class II averages.
Furthermore, the mean Tbol values of the non-envelope
Class Is (570 K in Perseus, 520 K in Serpens, and 500 K
in Ophiuchus) confirm that they fall at the warm end of
the Class I distribution. While it is possible that we are
simply not detecting the envelopes of these sources (e.g.,
if Menv . 0.1 M⊙ or they are very diffuse), we suggest
that the transition from Stage I (protostars that retain
an envelope) to Stage II (pre-main sequence stars with
no envelope) occurs closer to Tbol = 400 − 500 K than
the standard Tbol = 650 K, for these data.
24
Enoch et al.
When we add the "non-envelope Class I" sources to our
Class I samples, dropping the requirement that they be
detected at 1.1 mm, the Class 0 to Class I ratio becomes
0.5 in Perseus, 0.3 in Serpens, and 0.07 in Ophiuchus.
This represents a 25–30% decrease in the Class 0 lifetime
for each cloud. On the other hand, Crapsi et al. (2008)
suggest based on protostellar models that up to half of
sources observationally classified as Class I may actually
be Stage II source with disk inclinations & 65o. Our
"non-envelope Class I" objects may be just such sources;
if the Crapsi et al. (2008) picture is correct, the observed
number of Class I sources likely represents an upper limit
to the true number of Stage I protostars.
9.3. Evolution of Tbol with Time
In Table 8 we give the number of sources in each cloud
in our narrow Tbol bins ("early Class 0", "late Class 0",
etc.; §6). The ratio of the total number in each bin,
summed over the three clouds, to the total number of
embedded protostars is used to derive a lifetime for each
Tbol range. The mean Tbol within that range is also given.
With the exception of "early Class 0", there is no ev-
idence for a difference in the number of sources in each
Tbol bin, and little difference in the derived lifetimes. No-
tably, 75% of the embedded phase lifetime has elapsed
by the time the bolometric temperature reaches 300 K,
and the mean Tbol in "Late Class I" (400 K) is skewed
toward the lower temperature end of the bin. These
features support our earlier suggestion that the divid-
ing line between Stage I and Stage II is probably closer
to Tbol = 400 − 500 K than to 650 K. Tbol = 50 K is
reached at quite early times, which is not unexpected if
the temperature scale starts at approximately 10 K with
starless cores.
With lifetime measurements in several Tbol intervals,
together with our previous conclusion that the average
accretion rate is approximately constant through the em-
bedded phase, we can empirically characterize the evo-
lution of Tbol with time. Figure 15 plots the evolution
of the bolometric temperature with time, based on val-
ues from Table 8. The relationship is clearly non-linear;
given the small number of observed points, we fit a simple
function to Tbol as a function of time:
Tbol = 25 K + C(cid:18) t
105 yr(cid:19)n
,
(7)
where C is constant and we choose 25 K as the tempera-
ture at t = 0 because that is the minimum observed Tbol
in Class 0. The best fit is for C = 51 K and Tbol ∝ t1.8,
indicating a fairly steep evolution with time. The depen-
dence on time must flatten out significantly after the em-
bedded phase, however, to match the lifetime for Class II
(dashed line in Figure 15).
10. CONCLUSIONS
Utilizing large-scale 1.1 mm surveys (Enoch et al.
2006; Young et al. 2006; Enoch et al. 2007) together with
Spitzer IRAC and MIPS maps from the c2d Legacy
program (Evans et al. 2003), we have constructed an
unbiased census of deeply embedded protostars in the
Perseus, Serpens, and Ophiuchus molecular clouds. Our
sample includes a total of 39 Class 0 sources and 89
Class I sources, with approximate detection limits of
Menv & 0.1 M⊙ and Lbol & 0.05L⊙ for the envelope
mass and bolometric luminosity, respectively. We also
detect a few Class II and Herbig Ae/Be candidates at
1.1 mm, most likely evidence of fairly massive proto-
planetary disks.
Bolometric luminosities, temperatures, and envelope
masses are calculated for the candidate Class 0 and
Class I sources in each cloud. We compare protostel-
lar classification methods, concluding that, for deeply
embedded sources, the bolometric temperature Tbol is a
better measure of evolutionary state than the near- to
mid-infrared spectral index, αIR.
We also explore classifying sources into "early Class
0", "late Class I", etc., based on dividing them into nar-
rower Tbol bins. Average observed spectra in these bins
are compared to model predictions from Whitney et al.
(2003) for "early Stage 0", "late Stage 0", etc. In a broad
sense the Stage 0 and Stage II models match the observed
spectra fairly well. The agreement with both Stage I
models is quite poor, however, as the models severely
over-predict the near-IR and under-predict the mid-IR
flux at this stage, displaying a double-peaked SED that
is rarely observed.
Observed source properties are compared to pro-
tostellar evolutionary models using the bolometric
temperature-luminosity (Lbol − Tbol) diagram, the pro-
tostellar equivalent of the H-R diagram (Myers et al.
1998). Neither models with a constant mass accretion
rate (Young & Evans 2005), nor those with an exponen-
tially decreasing rate (Myers et al. 1998) fit the observed
sources.
In particular, there is a large population of low lu-
minosity Class I sources that aggravate the previously
noted "luminosity problem" for embedded protostars
(e.g. Kenyon et al. 1990). We interpret this result as ev-
idence for episodic accretion beginning at least by the
Class I phase, and possibly earlier. More than 50% of
Class I sources are inferred to have "sub-Shu" mass ac-
cretion rates ( M . 10−6 M⊙ yr−1, corresponding to
M . 10−7
Lbol . 1L⊙), and approximately 20% have
M⊙ yr−1. To build up of order a solar mass in 5.4×105 yr,
such sources must also have periods of "super-Shu" ac-
cretion ( M & 10−5 L⊙yr−1). Few very high luminosity
sources are observed (5%), suggesting that such rapid ac-
cretion periods must be short lived. An important caveat
to this analysis is that we may sometimes underestimate
the luminosity due to missing 70 or 160 µm photometry.
Finally, the relative number of Class 0 and Class I
sources are used to estimate the lifetime of the Class 0
phases. There are approximately half as many Class 0
as Class I sources in the three cloud sample, implying an
average Class 0 lifetime of 1.7 ± 0.3 × 105 yr (1.1 × 105
yr when approximate extinction corrections are applied).
This lifetime rules out drastic changes in the mass accre-
tion rate from Class 0 to Class I, particularly extremely
rapid accretion in Class 0. In Ophiuchus the fraction of
Class 0 sources is much smaller. While this difference
could be due to the lack of 160 µm flux measurements
in Ophiuchus, it may be that either the Class 0 phase is
shorter in that cloud (0.7 × 105 yr), or that a burst of
star formation is responsible for the large population of
Class I objects (e.g. as suggested by Visser et al. 2002).
Altogether, the large variation in Class I luminosities,
Protostars in Serpens, Perseus, and Ophiuchus
25
similar mean luminosities for the Class 0 and Class I
phases, and not dramatically different numbers of Class 0
and Class I sources (at least in Perseus and Serpens)
suggests a consistent picture of nearly constant average
mass accretion rate through the entire embedded phase
(Stage 0 and Stage I), with highly variable episodic ac-
cretion turning on by at least early Stage I, and possibly
sooner. Understanding embedded protostellar structure
and evolution well enough to reproduce the observations
with detailed models of spectra and evolution presents
an ongoing challenge.
The authors thank J. Hatchell, Y. Shirley, and the
anonymous referee for comments and suggestions that
helped to improve this paper, as well as M. Dunham for
many fruitful discussions. We are grateful to B. Whit-
ney for sharing the protostellar evolutionary model data
used here. We thank the Lorentz Center in Leiden for
hosting meetings that contributed to this paper. Part
of the work was done while in residence at the Kavli
Institute for Theoretical Physics in Santa Barbara, Cali-
fornia. Support for this work, part of the Spitzer Legacy
Science Program, was provided by NASA through con-
tracts 1224608 and 1230782 issued by the Jet Propul-
sion Laboratory, California Institute of Technology, un-
der NASA contract 1407. Additional support was pro-
vided by NASA through the Spitzer Space Telescope
Fellowship Program and obtained from NASA Origins
Grants NNG04GG24G and NNX07AJ72G to the Uni-
versity of Texas at Austin. Support for the development
of Bolocam was provided by NSF grants AST-9980846
and AST-0206158.
REFERENCES
Adams, F. C., Lada, C. J., & Shu, F. H. 1987, ApJ, 312, 788
Alves, J., Lombardi, M., & Lada, C. J. 2007, A&A, 462L, 17
Andr´e, P. 1994, The Cold Universe, editors Montmerle, T., Lada,
C. J., Mirabel, I. F., & Tran Thanh Van, J. Gif-sur-Yvette,
Editions Frontieres, p. 179
Andr´e, P., & Montmerle, T. 1994, ApJ, 420, 837
Andr´e, P., Ward-Thompson, D., & Barsony, M. 1993, ApJ, 406,
Hatchell, J., Fuller, G. A., Richer, J. S., Harries, T. J., & Ladd,
E. F. 2007a, A&A, 468, 1009
Hatchell, J., Fuller, G. A., & Richer, J. S. 2007b, A&A, 472, 187
Hatchell, J., & Fuller, G. 2008, A&A, in press (preprint:
arXiv:0803.1064 [astro-ph])
Herbig, G. H. & Kameswara, R. N. 1972, ApJ, 174, 401
Indebetouw, R., Whitney, B. A., Johnson, K. E., & Wood, K.
122
Aspin, C., Sandell, G., & Russell, A. P. G. 1994, A&AS 106 165
Beckwith, S. V. W. & Sargent, A. I. 1996, Nature, 383, 139
Beckwith, S. V. W., Sargent, A. I., Chini, R. S., & Gusten, R.
1990, AJ, 99, 924
Chabrier, G. 2003, PASP, 115, 763
Chen, H., Myers, P. C., Ladd, E. F., & Wood, D. O. S. 1995,
ApJ, 445, 377
Cieza, L., et al. 2007, ApJ, 667, 308
Cohen, M. & Kuhli, L. V. 1979, ApJS, 41, 743
Crutcher, R. M. 1999, ApJ, 520, 706
Crapsi, A., van Dishoeck, E. F., Hogerheijde, M. R.,
2006, ApJ, 636, 362
Jayawardhana, R., Hartmann, L., Calvet, N. 2001, ApJ, 548, 310
Jeans, J. H. 1928, Astronomy and Cosmogony, p. 340.
Cambridge, U.K.: Cambridge University Press
Jørgensen, J. K., Bourke, T. L., Myers, P. C., Schoier, F. L., van
Dishoeck, E. F., Wilner, D. J. 2005, ApJ, 631, L77
Jørgensen, J. K., Harvey, P. M., Evans, N. J., II, Huard, T. L.,
Allen, L. E., Porras, A., Blake, G. A., et al. 2006, ApJ, 645,
1246
Jørgensen, J. K., Johnstone, D., Kirk, H., & Myers, P. C. 2007,
ApJ, 656, 293
Jørgensen, J. K., et al. 2008, ApJ, in press (preprint:
Pontoppidan, K. M., & Dullemond, C. P. 2008, A&A, in press
arXiv:0805.0599)
Di Francesco, J., Evans, N. J., II, Caselli, P., Myers, P. C.,
Shirley, Y., Aikawa, Y., & Tafalla, M. 2007, in Protostars and
Planets V, editors B. Reipurth, D. Jewitt, and K. Keil, p. 17
Dunham, M. M., Crapsi, A., Evans, N. J., II, Bourke, T. L.,
Huard, T. L., Myers, P. C., & Kauffmann, J., 2008, ApJS, in
press (preprint: arXiv:0806.1754)
Ebert, R. 1955, Zeitschrift Astrophysics, 37, 217
Enoch, M. L., Young, K. E., Glenn, J., Evans, N. J., II, Golwala,
S., Sargent, A. I., Harvey, P., et al. 2006, ApJ, 638, 293
Enoch, M. L., Glenn, J., Evans, N. J., II, Sargent, A. I., Young,
K. E., & Huard, T. L. 2007, ApJ, 666, 982
Enoch, M. L., Evans, N. J., II, Sargent, A. I., Glenn, J.,
Rosolowsky, E., & Myers, P. C., ApJ, 684, 1240
Evans, N. J., II, Allen, L. E., Blake, G. A., Boogert, A. C. A.,
Bourke, T., Harvey, P. M., Kessler, J. E., et al. 2003, PASP,
115, 965
Evans, N. J., II, et al. 2007, Final Delivery of Data from the c2d
Legacy Project: IRAC and MIPS
Evans, N. J., II, et al. 2008, ApJ, submitted
Froebrich, D., Schmeja, S., Smith, M. D., & Klessen, R. S. 2006,
MNRAS, 368, 435
Greene, T. P., Wilking, B. A., Andr´e, P., Young, E. T., & Lada,
C. J. 1994, ApJ, 434, 614
Greene, T. P. & Young, E. T. 1992, ApJ, 395, 516
Habets, G. M. H. J., & Heintze, J. R. W. 1981, A&A, 46, 193
Hartmann, L., & Kenyon, S. J. 1985, ApJ, 299, 462
Harvey, P. M., Chapman, N., Lai, S.-P.. Evans, N. J., II, Allen, L.
E., Jørgensen, J. K., Mundy, L. G., et al. 2006, ApJ, 644, 307
Harvey, P. M., Merin, B., Huard, T. L., Rebull, L. M., Chapman,
N. Evans, N. J., II, & Myers, P. C. 2007a, ApJ, 663, 1149
Harvey, P. M., Rebull, L. M., Brooke, T., Spiesman, W. J.,
Chapman, N., Huard, T. L., Evans, N. J., II, et al. 2007b, ApJ,
663, 1139
Kenyon, S. J., & Hartmann, L. W. 1995, ApJS, 101, 117
Kenyon, S. J., Hartmann, L. W., Strom, K. M., & Strom, S. E.
1990, AJ, 99, 869
Kirk, H., Johnstone, D., & DiFrancesco, J. 2006, ApJ, 646, 1009
Lada, C. J. 1987, in IAU Symp. 115, Star Forming Regions,
editors M. Peimbert & J. Jugaku (Dordrecht: Reidel), p. 1
Lada, C. J., & Lada, E. A. 2003, ARA&A, 41, 57
Looney, L. W., Mundy, L. G., & Welch, W. J. 2000, ApJ, 529, 477
Mer´ın, B., et al. 2008, ApJ, in press (preprint: arXiv:0803.1504)
McCaughrean, M. J., Rayner, J. T., & Zinnecker, H. 1994, ApJ,
436, L189
McKee, C. F., & Zweibel, E. G. 1992, ApJ, 399, 551
Myers, P. C., Adams, F. C., Chen, H., Schaff, E. 1998, ApJ, 492,
703
Myers, P. C., & Ladd, E. F. 1993, ApJ, 413, 47
Natta, A., Testi, L., Calvet, N., Henning, Th., Waters, R., &
Wilner, D. 2007, in Protostars and Planets V, editors B.
Reipurth, D. Jewitt, and K. Keil, p. 767
Ossenkopf, V., & Henning, Th. 1994, A&A, 291, 943
Palla, F. & Stahler, S. W. 1991, ApJ, 375, 288
Pringle, J. E. 1989, MNRAS, 239, 361
Rebull, L. M., et al. 2007, ApJS, 171 447
Reipurth, B., & Bally, J. 2001, ARA&A, 39, 403
Robitaille, T. P.. Whitney, B. A., Indebetouw, R., Wood, K., &
Denzmore, P. 2006, ApJS, 167, 256
Sandell, G. & Knee, L. B. G. 2001, ApJ, 546, L49
Shirley, Y. L., Evans, N. J., II, & Rawlings, J. M. C. 2002, ApJ,
575, 337
Shu, F. H. 1977, ApJ, 214, 488
Spezzi, L., et al. 2008, ApJ, 680, 1295
Spitzer, L. Jr. 1978, Physical Processes in the Interstellar
Medium, New York: Wiley, p. 282
Tassis, K., & Mouschovias, T. Ch. 2005, ApJ, 618, 783
26
Enoch et al.
Vieira, S. L. A., Corradi, W. J. B., Alencar, S. H. P., Mendes, L.
T. S., Torres, C. A. O., Quast, G. R., Guimaraes, M. M., & da
Silva, L. 2003, AJ, 126, 2971
Visser, A. E., Richer, J. S., & Chandler, C. J. 2002, AJ, 124, 2756
Vorobyov, E. I. & Basu, S. 2006, ApJ, 650, 956
Walawender, J., Bally, J., Kirk, H., & Johnstone, D. 2005, AJ,
130, 1795
Walker, C. K., Lada, C. J., & Young, E. T. 1986, ApJ, 309, L47
Ward-Thompson, D. 1993, MNRAS, 265, 493
Ward-Thompson, D., Andr´e, P., Crutcher, R., Johnstone, D.,
Onishi, T., & Wilson, C. 2007, in Protostars and Planets V,
editors B. Reipurth, D. Jewitt, and K. Keil, p. 33
Whitney, B. A., Wood, K., Bjorkman, J. E., & Cohen, M. 2003,
ApJ, 598, 1079
Wu, J., Dunham, M. M., Evans, N. J. II, Bourke, T. L., & Young,
C. H., 2007, AJ, 133, 1560
Young, C. H. & Evans, N. J., II 2005, ApJ, 627, 293
Young, C. H., Shirley, Y. L., Evans, N. J., II, & Rawlings, J. M.
C. 2003, ApJS, 145, 111
Young, K. E., Enoch, M. L., Evans, N. J., II, Glenn, J., Sargent,
A., Huard, T. L., Aguirre, J., et al. 2006, ApJ, 644, 326
Protostars in Serpens, Perseus, and Ophiuchus
27
TABLE 1
Definition of Classes and Stages
Class/Stage
definition
Class 0
Class I
Class II
Early Class 0
Late Class 0
Early Class I
Late Class I
non-envelope Class I
Stage 0
Stage I
Stage II
Tbol ≤ 70 K
70 K< Tbol ≤ 650 K; 1.1 mm detection (Menv & 0.1M⊙)
650 K< Tbol ≤ 2800 K
Tbol ≤ 50 K
50 K < Tbol ≤ 100 K
100 K < Tbol ≤ 300 K
300 K < Tbol ≤ 650 K
70 K< Tbol ≤ 650 K; no 1.1 mm detection (Menv . 0.1M⊙)
M∗ < Menv
M∗ > Menv; Menv ≥ 0.1M⊙
circumstellar disk; Menv < 0.1M⊙
.
Bolometric temperatures, luminosities, and envelope masses of embedded
protostars in Perseus
TABLE 2
ID
c2d name/position
(SSTc2dJ...)
Tbol
(K)
Lbol
(L⊙)
Class 0
αIR
Menv
M⊙)
Bolocam ID
Other Names
Per-emb 1
Per-emb 2
Per-emb 3
Per-emb 4
Per-emb 5
Per-emb 6
Per-emb 7
Per-emb 8
Per-emb 9
Per-emb 10
Per-emb 11
Per-emb 12
Per-emb 13
Per-emb 14
Per-emb 15
Per-emb 16
Per-emb 17
Per-emb 18
Per-emb 19
Per-emb 20
Per-emb 21
Per-emb 22
Per-emb 23
Per-emb 24
J034356.53+320052.9
J033217.95+304947.6
J032900.52+311200.7
J032839.10+310601.8
J033120.96+304530.2
J033314.40+310710.9
J033032.68+302626.5
J034443.94+320136.1
J032951.82+313906.1
J033316.45+310652.5
J034356.85+320304.6
J032910.50+311331.0
J032912.04+311301.5
J032913.52+311358.0
J032904.05+311446.6
J034350.95+320324.8
J032739.09+301303.0
J032911.25+311831.3
J032923.49+313329.5
J032743.23+301228.8
J032910.68+311820.5
J032522.33+304514.0
J032917.16+312746.4
J032845.30+310542.0
24 (5)
25 (2)
30 (4)
31 (3)
33 (1)
36 (5)
37 (4)
40 (6)
41 (2)
47 (14)
49 (15)
51 (17)
54 (18)
54 (14)
56 (16)
56 (17)
59 (11)
59 (12)
60 (3)
60 (14)
63 (12)
63 (11)
66 (18)
67 (10)
1.5 (0.4)
1.3 (0.2)
0.69 (0.15)
0.22 (0.03)
1.2 (0.2)
0.64 (0.01)
0.15 (0.06)
3.0 (0.5)
0.46 (0.01)
0.27 (0.66)
0.45 (0.85)
4.2 (4.4)
1.6 (2.0)
0.49 (0.32)
0.53 (0.67)
0.38 (0.62)
4.2 (0.1)
2.8 (1.7)
0.36 (0.05)
1.7 (0.01)
2.8 (1.9)
1.7 (1.1)
0.29 (0.33)
0.43 (0.01)
0.5 (0.18)
1.07 (0.11)
2.16 (0.1)
1.68 (0.06)
0.98 (0.06)
2.22 (0.06)
2.08 (0.07)
0.96 (0.07)
3.44 (0.11)
1.73 (0.09)
1.37 (0.09)
2.58 (0.08)
0.55 (0.06)
2.41 (0.13)
1.43 (0.06)
1.51 (0.06)
2.68 (0.06)
1.94 (0.06)
1.51 (0.06)
2.39 (0.06)
1.95 (0.06)
2.34 (0.07)
1.75 (0.06)
1.11 (0.05)
3.02 (0.08) Bolo 103
2.88 (0.08) Bolo 66
0.29 (0.03) Bolo 41
0.36 (0.02) Bolo 30
0.95 (0.04) Bolo 65
Bolo 79a
1.04 (0.1)
0.99 (0.08) Bolo 62
0.63 (0.06) Bolo 116a
0.64 (0.04) Bolo 59
1.57 (0.16) Bolo 79a
1.78 (0.04) Bolo 104
7.75 (0.78) Bolo 48a
3.66 (0.37) Bolo 48a
Bolo 48a
0.5 (0.05)
1.16 (0.12) Bolo 46
1.15 (0.11) Bolo 102a
0.51 (0.03) Bolo 22
2.47 (0.07) Bolo 49b
0.64 (0.05) Bolo 57
0.5 (0.03)
Bolo 23
2.47 (0.07) Bolo 49b
1.41 (0.14) Bolo 5
0.5 (0.05)
Bolo 52
0.19 (0.02) Bolo 33a
HH 211
IRAS 03292+3039
IRAS 03282+3035
IRAS 03267+3128
IC 348-MMS
NGC 1333-IRAS 4A
NGC 1333-IRAS 4B
NGC 1333-IRAS 4C
RNO 15-FIR
L1455-IRS 4
L 1448-IRS2
28
Enoch et al.
TABLE 2 – Continued
ID
c2d name/position
(SSTc2dJ...)
Per-emb 25
Per-emb 26
Per-emb 27c
J032637.46+301528.0
J032538.82+304406.3
J032855.56+311436.6
Per-emb 28
Per-emb 29
Per-emb 30
Per-emb 31
Per-emb 32
Per-emb 33
Per-emb 34
Per-emb 35
Per-emb 36
Per-emb 37
Per-emb 38
Per-emb 39
Per-emb 40
Per-emb 41
Per-emb 42
Per-emb 43
Per-emb 44c
Per-emb 45
Per-emb 46
Per-emb 47
Per-emb 48
Per-emb 49
Per-emb 50d
Per-emb 51
Per-emb 52
Per-emb 53
Per-emb 54d
Per-emb 55
Per-emb 56
Per-emb 57
Per-emb 58
Per-emb 59
Per-emb 60
Per-emb 61
Per-emb 62
Per-emb 63
Per-emb 64
Per-emb 65
Per-emb 66
J034351.02+320307.9
J033317.85+310932.0
J033327.28+310710.2
J032832.55+311105.2
J034402.40+320204.9
J032536.48+304522.3
J033015.12+302349.2
J032837.09+311330.7
J032857.36+311415.7
J032918.27+312320.0
J033229.18+310240.9
J033313.78+312005.2
J033316.66+310755.2
J033320.30+310721.3
J032539.10+304358.0
J034202.16+314802.1
J032903.76+311603.7
J033309.57+310531.2
J032800.40+300801.3
J032834.50+310051.1
J032738.23+301358.8
J032912.94+311814.4
J032907.76+312157.2
J032834.53+310705.5
J032839.72+311731.9
J034741.56+325143.9
J032901.57+312020.7
J034443.33+320131.4
J034705.42+324308.4
J032903.33+312314.6
J032858.44+312217.4
J032835.04+302009.9
J032920.07+312407.5
J034421.33+315932.6
J034412.98+320135.4
J032843.28+311733.0
J033312.85+312124.1
J032856.31+312227.8
J034345.15+320358.6
Tbol
(K)
68 (12)
69 (7)
69 (1)
72 (25)
76 (20)
78 (6)
80 (13)
84 (29)
90 (18)
99 (13)
103 (26)
106 (12)
106 (21)
115 (21)
125 (47)
132 (25)
157 (72)
163 (51)
176 (42)
188 (9)
197 (93)
221 (7)
230 (17)
238 (14)
239 (68)
254 (23)
263 (115)
278 (119)
287 (8)
304 (63)
309 (64)
312 (1)
313 (200)
322 (88)
341 (179)
363 (240)
371 (107)
378 (29)
436 (9)
438 (8)
440 (191)
542 (110)
Lbol
(L⊙)
αIR
Menv
M⊙)
Bolocam ID
Other Names
0.95 (0.02)
4.4 (1.5)
19.0 (0.4)
1.09 (0.05)
2.16 (0.06)
3.03 (0.12)
Bolo 18
0.5 (0.08)
1.87 (0.19) Bolo 10a
2.81 (0.28) Bolo 38a
L 1448-C
NGC 1333-IRAS 2A
Class I
0.28 (0.37)
1.8 (2.5)
1.7 (0.01)
0.16 (0.01)
0.16 (0.31)
4.3 (3.7)
1.6 (0.1)
9.1 (0.3)
5.3 (1.0)
0.07 (0.35)
0.54 (0.01)
0.04 (0.08)
1.5 (1.0)
0.17 (0.36)
0.68 (0.85)
0.07 (0.06)
32.5 (7.1)
0.05 (0.06)
0.3 (0.07)
1.2 (0.1)
0.87 (0.04)
1.1 (0.7)
10.0 (3.0)
0.07 (0.1)
0.16 (0.21)
4.7 (0.9)
6.2 (2.6)
1.8 (0.8)
0.54 (0.09)
0.09 (0.45)
0.63 (0.47)
0.04 (0.06)
0.28 (1.05)
0.24 (0.16)
1.8 (0.4)
1.9 (0.4)
3.2 (0.6)
0.16 (0.16)
0.69 (0.22)
-0.28 (0.06)
3.33 (0.06)
1.93 (0.05)
0.78 (0.05)
1.53 (0.06)
2.62 (0.07)
1.7 (0.05)
2.35 (0.08)
1.6 (0.06)
1.26 (0.06)
0.4 (0.05)
1.44 (0.06)
1.57 (0.05)
0.88 (0.05)
2.36 (0.06)
1.47 (0.05)
1.21 (0.08)
1.13 (0.05)
0.95 (0.05)
0.83 (0.05)
-0.19 (0.05)
1.05 (0.05)
2.18 (0.1)
0.54 (0.05)
0.57 (0.05)
0.78 (0.05)
2.09 (0.08)
0.5 (0.05)
0.36 (0.05)
1.13 (0.05)
0.83 (0.05)
0.15 (0.05)
0.42 (0.05)
0.21 (0.05)
0.37 (0.05)
0.36 (0.05)
0.41 (0.05)
-0.14 (0.05)
-0.22 (0.05)
0.62 (0.06) Bolo 102a
3.84 (0.09) Bolo 80
0.59 (0.06) Bolo 84
0.19 (0.02) Bolo 25
0.63 (0.02) Bolo 106
4.47 (0.45) Bolo 8
0.46 (0.07) Bolo 60
0.44 (0.04) Bolo 29
0.74 (0.07) Bolo 38a
0.44 (0.04) Bolo 54a
0.45 (0.08) Bolo 68
0.41 (0.04) Bolo 78
1.12 (0.11)
1.5 (0.15)
Bolo 81
1.82 (0.18) Bolo 10a
0.07 (0.02)
3.25 (0.32) Bolo 43
0.21 (0.02)
0.17 (0.02)
0.12 (0.02)
0.34 (0.04) Bolo 21
0.77 (0.08) Bolo 49a
1.62 (0.16) Bolo 47
0.24 (0.04) Bolo 28
0.49 (0.05) Bolo 31a
1.67 (0.08) Bolo 122
1.87 (0.19) Bolo 42
0.32 (0.03) Bolo 116a
0.14 (0.02)
0.11 (0.02)
0.71 (0.07) Bolo 40
0.14 (0.02) Bolo 27
0.52 (0.05) Bolo 54a
0.17 (0.02) Bolo 113
0.12 (0.02)
0.25 (0.02) Bolo 31a
0.53 (0.07) Bolo 76
0.25 (0.02)
0.24 (0.02)
B1-c
L 1448-N
IRAS 03271+3013
NGC 1333-IRAS 1
B1-a
B1-b
SVS 13A
IRAS 03254+3050
L1455 FIR2
ASR 30
B5-IRS1
NGC 1333-IRAS 6
IRAS 03415+3152
IRAS 03439+3233
h NOTE – The envelope mass Menv is calculated from the 1.1 mm flux assuming a dust temperature TD = 15 K. Bolocam IDs are
from Papers I–III, and indicate that the protostellar source is within 1.0 core FWHM of a given Bolocam core. If no ID is listed, the
source was not originally identified as a distinct 1.1 mm core, but there is detectable (& 3σ) 1.1 mm emission at the position of the source.
Uncertianties for Tbol, Lbol, and Menv are given in parentheses; uncertainties for Tbol and Lbol are taken to be the difference of values
calculated by the midpoint and prismoidal methods (see §A). Uncertainties in Menv are from the photometric errors in the total flux only.
Absolute uncertainties in the mass may be a factor of 2 or more, from uncertainties in d, κ1mm and TD, but relative values should be much
more accurate.
a More than one protostar is associated with this Bolocam core or source is in a crowded region; mass is based on flux measured in a
30 − 40′′ diameter aperature.
b More than one protostar is associated with this Bolocam core; mass is based on dividing Bolocam core flux equally (source separation
is less than 30′′).
60µm fluxes.
c Saturated 24µm flux has been replaced with IRAS 25µm flux for the calculation of Lbol and Tbol.
d Saturated at either 24µm or 70µm so Lbol is likely an underestimate, but the IRAS beam is too confused to measure reliable 25µm or
Protostars in Serpens, Perseus, and Ophiuchus
29
Bolometric temperatures, luminosities, and envelope masses of cold protostars in Serpens
TABLE 3
ID
c2d name/position
(SSTc2dJ...)
Tbol
(K)
Lbol
(L⊙)
αIR
Menv
(M⊙)
Bolocam ID
Other Names
Ser-emb 1
Ser-emb 2
Ser-emb 3
Ser-emb 4
Ser-emb 5
Ser-emb 6
Ser-emb 7
Ser-emb 8b
Ser-emb 9
Ser-emb 10c,d
Ser-emb 11
Ser-emb 12b
Ser-emb 13
Ser-emb 14
Ser-emb 15
Ser-emb 16d
Ser-emb 17
Ser-emb 18
Ser-emb 19
Ser-emb 20b
Ser-emb 21
Ser-emb 22
Ser-emb 23
Ser-emb 24
Ser-emb 25
Ser-emb 26
Ser-emb 27b
Ser-emb 28d
Ser-emb 29
Ser-emb 30
Ser-emb 31
Ser-emb 32
Ser-emb 33e
Ser-emb 34
J182909.24+003132.3
J182952.44+003611.7
J182854.84+002952.5
J183000.72+011301.4
J182854.84+001832.6
J182949.56+011521.9
J182854.12+002929.9
J182948.12+011644.9
J182855.92+002944.7
J182845.12+005203.5
J182906.72+003034.3
J182952.08+011547.8
J182902.04+003120.6
J183005.40+004104.5
J182954.24+003601.3
J182844.76+005125.7
J182906.36+003043.2
J182952.80+011456.0
J183000.00+011311.6
J182949.20+011619.8
J182951.00+011640.6
J182957.48+011300.5
J182957.84+011251.4
J183000.00+011158.9
J182851.24+001927.3
J182958.92+011426.2
J182956.76+011446.5
J182844.04+005337.9
J183002.88+011228.2
J182957.84+011405.7
J182931.92+011842.9
J183005.76+003931.7
J182916.08+001822.7
J182902.76+003009.5
Class 0
39 (2)
42 (10)
51 (12)
54 (16)
56 (12)
56 (12)
58 (13)
58 (16)
66 (21)
1.6 (0.1)
1.0 (0.1)
2.6 (0.1)
1.2 (1.6)
0.16 (0.1)
11.0 (6.0)
3.1 (0.1)
2.1 (2.4)
0.34 (0.63)
Class I
75 (14)
77 (12)
85 (9)
86 (31)
100 (32)
101 (43)
110 (26)
117 (21)
120 (15)
129 (29)
131 (3)
141 (37)
157 (39)
190 (23)
249 (64)
250 (106)
303 (139)
322 (25)
360 (244)
374 (101)
560 (85)
430 (15)
431 (207)
437 (10)
631 (410)
1.09 (0.05)
1.9 (1.2)
2.6 (1.2)
0.05 (0.09)
0.08 (0.1)
0.17 (0.27)
0.04 (0.6)
1.5 (1.3)
1.2 (0.8)
2.0 (1.9)
4.4 (2.1)
1.7 (1.5)
2.6 (2.2)
3.1 (1.5)
1.8 (1.2)
0.06 (0.07)
0.62 (0.73)
5.8 (1.7)
0.19 (1.5)
2.2 (1.3)
13.8 (4.2)
5.3 (1.0)
0.05 (0.07)
1.7 (0.2)
0.12 (0.28)
2.27 (0.08)
0.77 (0.05)
1.91 (0.06)
1.68 (0.08)
0.9 (0.06)
2.65 (0.07)
1.36 (0.06)
1.37 (0.06)
1.89 (0.06)
1.33 (0.06)
1.66 (0.06)
1.54 (0.06)
0.24 (0.06)
1.3 (0.07)
-0.18 (0.06)
1.07 (0.05)
1.7 (0.06)
1.49 (0.05)
2.57 (0.06)
3.8 (0.11)
1.04 (0.05)
1.01 (0.05)
0.67 (0.05)
1.15 (0.05)
0.45 (0.05)
0.46 (0.05)
0.3 (0.05)
0.45 (0.05)
0.21 (0.05)
0.28 (0.09)
0.26 (0.05)
-0.39 (0.05)
-0.07 (0.05)
-0.14 (0.05)
1.16 (0.02) Bolo 15
0.54 (0.03) Bolo 24a
1.54 (0.15) Bolo 8a
2.56 (0.03) Bolo 28a
0.24 (0.02) Bolo 7a
7.98 (0.07) Bolo 23a
1.67 (0.17) Bolo 8a
3.72 (0.37) Bolo 22a
1.19 (0.12) Bolo 8a
0.29 (0.03) Bolo 3a
1.39 (0.14) Bolo 14a
1.23 (0.12) Bolo 23a
0.17 (0.02)
0.19 (0.02)
0.48 (0.05) Bolo 24a
0.24 (0.02) Bolo 3a
1.38 (0.14) Bolo 14a
0.74 (0.07) Bolo 23a
2.56 (0.03) Bolo 28a
1.53 (0.15) Bolo 22a
1.84 (0.18) Bolo 22a
2.35 (0.24) Bolo 25a
1.06 (0.11) Bolo 25a
1.14 (0.11) Bolo 29
0.16 (0.02) Bolo 7a
Bolo 26a
2.0 (0.2)
0.69 (0.07)
0.26 (0.03) Bolo 2
1.32 (0.13) Bolo 28a
Bolo 26a
1.9 (0.19)
2.12 (0.07) Bolo 20
0.2 (0.02)
Bolo 32
0.54 (0.04) Bolo 17
0.22 (0.02) Bolo 13
FIRS1
S68N
IRAS 1862+0050
SMM 10 IR
IRAS 18274+0112
IRAS 18267+0016
Note. - Columns are as in Table 2, and uncertainties for Tbol, Lbol, and Menv are given in parentheses.
a More than one protostar is associated with this Bolocam core or source is in a crowded region; mass is based on flux measured in a
30 − 40′′ diameter aperature.
b Saturated at either 24µm or 70µm so Lbol is likely an underestimate, but the IRAS beam is too confused to measure reliable 25µm or
60µm fluxes.
c Missing 70µm flux has been replaced with IRAS 60µm flux for the calculation of Lbol and Tbol.
d Source lies outside the 70µm map.
e Saturated 24µm flux has been replaced with IRAS 25µm flux for the calculation of Lbol and Tbol.
30
Enoch et al.
TABLE 4
Bolometric temperatures, luminosities, and envelope masses of cold protostars in Ophiuchus
ID
c2d name/position
(SSTc2dJ...)
Tbol
(K)
Lbol
(L⊙)
αIR
Menv
(M⊙)
Bolocam ID
Other Names
Oph-emb 1
Oph-emb 2a
Oph-emb 3
J162821.72-243623.4
J163222.56-242831.8
J162625.80-242428.8
35 (2)
54 (8)
57 (10)
0.25 (0.01)
6.9 (1.2)
0.41 (0.15)
1.23 (0.06)
5.03 (0.2)
1.65 (0.07)
0.51 (0.04) Bolo 26
1.16 (0.02) Bolo 36
0.24 (0.02) Bolo 9
IRAS 16253-2429
IRAS 16293-2422B
VLA 1623 (?)
Class 0
Oph-emb 4
Oph-emb 5
Oph-emb 6
Oph-emb 7
Oph-emb 8c,d
Oph-emb 9
Oph-emb 10
Oph-emb 11
Oph-emb 12
Oph-emb 13c
Oph-emb 14c
Oph-emb 15
Oph-emb 16c,d
Oph-emb 17
Oph-emb 18
Oph-emb 19
Oph-emb 20
Oph-emb 21c
Oph-emb 22
Oph-emb 23
Oph-emb 24
Oph-emb 25
Oph-emb 26
Oph-emb 27
Oph-emb 28
J163136.84-240419.9
J162721.96-242727.7
J162705.40-243629.5
J163151.96-245725.9
J162621.48-242304.2
J162625.44-242301.3
J163200.96-245642.7
J162717.64-242856.2
J162724.48-244103.1
J162728.08-243933.4
J162727.00-244050.5
J163152.32-245536.1
J162709.36-243718.4
J163135.76-240129.2
J162857.72-244054.8
J162728.44-242720.8
J162706.84-243815.0
J162702.16-243727.1
J162640.56-242714.4
J162648.48-242838.6
J162737.08-244237.8
J163143.68-245524.6
J162730.24-242743.2
J162739.96-244314.8
J162721.60-244143.0
Class I
77 (25)
87 (28)
106 (14)
124 (39)
133 (34)
135 (49)
145 (38)
190 (1)
191 (30)
193 (31)
211 (7)
240 (42)
257 (60)
290 (18)
304 (96)
309 (85)
310 (23)
331 (2)
372 (160)
428 (84)
438 (87)
496 (68)
528 (99)
551 (54)
595 (57)
0.18 (0.05)
0.08 (0.08)
0.15 (0.09)
0.1 (0.13)
18.3 (1.4)
0.1 (0.3)
2.7 (0.3)
0.52 (0.06)
0.33 (0.16)
7.2 (0.5)
3.8 (0.7)
0.13 (0.07)
17.9 (0.6)
1.5 (0.1)
0.03 (0.03)
0.47 (0.28)
0.6 (0.16)
5.2 (1.3)
0.06 (0.09)
0.12 (0.05)
0.13 (0.06)
0.28 (0.09)
0.93 (0.27)
0.74 (0.15)
1.2 (0.2)
-0.27 (0.05)
-0.05 (0.05)
1.27 (0.05)
0.82 (0.05)
1.46 (0.09)
0.87 (0.05)
1.39 (0.07)
0.25 (0.05)
1.01 (0.05)
2.29 (0.11)
1.17 (0.08)
1.07 (0.05)
1.69 (0.15)
0.14 (0.05)
0.67 (0.05)
-0.03 (0.05)
0.61 (0.05)
1.53 (0.11)
0.45 (0.05)
0.02 (0.05)
0.13 (0.05)
0.23 (0.05)
-0.12 (0.05)
-0.15 (0.05)
-0.03 (0.05)
0.03 (0.01)
0.12 (0.01) Bolo 17
0.09 (0.01) Bolo 14b
0.05 (0.01) Bolo 33
0.22 (0.02) Bolo 5
0.53 (0.05) Bolo 8
0.05 (0.01) Bolo 35
0.07 (0.01)
0.12 (0.01) Bolo 18b
0.11 (0.01)
0.36 (0.04) Bolo 18b
0.08 (0.01)
0.07 (0.01) Bolo 14b
0.14 (0.02) Bolo 30
0.04 (0.01)
0.29 (0.03) Bolo 20b
0.05 (0.01) Bolo 13
0.08 (0.01)
0.06 (0.01)
0.03 (0.01)
0.04 (0.01)
0.04 (0.01)
0.19 (0.02) Bolo 20b
0.04 (0.01)
0.04 (0.01)
LFAM26/GY197
GSS 30
GY 30
L1689S/IRS67
CRBR85
IRS 44/GY269 )
IRS 43/GY265
EL 29/GY214
IRAS 16295-2355
IRS45/GY273
WL17/GY205
WL16/GY182
GY91/CRBR42
WL2/GY128
GY301
IRS47/GY279
IRS51/GY315
IRS42/GY252
Note. - Columns are as in Table 2, and uncertainties for Tbol, Lbol, and Menv are given in parentheses.
a Oph-emb 2 is saturated at 160µm; when the 100µm IRAS flux is included the resulting Lbol is 16 ± 6L⊙, consistent with the luminosity
calculated by Walker et al. (1986) when scaled to a distance of 125 pc.
b More than one protostar is associated with this Bolocam core or source is in a crowded region; mass is based on flux measured in a
30 − 40′′ diameter aperature.
c Saturated 24µm flux has been replaced with IRAS 25µm flux for the calculation of Lbol and Tbol.
d Saturated 70µm flux has been replaced with IRAS 60µm flux for the calculation of Lbol and Tbol.
Numbers of Protostars by Class and the Class 0 lifetime
TABLE 5
Cloud
NClass0
NClassI
NClass0/NClassI
Perseus
Serpens
Ophiuchus
27 (27)
9 (13)
3 (4)
Three cloud sample
39 (44)
39 (41)
25 (23)
25 (24)
89 (88)
0.70 (0.66)
0.4 (0.6)
0.12 (0.17)
0.44 (0.50)
τClass0
(yr)
2.2 ± 0.4 × 105
1.7+0.5
−0.9 × 105
0.7+0.4
−0.5 × 105
1.72 ± 0.25 × 105
Note. - Numbers of Class 0 and Class I sources are based on Tbol classifications (Table 1)
including the requirement that Class I sources be detected at 1.1 mm. Numbers in parentheses
indicate how the results change if we utilize a different method for calculating Tbol (prismoidal
versus midpoint integration, see §A). The Class 0 lifetime, τClass0, assumes that the entire
embedded phase lasts for 5.4×105 years (τemb = τClass0 +τClassI = 5.4×105 yr; Evans et al.
2008), and NClass0/(NClass0 +NClassI ) = τClass0/τemb. The lifetime is calculated using the
average of the two NClass0/NClassI ratios given, and the uncertainty is from the difference
between them added in quadrature with the √N uncertainties from counting statistic.
Protostars in Serpens, Perseus, and Ophiuchus
31
Class II sources detected at 1.1 mm in Perseus, Serpens, and Ophiuchus
TABLE 6
c2d name/position
(SSTc2dJ...)
Tbol
(K)
Lbol
(L⊙)
αIR
a
M1mm
(M⊙)
Bolocam IDb
other names
J032747.66+301204.6
J034400.00+320154.1
J032900.00+312146.8
J034109.13+314438.0
J032856.64+311835.6
J034548.27+322411.8
J032917.66+312245.1
662 (15)
700 (17)
856 (26)
873 (30)
993 (22)
1096 (23)
1297 (22)
J182900.00+003003.1
J183006.12+004233.9
J182900.96+002931.6
J182901.32+002933.0
J182901.68+002954.6
J182901.68+002946.5
J163151.96-245615.7
J162658.56-244531.6
J162624.00-241613.4
J163945.36-240203.8
J162623.28-242100.7
806 (5)
871 (2)
892 (2)
892 (6)
1158 (3)
1212 (5)
690 (71)
824 (14)
958 (35)
970 (30)
1080 (100)
2.80 (0.05)
1.60 (0.03)
0.33 (0.01)
1.70 (0.06)
0.42 (0.01)
5.00 (0.09)
1.50 (0.02)
0.19 (0.01)
3.60 (0.01)
1.20 (0.01)
0.36 (0.01)
1.80 (0.01)
0.70 (0.01)
1.2 (0.3)
1.3 (0.1)
1.9 (0.2)
0.84 (0.09)
2.0 (0.4)
Perseus
-0.09 (0.05)
-0.34 (0.05)
-0.83 (0.05)
-0.72 (0.05)
-0.78 (0.05)
-0.72 (0.05)
-0.77 (0.05)
Serpens
-0.61 (0.05)
-0.42 (0.05)
-0.42 (0.05)
-0.54 (0.05)
-1.04 (0.05)
-0.65 (0.05)
Ophiuchus
-0.17 (0.05)
-0.45 (0.05)
-0.71 (0.05)
-0.73 (0.05)
-0.62 (0.05)
0.82 (0.06) Per-Bolo 24
0.40 (0.04) Per-Bolo 106∗
0.90 (0.09) Per-Bolo 40∗
0.16 (0.02) Per-Bolo 90
0.10 (0.02) Per-Bolo 39∗
0.10 (0.02) Per-Bolo 120
0.19 (0.02) Per-Bolo 54∗
RNO 15c (1)
LkHα 353 (2)
IRAS 03380+3135
ASR 120 (3)
LkHα 330 (2)
LkHα 270 (2)
0.35 (0.03)
0.17 (0.01)
0.23 (0.02)
0.23 (0.02)
0.23 (0.02)
0.25 (0.02)
Ser-Bolo 13∗
Ser-Bolo 33
Ser-Bolo 13∗
Ser-Bolo 13∗
Ser-Bolo 13∗
Ser-Bolo 13∗
∗
0.05 (0.01)
0.03 (0.01)
0.04 (0.01)
0.03 (0.01)
0.15 (0.01) Oph-Bolo 7∗
CoKu Ser G3 (4)
CoKu Ser G4 (4)
GY 186 (5)
GY 20 (5)
Note. - References: (2) Herbig & Kameswara 1972; (3) Aspin et al. 1994; (4) Cohen & Kuhi 1979; (5) Greene & Young
1992.
a The 1.1 mm mass (M1mm) is calculated for an optically thin disk and a dust temperature of TD = 15 K. Our assumption
for the dust opacity may be incorrect for typical Class II disks, however, where grain growth is a possibility (e.g., Natta et al.
2007).
b Bolocam IDs with "*"s indicate that the source is in a confused region at 1.1 mm, and the association of the Spitzer
object with the given 1.1 mm emission is not secure.
c Typically classified as a Class I source, RNO 15 has a bolometric temperature just above the Class I cutoff in our sample.
The spectral index of RNO 15 is αIR = −0.09, placing it in the "flat spectrum" category, or close to the Class I boundary.
Effect of Extinction Corrections on Class Statistics
TABLE 7
Cloud
NClass0
NClassI
NClass0/NClassI
Perseus
Serpens
Ophiuchus
Three cloud sample
21
4
3
29
42
26
20
88
0.5
0.15
0.15
0.33
τClass0
(yr)
1.5 × 105
0.6 × 105
0.6 × 105
1.1 × 105
Note. - Same as Table 5 but for classifications based on T ′
bol, the extinction-
corrected bolometric temperature. In addition, an embedded phase lifetime of 4.4 ×
105 yr is assumed, derived based on extinction-corrected classifications (Evans et al.
2008).
TABLE 8
Lifetimes for narrow Tbol bins
Phase
Tbol range
N(Per) N(Ser) N(Oph)
(K)
Early Class 0
T < 50
Late Class 0
50 < T < 100
Early Class I
100 < T < 300
Late Class I
300 < T < 650
11
23
19
13
2
11
12
9
1
4
12
11
N(total)/
N(emb,tot)
14/128 = 0.11
38/128 = 0.30
43/128 = 0.34
33/128 = 0.26
Mean Tbol
Lifetime
(K)
40
70
180
400
(yr)
0.6 × 105
1.6 × 105
1.8 × 105
1.4 × 105
Note. - For each Tbol bin, N(total) is the total number of sources in all three clouds; N(emb,tot)
is the total number of embedded protostars in the three cloud sample (128). Only sources detected at
1.1 mm are included. The mean Tbol of protostars in a given Tbol range is also given. Lifetimes are
calculated from N(total)/N(emb, tot)×5.4×105 yr. Cloud-to-cloud variations are typically 0.2−0.3×105
yr.
32
Enoch et al.
APPENDIX
CALCULATING THE BOLOMETRIC LUMINOSITY AND TEMPERATURE
Determination of the bolometric temperature and luminosity of any given source can depend strongly on the method
used to approximate the integrations over frequency in equations 1–3. As Sν is sampled at a finite number of frequencies
(usually 6 to 10 here), the SED must be interpolated over the intermediate frequencies. Here we use two different
methods; the first (midpoint) method utilizes a simple linear interpolation for the midpoint flux, while the second
(prismoidal) method calculates a color temperature for each pair of flux points, and uses a modified blackbody based
on that color temperature to estimate the midpoint flux density.
In both cases the SED is extrapolated from the
longest wavelength flux using Sν ∝ ν2, and flux upper limits are removed from the fit (i.e., we interpolate over them).
Examples of the midpoint and prismoidal interpolations are shown in Figure 16 for two sources with Class 0-like (left)
and Class I-like (right) SEDs.
Fig. 16.- Examples of the interpolated midpoints used for the integration of the SED in the midpoint (solid line) and prismoidal (dashed
line) methods. Observed data points are overlaid as diamonds. A typical Class 0 source is shown on the left, and a Class I source on
the right; both sources shown are lacking data at 160 µm. After the midpoint is calculated, the integration for both methods is done in
linear νSν space. The prismoidal method provides a good approximation of the dust emission peak for cold sources (left), but probably
overestimates the long-wavelength flux for warmer, flatter SED sources (right), resulting in a bias toward lower Tbol.
Figure 17 compares the results of using the midpoint and prismoidal methods to calculate Lbol and Tbol for the
embedded protostar sample in Perseus. The midpoint and prismoidal Tbol values agree fairly well, usually to within
20%, with no strong systematic bias. There is some tendency for the prismoidal Tbol to be lower for colder sources.
On the other hand, Lbol calculated by the prismoidal method is generally higher than the midpoint Lbol.
To further test this behavior, we plot in Figure 18 the fractional errors due solely to finite sampling, for Lbol and
Tbol calculated from input blackbody spectra. Blackbody sources with input Tbol from 5–5000 K are sampled at the
observed wavelengths (λ =1.25, 1.65, 2.17, 3.6, 4.5, 5.8, 8, 24, 70, 160, 1100 µm), and the bolometric temperature and
luminosity calculated by the midpoint (left) and prismoidal (right) methods. Both Lbol and Tbol have typical errors
of approximately 20%, and the measured Tbol is consistently over-estimated by approximately 20%. The midpoint
Tbol is more variable but somewhat more accurate than the prismoidal Tbol, while Lbol is somewhat less accurate. Of
course, the blackbody spectra tested here are not necessarily representative of the more complicated observed SEDs.
Based on Figures 17–18 and on examinations of the midpoint and prismoidal fits to observed SEDs, from which we
find that the prismoidal method often provides a poor fit to "flatter" SEDs at long wavelengths (Figure 16), we use
the midpoint method throughout this paper. Differences between the midpoint and prismoidal methods are often used
as a measure of uncertainty.
In general, measured 160 µm fluxes are uncertain by up to a factor of 2, due to unquantified saturation effects
Protostars in Serpens, Perseus, and Ophiuchus
33
Fig. 17.- Comparison of the midpoint and prismoidal methods for calculating Lbol and Tbol, for sources in Perseus. Left: Midpoint and
prismoidal Tbol (black asterisks) and Lbol (triangles) values are plotted for all candidate embedded protostars in Perseus. Right: Fractional
difference between the midpoint and prismoidal methods, plotted as (midpoint-prismoidal)/prismoidal. The two methods generally agree
to within 20% for Tbol but Lbol calculated by the prismoidal method is higher than the midpoint Lbol by as much as 100%. There is some
tendency for the prismoidal Tbol to be lower for colder sources.
and calibration uncertainties, and in many cases we are unable to measure a 160 µm flux at all. In addition, the
"point-source" 160 µm fluxes of extended sources will likely be underestimated, because they are measured with a
PSF fit rather than aperture photometry. Missing or severely underestimated fluxes at 160 µm can significantly affect
both Lbol and Tbol for cold sources that peak near 100 µm. As we are primarily interested in the coldest sources, we
attempt here to quantify the effects of missing or underestimated 160 µm flux. One way of approaching this problem
is to examine sources for which we have additional information near the peak of the SED, e.g., from IRAS or SHARC
350 µm observations (Wu et al. 2007).
We first look at two example sources, one cold (IRAS 03282+3039; Tbol ∼ 33 K), and one warmer (Tbol ∼ 163 K).
IRAS 03282+3039 is isolated, so we can use the IRAS fluxes without worrying about confusion due to the large IRAS
beam. We calculate the "true" Lbol and Tbol by including the IRAS 60 and 100 µm fluxes, a SHARC II 350 µm
flux, and a MIPS 160 µm flux measured in a large aperture, which is almost two times higher than the PSF-fit flux.
Using the underestimated point source 160 µm flux, and no longer including IRAS or SHARC II data, causes an
underestimate of Lbol by 35% and an overestimate of Tbol by 15% compared to the "true" values. Omitting the 160 µm
point altogether, as would be appropriate for saturated sources, results in an underestimate of Lbol by 7% and an
overestimate of Tbol by 9%.
Thus we conclude that while the 160 µm point is important for characterization of the SED of embedded objects,
our integration method can interpolate over a missing 160 µm flux to estimate Lbol and Tbol to within 20% for cold
sources. Severely underestimated 160 µm flux densities will cause larger errors of up to 50%. Using a similar procedure
for the warmer source (Tbol = 163 K), we find that omitting the 160 µm flux results in errors in the opposite sense
compared to the colder source: an overestimate of Lbol by 28% and an underestimate of Tbol by 18%.
For a more general result, we compare Tbol calculated with the 160 µm flux included in the SED to the value found
by omitting the 160 µm point, for all sources in Perseus with a measured 160 µm flux (Figure 19, left panel). For this
plot, we have adopted classifications from Chen et al. (1995), as discussed in §4. Although Tbol calculated without the
160 µm flux is always overestimated compared to Tbol calculated with the 160 µm flux, it is never a large enough effect
to shift the source classification of objects with Tbol & 100 K (e.g., from Class I to Class II). This is not the case for
Class 0 sources, however, as can be seen in the right panel of Figure 19, where the scale has been adjusted to highlight
the lowest Tbol values. Five sources that have Tbol < 70 K (Class 0) when including the 160 µm flux are shifted to
650 < Tbol > 70 K (Class I) when the 160 µm flux is omitted. If a published 350 µm flux is available (diamonds), the
errors in Tbol resulting from excluding the 160 µm flux are almost completely eliminated.
Based on Figures 17–19 and the above examples, we estimate overall uncertainties for measured Lbol and Tbol values
of 20 − 50%, depending on whether or not a 160 µm flux is available. If we are unable to measure a 160 µm flux, Tbol
34
Enoch et al.
Fig. 18.- Characterization of sampling errors for Tbol and Lbol calculated using the midpoint (left) and prismoidal (right) methods. The
fractional difference between the input (Tbol,Lbol) of blackbody spectra and the measured values when SEDs are sampled at the observed
wavelengths is plotted as a function of the input Tbol. Both Lbol and Tbol have typical errors of approximately 20%, and the measured
Tbol is consistently over-estimated by approximately 20%. Source SEDs, which are considerably more complex than the blackbodies tested
here, may behave differently.
Fig. 19.- Characterization of the systematic errors introduced into the measured Tbol when a 160 µm flux is not available. Left: For all
sources in Perseus with a reliable 160 µm flux, we calculate Tbol with and without the 160 µm point (asterisks). Tbol is always overestimated
when the 160 µm flux is missing, but it does not change the classification of Class I sources for these data. Right: Similar, but with the
scale adjusted to highlight lower Tbol sources. Not having a 160 µm point does change the classification of approximately half of the Class 0
(Tbol < 70 K) sources. Errors are substantially reduced when a flux at 350 µm is available (diamonds).
will almost certainly be an overestimate for very cold sources, which may affect our classification of the most deeply
embedded protostars.
|
astro-ph/0412006 | 1 | 0412 | 2004-11-30T23:22:12 | X-ray and Radio Monitoring of GX 339-4 and Cyg X-1 | [
"astro-ph"
] | Previous work by Motch et al. (1985) suggested that in the low/hard state of GX339-4, the soft X-ray power-law extrapolated backward in energy agrees with the IR flux level. Corbel and Fender (2002) later showed that the typical hard state radio power-law extrapolated forward in energy meets the backward extrapolated X-ray power-law at an IR spectral break, which was explicitly observed twice in GX339-4. This has been cited as further evidence that jet synchrotron radiation might make a significant contribution to the observed X-rays in the hard state. We explore this hypothesis with a series of simultaneous radio/X-ray hard state observations of GX339-4. We fit these spectra with a simple, but remarkably successful, doubly broken power-law model that indeed requires a spectral break in the IR. For most of these observations, the break position as a function of X-ray flux agrees with the jet model predictions. We then examine the radio flux/X-ray flux correlation in Cyg X-1 through the use of 15 GHz radio data, obtained with the Ryle radio telescope, and Rossi X-ray Timing Explorer data, from the All Sky Monitor and pointed observations. We find evidence of `parallel tracks' in the radio/X-ray correlation which are associated with `failed transitions' to, or the beginning of a transition to, the soft state. We also find that for Cyg X-1 the radio flux is more fundamentally correlated with the hard, rather than the soft, X-ray flux. | astro-ph | astro-ph |
X-ray and Radio Monitoring of GX 339−4 and Cyg X-1
Michael Nowak
Massachusetts Institute of Technology - Chandra X-ray Science Center
Abstract. Previous work by Motch et al. (1985) suggested that in the low/hard
state of GX 339−4, the soft X-ray power-law extrapolated backward in energy agrees
with the IR flux level. Corbel and Fender (2002) later showed that the typical
hard state radio power-law extrapolated forward in energy meets the backward
extrapolated X-ray power-law at an IR spectral break, which was explicitly observed
twice in GX 339−4. This has been cited as further evidence that jet synchrotron
radiation might make a significant contribution to the observed X-rays in the hard
state. We explore this hypothesis with a series of simultaneous radio/X-ray hard
state observations of GX 339−4. We fit these spectra with a simple, but remarkably
successful, doubly broken power-law model that indeed requires a spectral break in
the IR. For most of these observations, the break position as a function of X-ray flux
agrees with the jet model predictions. We then examine the radio flux/X-ray flux
correlation in Cyg X-1 through the use of 15 GHz radio data, obtained with the Ryle
radio telescope, and Rossi X-ray Timing Explorer data, from the All Sky Monitor
and pointed observations. We find evidence of 'parallel tracks' in the radio/X-ray
correlation which are associated with 'failed transitions' to, or the beginning of a
transition to, the soft state. We also find that for Cyg X-1 the radio flux is more
fundamentally correlated with the hard, rather than the soft, X-ray flux.
1. Introduction
Both Cyg X-1 and GX 339−4 in their spectrally hard, radio-loud states
have served as canonical examples of the so-called 'low state' (or 'hard
state') of galactic black hole candidates [16, 14]. In this state the X-ray
spectrum is reasonably well-approximated by a power-law with photon
spectral index of Γ ≈ 1.7, with the power-law being exponentially cutoff
at high energies (≈ 100 keV). Such spectra have been attributed to
Comptonization of soft photons from an accretion disk by a hot corona;
however, it recently has been hypothesized that the X-ray spectra of
hard state sources might instead be due to synchrotron and synchrotron
self-Compton (SSC) radiation from a mildly relativistic jet [10, 12]. Jet
models have been prompted in part by multi-wavelength (radio, optical,
X-ray) observations of hard state systems.
In hard states of GX 339−4, the 3 -- 9 keV X-ray flux (in units of
10−10 erg cm−2 s−1) is related to the 8.6 GHz radio flux (in mJy) by
Fx ≈ 0.46F 1.42
[2]. This correlation was seen to hold over several
decades in X-ray flux, and also to hold for two hard state epochs
that were separated by a prolonged, intervening soft state outburst.
r
c(cid:13) 2018 Kluwer Academic Publishers. Printed in the Netherlands.
mnowak_talk.tex; 18/11/2018; 5:41; p.1
2
Michael Nowak
It further has been suggested that the Fx ∝ F 1.4
correlation is a
universal property of the low/hard state of black hole binaries [5].
This specific power-law dependence of the radio flux upon the X-ray
flux naturally arises in synchrotron jet models [3, 2, 12, 7], where the
optically thin synchrotron spectrum, occurring above an IR spectral
break, is presumed to continue all the way through the X-ray.
r
Interestingly, nearly 20 years ago Motch et al. (1985) [13] noted
that for a set of simultaneous IR, optical, and X-ray observations of
the GX 339−4 hard state, the extrapolation of the X-ray power-law
to low energy agreed with the overall flux level of the optical/IR data.
Corbel and Fender (2002) [1] reanalyzed these observations (which did
not include simultaneous radio data), as well as a set of (not strictly
simultaneous) radio/IR/X-ray observations from the 1997 GX 339−4
hard state. They showed that the low energy extrapolation of the X-ray
power-laws, and the high energy extrapolation of the radio power-law,
coincided with a spectral break in the IR.
2. Observations of GX 339−4
We consider a set of ten simultaneous radio/X-ray observations of
GX 339−4, eight of which come from the 1997 or 1999 hard state
[17, 14] and and two of which come from the 2002 hard state [8]. All X-
ray observations were performed with the Rossi X-ray Timing Explorer
(RXTE). Note that five of these observations are further labeled A --
E, as we single these out for special discussion. A and B occurred
immediately after the 1999 soft-to-hard state transition [14] and have
optically thin radio spectra (αr < 0). C has a very 'inverted' radio spec-
trum (see below). D has only a single radio point, and hence we cannot
extrapolate its radio power-law without making further assumptions. E
has the brightest X-ray flux in our sample, and is one of the brightest
hard X-ray states observed in GX 339−4 to date.
To analyze the X-ray spectra of these observations, RXTE response
matrices were created using the software tools available in HEASOFT
5.3, which we find yield extremely good agreement between the Propor-
tional Counter Array (PCA) and High Energy X-ray Timing Explorer
(HEXTE) when fitting power-law models to the Crab pulsar plus neb-
ula system. This is true for both the power-law normalization and slope,
both of which must be determined very accurately when extrapolating
over large energy ranges.
The radio data for observation E were obtained with the Australia
Telescope Compact Array (ATCA) at 4.8 GHz and 8.6 GHz. The radio
mnowak_talk.tex; 18/11/2018; 5:41; p.2
Monitoring of GX 339−4 and Cyg X-1
3
Obs. E (40031−03−01)
Soft−to−Hard
X−ray Break
Figure 1. Unfolded spectra of an X-ray spectrum of GX 339−4 fit with an absorbed,
exponentially cutoff, broken power-law and a gaussian line. Residuals are from the
proper forward folded model fit.
data for observation D were also obtained with ATCA, but only at 5
GHz. All other radio data can be found in [14].
3. A Rant on the Nature of Evil
The observations were analyzed with the Interactive Spectral In-
terpretation System (ISIS) [9]. For our purposes, there are several
major reasons for our use of ISIS. Data input without a response
matrix (i.e., the radio data) are automatically presumed to have an
associated diagonal response with one cm2 effective area and one second
integration time. We convert the radio data from mJy to photon rate
in narrow bands around the observation frequencies, and use this as
input for the simultaneous radio/X-ray fits.
The other major reason for using ISIS is that it treats 'unfolded
spectra' (shown in Fig. 1) in a model-independent manner. The un-
folded spectrum in an energy bin denoted by h is defined by: Funfold(h) =
([C(h) − B(h)]/∆t)/(R R(h, E)A(E)dE), where C(h) is the total de-
tected counts, B(h) is the background counts, ∆t is the integrated
observation time, R(h, E) is the unit normalized response matrix de-
scribing the probability that a photon of energy E is detected in bin h,
mnowak_talk.tex; 18/11/2018; 5:41; p.3
4
Michael Nowak
Obs. E (40031−03−01)
B
Radio−to−Soft
X−ray Break
E
D
C
2
C1
Figure 2. Left: An unfolded, simultaneous radio/X-ray observation of GX 339−4,
fit with an absorbed, exponentially cutoff, doubly broken power-law and a gaussian
line. Right: Results of broken power-law fits to GX 339−4, showing the location
of the break between the radio and soft X-ray power-law as a function of X-ray
flux. Dashed lines show the approximate integrated X-ray flux and approximate IR
spectral break energy previously observed in GX 339−4 (Corbel & Fender 2002).
Dotted lines are Eb−r ∝ Fx
0.38, Eb−r ∝ Fx
0.74.
and A(E) is the detector effective area at energy E. Contrary to un-
folded spectra produced by XSPEC, this definition produces a spectrum
that is independent of the fitted model. Any unfolded spectrum should
be considered something of a sin; however, ISIS unfolded spectra are
only venial sins, whereas XSPEC unfolded spectra should rightly be
classified as cardinal sins (dictum vel factum vel concupitum contra
legem aeternam). In Fig. 1, however, the plotted residuals are those
obtained from a proper forward-folded fit.
4. Radio-to-X-ray Break Energy Correlations
We obtain surprisingly good fits for nine of the ten radio/X-ray spectra
using the following simple model (using the ISIS/XSPEC model defini-
tions): absorption (the phabs model, with NH fixed to 6 × 1021 cm2)
and a high energy, exponential cutoff (the highecut model) multiplying
a doubly broken power-law (the bkn2pow model, with the first break
being in the far IR to optical regime, and the second break being
constrained to the 9 -- 12 keV regime) plus a gaussian line (with energy
fixed at 6.4 keV). When considering just the X-ray spectra, a singly
broken power-law fits all ten spectra, with better results than any of
the Comptonization models that we have tried.
In Fig. 2 we show the fitted radio-to-X-ray break location as a
function of 3 -- 9 keV integrated flux. We also show in this figure the
approximate integrated 3 -- 9 keV flux and the IR break location for the
mnowak_talk.tex; 18/11/2018; 5:41; p.4
Monitoring of GX 339−4 and Cyg X-1
5
1997 observation discussed by Corbel and Fender (2002) [1]. For our
GX 339−4 observations of comparable 3 -- 9 keV flux, the doubly broken
power-law models do indeed produce a break in the IR. The model
fits presented here have predicted radio-to-X-ray breaks ranging all the
way from the far IR to the blue end of the optical (and into the X-ray,
if one also considers observation B, which has an 'optically thin' radio
spectrum).
The data point labeled C1, with a break in the far IR, has an ex-
tremely 'inverted' radio spectrum (αr = 0.58). This drives the fitted
break to low energies, and hence leads to deviations from the overall
observed trends shown in Fig. 2. Such an inverted spectrum is very
unlikely to be intrinsic to the radio jet, and is most likely a signature
of free-free absorption at low frequencies [4]. If we instead consider
only the highest observed radio frequency (8.6 GHz, which is likely less
affected by free-free absorption), and fix the radio spectral slope at this
point to αr = 0.1, similar to the other observations, we obtain an IR
break frequency (labeled C2) that is consistent with the other inferred
breaks.
To assess the correlation of radio-to-X-ray break energy with inte-
grated X-ray flux, we exclude the data points from observation C (likely
free-free absorbed), observation B (which has an optically thin radio
spectrum), and observation D (which is consistent with the trends if
we assume a radio slope of αr = 0.1). A regression fit to the remaining
six data points suggests that the radio-to-X-ray break energy, in eV,
scales with the 3 -- 9 keV integrated flux as 0.95Fx
0.74±0.05.
Using the scale invariance Ansatz to describe the jet physics [7, 6], we
show elsewhere [15] that the predicted scaling between the integrated
X-ray synchrotron flux and the radio-to-X-ray break frequency where
the jet becomes optically thin to synchrotron self-absorption scales as
2(p+6)/(p+4)/(p+5), where p is the power-law index of the electron
νb ∝ Fx
spectrum, and we have used for the X-ray spectral slope αx = (1 − p)/2
from standard synchrotron theory. For the usual range of −0.65 <
to νb ∝ F 0.38
αx < −0.5 of synchrotron spectra, we obtain νb ∝ F 0.36
.
This prediction is flatter than the observed dependence of extrapolated
break frequency upon X-ray flux. However, if one also excludes the
highest flux point, then the scaling becomes more consistent with the
jet synchrotron prediction, i.e., νb ∝ Fx
0.38±0.16 (Fig. 2).
x
x
5. Radio/X-ray Correlations in Cyg X-1
We now turn to radio/X-ray observations of Cyg X-1 [16, 5]. Again,
these spectra are remarkably well-fit by a simple, exponentially cut-
mnowak_talk.tex; 18/11/2018; 5:41; p.5
6
Michael Nowak
Figure 3. Left: Unfolded spectra of an X-ray spectrum of Cyg X-1 fit with an
absorbed, exponentially cutoff, broken power-law and a gaussian line. Right: Corre-
lation of the soft X-ray spectral slope with the hard minus soft X-ray slope for all
our Cyg X-1 observations.
off broken power-law (Fig. 3). The degree of the break here indicates
that the broken power-law models are not simply mimicking reflection,
but are suggesting two separate continuum components. In fact, the
correlation between the soft X-ray slope and the hard minus soft X-
ray slope is more pronounced than any correlation between reflection
fraction and spectral slope (Fig. 3). We shall elaborate upon these point
further in an upcoming paper (Wilms et al., in prep.).
The associated radio data are 15 GHz observations performed at the
Ryle Telescope, Cambridge (UK) [16]. (These are single channel obser-
vations, so a radio spectral slope cannot be determined.) Most of these
observations have occurred simultaneously with pointed RXTE obser-
vations [16], and nearly all have very good contemporaneous coverage
by the RXTE All Sky Monitor (ASM ).
In Fig. 4 we plot the daily average ASM count rate vs. the daily
average 15 GHz flux. Ranging from approximately 10 -- 50 cps in the
ASM there is a clear log-linear correlation between the radio flux and
the ASM count rate. As for GX 339−4, the radio flux rises more slowly
than the ASM count rate (Fr scales approximately as the 0.8 power of
the ASM count rate). Cyg X-1, however, shows much more scatter in
the amplitude of the correlation than does GX 339−4.
As noted elsewhere [5], there is a sharp roll-over for higher ASM
count rates. However, one can clearly discern on the shoulder of this
roll-over (i.e., the upper right corner of Fig. 4) four 'spokes', consisting
of 2 -- 5 data points each. In these spokes, the radio/X-ray correlation
appears to hold to high count rates. We have confirmed [15] that each
of these times are associated with 'failed transitions' to the soft state
mnowak_talk.tex; 18/11/2018; 5:41; p.6
Monitoring of GX 339−4 and Cyg X-1
7
Figure 4. Left: 15 GHz Ryle radio flux (mJy) vs. Cyg X-1 daily mean ASM count
rate. Right: 20 -- 200 keV flux (units of 10−9 ergs cm s−1) vs. the daily average 15 GHz
Ryle radio flux (mJy) for pointed observations of Cyg X-1.
[16], except for the lowest amplitude of these spokes, which occurs
immediately preceding a prolonged soft state outburst.
In Fig. 4 we also plot the daily average ASM count rate vs. the 20 --
200 keV flux from our pointed RXTE observations taken during the
same 24 hour period [16]. We see that hard X-ray/ASM correlation
traces a similar pattern to the radio/ASM correlation. Indeed, when
we plot the hard X-ray flux vs. the daily average radio flux we obtain
a log-linear relationship, as shown in Fig. 4. In Cyg X-1, the radio flux
density appears fundamentally to be tied to the hard X-ray emission.
6. Summary
We have considered ten simultaneous RXTE/radio hard state observa-
tions of GX 339−4, and over one hundred RXTE/radio observations
of Cyg X-1. We have fit the former spectra with a very simple, but
remarkably successful, phenomenological model consisting of a doubly
broken power-law with an exponential roll-over plus a gaussian line.
For GX 339−4, the break between the radio and soft X-ray power-
laws occurs in the IR to optical range, in agreement with prior work
[13, 1]. In contrast to prior works, we have fit the X-ray data in 'detector
space' and provided a quantitative assessment of the extrapolated break
location.
The scaling of the radio-to-X-ray break location with integrated X-
ray flux agrees reasonably well with predictions of jet models wherein
a large fraction of the soft X-ray flux is due to synchrotron emission
from the jet. At least some fraction of the observed soft X-rays may be
attributable to emission from the jet, as opposed to disk or corona. On
mnowak_talk.tex; 18/11/2018; 5:41; p.7
8
Michael Nowak
the other hand, we have evidence in the Cyg X-1 failed state transitions
and soft state transition, that the correlation between radio flux and
integrated X-ray flux can take on different amplitudes during different
hard state episodes. There is also evidence in Cyg X-1 that the radio/X-
ray correlation is more fundamental to the hard X-ray band. In jet
models, this band, which essentially encompasses the third, highest
energy, power-law component in our model fits (and also encompasses
the exponential cutoff), is possibly attributable to the synchrotron self-
Compton (SSC) emission from the base of the jet [12, 11]. It is therefore
quite reasonable to expect a strong coupling between the radio and
hard X-ray flux; however, these models are more complex than simple
pure synchrotron models, and are only now beginning to be explored
quantitatively [12, 11].
The results presented here suggest, at the very least, some obvious
observational strategies. Given the break energy correlations, it would
be extremely useful to have not only a radio amplitude for each X-ray
observation, but also a radio slope. Furthermore, the predicted break
for the brightest observation of GX 339−4, E, occurs in the blue end of
the optical. Thus, ideally multi-wavelength observations would consist
of radio, broad band X-ray, and IR through optical coverage. This is
an admittedly difficult task, but BHC are demonstrating via spectral
correlations that all these energy regimes are fundamentally related to
activity near the central engine.
Finally, it is important to obtain multi-wavelength observations of
multiple episodes of each of the spectral states. For example, if there
are indeed 'parallel tracks' in the radio/X-ray correlations, it would
be interesting to determine whether the amplitude of the radio/X-ray
correlation is related to the flux at which the outbursting source transits
from the low/hard to high/soft state. If such observations can be made
with more quantitative detail, we will have vital clues to determining
the relative contributions of coronae and jets, and the coupling between
these two components, for black hole binary systems.
Acknowledgements
This talk later evolved into the paper cited as Nowak et al. (2004).
I would therefore like to thank my coauthors, J. Wilms, S. Heinz, G.
Pooley, K. Pottschmidt, and S. Corbel. It is also a pleasure to acknowl-
edge useful conversations with Sera Markoff and Jeroen Homan. This
work has been supported by NASA grants SV3-73016 and GO4-5041X,
and NSF grant INT-0233441.
mnowak_talk.tex; 18/11/2018; 5:41; p.8
Monitoring of GX 339−4 and Cyg X-1
9
References
1. Corbel, S. and R. P. Fender: 2002, 'Near-Infrared Synchrotron Emission from
the Compact Jet of GX 339−4'. ApJ 573, L35 -- L39.
2. Corbel, S., M. A. Nowak, R. P. Fender, A. K. Tzioumis, and S. Markoff: 2003,
'Radio/X-ray correlation in the low/hard state of GX 339−4'. A&A 400,
1007 -- 1012.
3. Falcke, H. and P. L. Biermann: 1995, 'The jet-disk symbiosis. I. Radio to X-ray
emission models for quasars.'. A&A 293, 665 -- 682.
4. Fender, R. P.: 2001, 'Powerful jets from black hole X-ray binaries in low/hard
X-ray states'. MNRAS 322, 31 -- 42.
5. Gallo, E., R. P. Fender, and G. G. Pooley: 2003, 'A universal radio-X-ray
correlation in low/hard state black hole binaries'. MNRAS 344, 60 -- 72.
6. Heinz, S.: 2004, 'Constraints on the role of synchrotron X-rays from jets of
accreting black holes'. MNRAS. in press (astro-ph/0409029).
7. Heinz, S. and R. A. Sunyaev: 2003, 'The non-linear dependence of flux on black
hole mass and accretion rate in core-dominated jets'. MNRAS 343, L59 -- L64.
8. Homan, J., M. Buxton, S. Markoff, C. Bailyn, E. Nespoli, and T. Belloni:
2004, 'Multi-wavelength Observations of the 2002 Outburst of GX 339−4: Two
Patterns of X-ray Optical/Near Infrared Behavior'. ApJ. submitted.
9. Houck, J. C. and L. A. Denicola: 2000, 'ISIS: An Interactive Spectral Interpre-
tation System for High Resolution X-Ray Spectroscopy'. In: ASP Conf. Ser.
216: Astronomical Data Analysis Software and Systems IX, Vol. 9. p. 591.
10. Markoff, S., H. Falcke, and R. Fender: 2001, 'A jet model for the broad-
band spectrum of XTE J1118+480: optically thin synchrotron X-rays in the
Low/Hard spectral state'. ApJ 372, L25 -- L28.
11. Markoff, S. and M. Nowak: 2004,
'Constraining X-ray Jet Models with
Reflection'. ApJ 609, 972 -- 976.
12. Markoff, S., M. Nowak, S. Corbel, R. Fender, and H. Falcke: 2003, 'Exploring
the role of jets in the radio/X-ray correlations of GX 339−4'. A&A 397,
645 -- 658.
13. Motch, C., S. A. Ilovaisky, C. Chevalier, and P. Angebault: 1985, 'An IR, optical
and X-ray study of the two state behaviour of GX 339−4'. Space Sci. Rev. 40,
219.
14. Nowak, M. A., J. Wilms, and J. B. Dove: 2002, 'Coronal-temporal correlations
in GX 339−4: Hysteresis, Possible Reflection Changes, and Implications for
ADAFs'. MNRAS 332, 856 -- 878.
15. Nowak, M. A., J. Wilms, S. Heinz, G. Pooley, K. Pottschmidt, and S. Corbel:
2004, 'Is the 'IR Coincidence' Just That?'. ApJ. submitted.
16. Pottschmidt, K., J. Wilms, M. A. Nowak, G. G. Pooley, T. Gleissner, W. A.
Heindl, D. M. Smith, R. Remillard, and R. Staubert: 2003, 'Long Term Vari-
ability of Cyg X-1 (1998 to 2001) I. Systematic spectral-temporal correlations
in the hard state'. A&A 407, 1039 -- 1058.
17. Wilms, J., M. A. Nowak, J. B. Dove, R. P. Fender, and T. di Matteo: 1999,
'Low Luminosity states of the black hole candidate GX 339−4. I. ASCA and
simultaneous radio/RXTE observations'. ApJ 522, 460 -- 475.
mnowak_talk.tex; 18/11/2018; 5:41; p.9
|
astro-ph/0208455 | 1 | 0208 | 2002-08-26T13:38:27 | Be/X-ray Binaries | [
"astro-ph"
] | A review of basic properties of Be/X-ray binaries is presented. These systems (called also hard X-ray transients), which form the most numerous class of massive X-ray binaries in the Galaxy, are composed of Be stars and neutron stars (X-ray pulsars) on wide (orbital periods 17-263 d), usually eccentric (eccentricities 0.1-0.9) orbits. The systems contain two quasi-Keplerian (ratio of radial to orbital velocity components less than 0.01) discs: decretion disc around Be star and accretion disc around neutron star. Both discs are temporary: decretion disc disperses and refills on time scales of years (dynamical evolution of the disc, formerly known as the "activity of a Be star"), while accretion disc disperses and refills on time scales of weeks to months (which is related to the orbital motion on an eccentric orbit and, on some occasions, also to the major instabilities of the other disc). Accretion disc might be absent over a longer period of time (years), if the other disc is very weak or absent. The X-ray emission of Be/X-ray binaries has distinctly transient nature and is controlled by the centrifugal gate mechanism, which, in turn, is operated both by the periastron passages (Type I bursts) and by the dynamical evolution of the decretion disc (both types of bursts). The X-ray pulsars in these systems rotate at equilibrium periods (with the possible exception of the slowest pulsars). Be/X-ray binaries are excellent laboratories for investigation of both the evolutionary processes in the two discs and of the evolution of the neutron star rotation. | astro-ph | astro-ph |
Be/X- RAY BINARIES
JANUSZ ZI ´O LKOWSKI
Copernicus Astronomical Center, ul. Bartycka 18, 00-716 Warsaw, Poland
ABSTRACT. A review of basic properties of Be/X-ray binaries is presented. These systems (called
also hard X-ray transients), which form the most numerous class of massive X-ray binaries in the
Galaxy, are composed of Be stars and neutron stars (X-ray pulsars) on wide (Porb ∼ 17 − 263 d),
usually eccentric (e ∼ 0.1 − 0.9) orbits. The systems contain two quasi-Keplerian ( vr /vorb <∼ 10−2)
discs: decretion disc around Be star and accretion disc around neutron star. Both discs are temporary:
decretion disc disperses and refills on time scales ∼ years (dynamical evolution of the disc, formerly
known as the "activity of a Be star"), while accretion disc disperses and refills on time scales ∼ weeks
to months (which is related to the orbital motion on an eccentric orbit and, on some occasions, also to
the major instabilities of the other disc). Accretion disc might be absent over a longer period of time (∼
years), if the other disc is very weak or absent. The X-ray emission of Be/X-ray binaries has distinctly
transient nature and is controlled by the centrifugal gate mechanism, which, in turn, is operated both
by the periastron passages (Type I bursts) and by the dynamical evolution of the decretion disc (both
types of bursts). The X-ray pulsars in these systems rotate at equilibrium periods (with the possible
exception of the slowest pulsars). Be/X-ray binaries are excellent laboratories for investigation of both
the evolutionary processes in the two discs and of the evolution of the neutron star rotation.
1. Introduction
At present, there are no doubts, that Be/X-ray systems (or hard X-ray transients) dom-
inate the population of the massive X-ray binaries. These systems, whose primaries
are Be stars and secondaries are neutron stars were initially believed to be just atyp-
ical cases of massive X-ray binaries (typical cases were supergiant systems). The final
Uhuru catalogue (Forman et al., 1977) listed just 1 such system (as opposed to 6 su-
pergiant systems). Catalogue of Bradt et al. (1978), which was compilation of Uhuru,
SAS-3, Ariel, Copernicus and HEAO-A sources, listed 5 Be/X-ray systems (and 8 su-
pergiant systems). During my previous Vulcano talk on Be/X-ray binaries, 9 years ago
(Zi´o lkowski, 1992), I listed 13 such binaries. Van Paradijs' (1995) catalogue contained
14 Be/X-ray systems (and 14 supergiant systems). In this review, I will present a list of
63 Be/X-ray binaries (while the number of the presently known supergiant systems is
20).
As one can see, the number of the known Be/X-ray systems is growing fast and, due
to the transient nature of their emission, it is likely to continue its fast growth in the
future. At the same time, the number of the known supergiant systems (in our Galaxy)
is already saturated and is not expected to increase substantially,
One may notice that the situation in the field of the low mass X-ray binaries is
similar. The fastest growing class of these systems is the group of soft X-ray transients
(or X-ray Novae). These systems, composed typically of a black hole and a low mass
optical companion, are being discovered at a high rate (again, due to the transient nature
of their X-ray emission). In near future, they will, probably, dominate the population of
the low mass X-ray binaries. It seems, that we have already detected almost all strong
permanent X-ray sources in the Galaxy. In the future, we will detect mostly transient
sources (both massive hard X-ray transients and low mass soft X-ray transients).
2. Basic Properties
In this section, I will list basic characteristics of Be/X-ray binaries and confront them
with the properties of the other massive X-ray systems.
(1) The optical component is a Be star (in the other massive systems it is usually an
OB supergiant).
(2) The X-ray component is a strongly magnetized neutron star (an X-ray pulsar). In
most of the supergiant systems, it is also a neutron star, but in five systems it is a
black hole. No Be/black hole binary was found yet.
(3) Orbits are elliptical with substantial eccentricities - usually e >∼ 0.3 (in the other
systems orbits are usually circular).
(4) The orbital periods are rather long - between 17 and 263 days (in the other systems
they are almost always shorter than 10 days).
(5) The optical components are not substantially evolved and they are substantially
smaller than their Roche lobes. In the other massive systems, the optical components
(supergiants) are substantially evolved (they are overluminous for their masses or
undermassive for their luminosities) and they approximately fill their Roche lobes.
(6) X-Ray emission (with a few exceptions) has distinctly transient nature with rather
short active phases (a flaring behaviour). There are two types of flares, which are
classified as Type I outbursts (smaller and regularly repeating) and Type II out-
bursts (larger and irregular). In the supergiant systems the emission is rather per-
manent, although it may be strongly variable.
(7) X-Ray emission has a hard spectrum, with typical kT >∼ 15 keV (this property is
shared with the other massive systems). Because of this property (and the transient
nature of their X-ray emission), the Be/X-ray binaries are also called hard X-ray
transients, as opposed to the so-called soft X-ray transients (or X-ray Novae), which
have typical kT <∼ 1 keV and form a distinct group within the class of the low mass
X-ray binaries.
(8) In all but three systems that are confirmed or probable members of the class of
Be/X-ray binaries, the neutron star is seen as an X-ray pulsar. The range of observed
rotational periods covers more than four decades - from 0.03 sec to 1404 sec. In the
supergiant systems the neutron star is also usually seen as an X-ray pulsar and the
observed range of pulse periods is similarly wide (0.7 sec to 10000 sec).
The list of the known Be/X-ray binaries and some of their parameters is given in
Table 1. The list includes all X-ray systems containing a Be star and having at least one
period (orbital or spin of the neutron star) determined.
Tab. 1 − Be/X-Ray Binaries
Name
Pspin
[s]
Porb
[d]
e
Sp. type
Ref
SAX J0635+0533
A
SMC
0538−7
X−2
RX J0059.2−7138
AX J0105−722
4U
RX J0502.9−6626
0115+63
V
0332+53
GRO J1750−27
XTE J0052−723
RX J0051.8−7231
AX J0049−732
2S
GS
1553−54
0834−430
EXO 0531.1−6609
RX J0052.1−7319
XTE J1946+274
2S
1417−62
GRO J1948+32
RX J0117.6−7330
XTE J1543−568
XTE J0111.2−7317
RX J0812.4−3114
2030+37
EXO
XTE J0053−724
SAX J0054.9−7226
Cep
RX J0529.8−6556
RX J0049.1−7250
AX J0051−722
X−4
GRO J1008−57
1845−024
2S
RX J0544.1−7100
AX J0057.4−7325
0535+26
0728−25
A
4U
AX J1820.5−1434
RX J0052.9−7158
SAX J1324.4−6200
AX J0051.6−7311
0.034
0.069
2.37
2.763
3.343
3.61
4.06
4.37
4.45
4.782
8.9
9.132
9.26
12.3
13.7
15.3
15.8
17.6
18.76
22.07
27.1
30.95
31.89
41.7
46.63
58.97
66.3
69.5
74.68
90.65
93.5
94.8
96.08
101.42
103.5
103.2
152.26
169.3
170.84
172.4
< 0.09
0.12
30.6
105.8
25.4
172
42.12
41.7
76.6
< 0.03
81
46.03
139
65
120
247.5
242.18
110.3
34.5
0.41
0.66
0.88
0.47
16.7
> 0.5
24.31
0.34
34.25
29.82
0.31
0.446 B1 Ve
< 0.25
B0.5 IIIe
B2 V-IIIe
B1.5 Ve
B1 IIIe
Be ?
B0.2 Ve
B0 IIIe
O8.5 Ve
Be ?
Be ?
Be
Be ?
Be ?
B0-2 V-IIIe
Be ?
Be
Be ?
B0e
B0.5 IIIe
B0.7 Ve
B0-2 V-IIIe
B0.2 IVe
B0e
B1 Ve
B0-1 V-IIIe
B1e
B2 V-IIIe
Be
Be
O9e-B1e
Be ?
Be
Be ?
O9.7 IIIe
O8.5 Ve
Be ?
Be
Be
Be
1
1,2
1,3,4
1
1
1,5
1,6
1,5
1,5
4,7
1
1
1,5
1,5
1
1
1,2
1,5,6
1,8,9
1
2,10
1,11
1,12,13
1,2
1
1
1,2
1
1
1,14
1,15
1,16
1
4,17
1,2,18
1,2
1,19
1,19
1
1,20
Tab. 1 − Be/X-Ray Binaries (cont.)
Name
Pspin
[s]
Porb
[d]
e
Sp. type
Ref
GRO J2058+42
RX J0440.9+4431
AX J1749.2−2725
J1858+034
XTE
GX
AX J0058−720
4U
AX J0051−733
304−1
1145−61
SAX J0103.2−7209
4U
A
2206+54
1118−61
SAX J1452.8−5949
RX J0101.3−7211
AX J170006−4157
AX J0049.4−7323
4U
RX J1037.5−5647
SAX J2239.3+6116
RX J0146.9+6121
1E 0236.6+6100
0352+30
γ
Cas
RX J0535.0−6700
198
202.5
220.38
221.0
272
280.4
292.4
323.2
345.2
392
406.4
437.4
455
714.5
755.5
837.7
862
1247.2
1404.2
110
132.5
187.5
1.416 ??
9.57 ?
> 0.5
> 0.5
250.3
262.6
26.49
203.59
241
?
0.11
0.26
Be
B0 V-IIIe
Be
Be ?
B0.7 Ve
Be ?
B0.2 IIIe
Be
O9-B1 V-IIIe
B1e
O9.5 V-IIIe
Be ?
Be
Be
Be
B0 Ve
B0 Ve
B0 V−B2 IIIe
B1 Ve
B0e
B0.5 Ve
Be
1
1
1,19
1
1,2
1,19
1,2
1
1
1,21
1,5
1
22
1,19
23
1,24
1
1,25
1
1,26
1,27
1
NOTES:
e − eccentricity
Sp. type − spectral type of the optical component
Ref − references
REFERENCES:
(1) Liu et al. 2000; (2) Okazaki and Negueruela 2001; (3) Corbet aand Marshall 2000; (4)
Mereghetti 2001; (5) Bildsten et al. 1997; (6) Negueruela 1998; (7) Corbet et al. 2001; (8)
Levine and Corbet 2000; (9) Negueruela et al. 2000a; (10) Marshall et al. 2000; (11) Yokogawa
et al. 2000a; (12) Corbet and Peele 2000; (13) Reig et al. 2000; (14) Israel et al. 1998 (15)
Negueruela and Okazaki, 2000; (16) Finger et al. 1999; (17) Torii et al. 2000; (18) Negueruela
et al. 2000b; (19) Haberl and Sasaki 2000; (20) Yokogawa et al. 2000b; (21) Corbet et al. 2000;
(22) Sasaki et al. 2001; (23) Ueno et al. 2000; (24) Delgado-Marti et al. 2001; (25) In't Zand
et al. 2001; (26) Zamanov et al. 2001; (27) Harmanec et al. 2000.
3. Temporal Behaviour of the X-Ray Emission
Be/X-ray binaries exhibit three types of behaviour:
(1) Type I outbursts
They last days to weeks and are connected with the periastron passages of neutron
stars orbiting Be stars on alongated orbits. They repeat after the time intervals equal
to the orbital periods or their multiples. The recurrency is not very strict. On some
occasions, single outbursts are missing (or are unusually weak). On some others, the
outbursts disappear for extended periods (∼ years or decades). Almost all Be/X-ray
binaries show Type I outbursts on some occasions, but the patterns are very different.
Some systems (like A 0535+26) produce Type I outbursts during almost every periastron
passage for a long interval of time (years or decades) and then the period of the missing
outbursts follows. In some others (like 4U 0115+63 or V 0332+53) the outbursts are
very rare: less than 1 % of all periastron passages results in Type I outbursts. In the
second group, the typical pattern is composed of short series of the few Type I outbursts,
separated by long periods (years or decades) of inactivity, during which only Type II
outbursts occur.
(2) Type II outbursts
They last several weeks and are not correlated with any particular orbital phase.
These bursts are typically much stronger (by an order of magnitude or more) than
Type I bursts. They occur irregularly, but the typical recurrence time is of the order
of several years (although the intervals as long as two decades were also noted). Type
II bursts are usually preceded by an increased activity of the Be companion (enhanced
emission lines). This last fact, together with the lack of the correlation with the orbital
phase, indicates that Type II bursts are connected with the activity of a Be star (while
Type I bursts are clearly connected with the periastron passages).
(3) Persistent emission
Five Be/X-ray binaries are seen as permanent X-ray sources (with the emission level
variable only by a factor of few). All of them are probably wide orbit (Porb >∼ 200 days)
systems, containing very slow pulsars (Pspin in the range 202 to 1404 sec). Some of these
pulsars might rotate slower than at equilibrium periods (Zi´o lkowski, 2001).
4. Why the Emission is Transient?
In order to understand the temporal behaviour of the X-ray emission from Be/X-ray
binaries, we have to invoke the, so called, accretion gate mechanism which plays a
fundamental role in switching on and off the X-ray emission. Then, we have to consider
the mechanisms that operate the accretion gate by controlling the supply of the external
matter to the magnetosphere of the neutron star.
4.1. The accretion gate mechanism
The accretion gate mechanism operates due to changes in the mass flux of the mat-
ter falling on the magnetosphere of the neutron star. The size of the magnetosphere is
determined by the instantaneous balance between the magnetic pressure and the dy-
namic pressure of the infalling matter. Therefore the magnetosphere changes its size.
When the flux of the infalling matter decreases, the magnetosphere expands. When the
flux increases - the magnetosphere gets squeezed. For example, the magnetosphere of
a neutron star traveling along an elongated orbit around a Be star is much smaller
at periastron than at apoastron. We have good reasons to believe that neutron stars
in Be/X-ray binaries, similarly as many of the other X-ray pulsars, are rotating at so
called equilibrium periods (Zi´o lkowski 1980, 1985; Joss and Rappaport 1984, Stella et al.
1986, Giovannelli and Zi´o lkowski 1990, and references therein). The equilibrium period
is defined as a period at which the outer edge of magnetosphere rotates with the Keple-
rian velocity. At this period the accelerating accretion torque and the braking propeller
torque should balance each other and the period, in the first approximation, should
be constant. For a given magnetosphere (i.e. for a given magnetic field strength) the
equilibrium period depends mainly on the mass flux of the infalling matter (because it
determines the size of the magnetosphere which is assumed to corotate with the neutron
star). This dependence is given as Peq ∝ M −3/7 (Davidson and Ostriker 1973, van den
Heuvel 1977, Lamb 1977). Returning to our neutron star on an elongated orbit, we see
that its instantaneous equilibrium period varies along the orbit as it tracks the variable
mass flux. At periastron Peq is much smaller than at apastron. As a result, we find that
in a typical situation Pspin > Peq ("slow" pulsar) at periastron but Pspin < Peq ("fast"
pulsar) at apoastron. It means that at periastron the accretion gate is open and the ro-
tation of the pulsar (neutron star) is accelerated, while at apoastron the accretion gate
is closed and the pulsar is braked. The real rotation period Pspin will adjust itself so
that the action of the accretion torque along the inner part of the orbit and the braking
torque along the outer part, integrated over the full orbital cycle, will cancel each other.
Therefore the statement that X-ray pulsars in Be/X-ray binaries rotate at equilibrium
periods should be read as "equilibrium periods averaged over orbital cycle (or, in fact,
over many orbital cycles, because of the intrinsic variability of a Be star)". One should
remember, however, that along the orbit (except for the two points) the real rotation
period Pspin substantially differs from the instantaneous value of Peq and, therefore, the
pulsar is always, either "fast" or "slow".
The picture presented above applies directly only to those Be/X-ray binaries which
exhibit (more or less regularly) Type I outbursts. However, the accretion gate mechanism
plays a crucial role in all Be/X-ray binaries.
4.2. Two states of an X-ray pulsar
The previous paragraph leads us to the picture of two possible states of an X-ray pulsar
in a binary system:
(1) Accretor
This state typically occurs when Pspin > Peq. Since the velocity of the outer edge of
magnetoshere is smaller then the Keplerian velocity, there is no centrifugal barrier. The
matter can approach the neutron star and get accreted (accretion gate open). Due to
accretion taking place, the neutron star is a strong X-ray emitter and also experiences
a substantial spin-up of its rotation (the accreted matter carries a substantial angular
momentum from the inner edge of the accretion disc). The characteristic properties of
the accretor state are therefore: (1) "slowness" of the pulsar (Pspin > Peq), (2) strong
X-ray emission, (3) rapid spin-up.
(2) Propeller
This is a typical state when Pspin < Peq.. In this case the velocity of the outer edge of
magnetosphere is larger than the Keplerian velocity and there exist a centrifugal barrier
against accretion. The matter that would like to get accreted is rather expelled by the
magnetosphere acting as a propeller (llarionov and Sunyaev 1975, Davies et al. 1979).
The accretion gate is closed. Only marginal amount of matter can get accreted and so
the X-ray luminosity is low or undetectable. However the neutron star loses rotational
energy because magnetospheric propeller has to work on the infalling matter to get it
expelled and so it experiences a slow-down of its rotation. The characteristic properties
of the propeller state are therefore: (1) "fastness" of the pulsar (Pspin < Peq.), (2) low
or undetectable X-ray emission, (3) significant spin-down.
4.3. The Variability of the Supply of the External Matter to the Neutron Star
From the last two paragraphs, it is clear that to operate (to open or to close) the
accretion gate, one has to adjust accordingly (to increase or to decrease) the flux of
the matter falling on the magnetosphere of the neutron star. In the Be/X-ray binaries,
the matter in question comes from the decretion disc around the Be component (known
earlier as "an envelope of a Be star"). There exist two mechanisms that might control
the supply of that matter:
(1) The eccentricity of the orbit of the neutron star
The neutron star moving on an elongated orbit around the Be star is far from the
decretion disc near the apoastron and quite close to it (or may even enter it) near the
periastron. This may result in the larger mass flux of the infalling matter (and the open
accretion gate) near the periastron and the smaller mass flux (and the closed gate)
near apoastron. This mechanism is, most likely, responsible for Type I outbursts, which
repeat, more or less regularly, at periastron passages.
(2) The dynamical evolution of the decretion disc
Both the observational evidence and the theoretical considerations (see section 6)
indicate that the decretion disc is not a stable structure and evolves dynamically on
a time scale of, typically, several years. This evolution leads to cyclical dispersals and
refillings of the disc. Such variability produces on some occasions large infalls of matter
on the neutron star (Type II bursts). On some other occasions (e.g. empty disc) this
results in so low supply of the matter to the vicinity of the neutron star that even in
a system producing, rather regularly, Type I bursts, the accretion gate remains closed
during periastron passages (the case of the missing Type I bursts).
5. Accretion Discs around Neutron Stars
Accretion discs around neutron stars in the Be/X-ray binaries have all properties of
the accretion discs in other X-ray systems, except for their, possibly, transient nature.
Accretion discs are, certainly, present during Type II outbursts. This might be directly
inferred from the observed spin-ups of neutron stars during these bursts. The time
scales of the spin-ups and their correlation with the X-ray luminosities indicate that
the accreted angular momentum corresponds to the orbital angular momentum at the
inner edge of the disc. The best studied case is the major February 1994 outburst of
A 0535+26 (Finger et al., 1996; Bildstein et al., 1997), during which a spin-up on time
scale ∼ 25 years was seen at the peak of the outbursts. The same outburst provided an
independent evidence of the presence of an accretion disc, based on the interpretation
of the QPO detected during that event. Both the frequency of the QPO (interpreted
as the Keplerian frequency at the inner edge of the disc) and its correlation with the
spin-up rate and with the X-ray luminosity (Finger et al., 1996; Bildstein et al., 1997)
fully support the model based on an accretion disc. The accretion discs must be present
also during most of Type I bursts. The system A 0535+26 demonstrated spin-ups on
time scales ∼ 100 years during many Type I bursts (Zi´o lkowski, 1985; Giovannelli and
Sabau Graziati, 1992). Similar spin-ups were observed during all (2S 1845-024; Finger
et al., 1999) or, at least, some Type I bursts of many other Be/X-ray systems (e.g. 4U
0115+63, GS 0834-430, 2S 1417-62, EXO 2030+375, 4U 1145-61 − see Bildstein et al.,
1997). Neutron stars in many Be/X-ray systems are known to experience spin-downs
between the outbursts. This is in agreement with the predictions of the simple model,
since they are expected to enter a propeller phase between the outbursts (see sections
4.1 and 4.2 above). For the system A 0535+26 the time scales of these spin-downs are
of the order of 1000 years (Zi´o lkowski, 1985; Giovannelli and Sabau Graziati, 1992).
In principle, the accretion disc may survive during the propeller phase. However, it
seems that the accretion discs in the Be/X-ray binaries are not persistent. Clark et
al. (1999) searched for the optical/infrared contribution from the accretion disc in A
0535+26 during the X-ray quiescent phase (propeller state). They found no significant
contribution which implied that the accretion disc was either absent, or, if present, then
only as a weak remnant.
Let us summarize our considerations. The accretion discs are, probably, absent during
the X-ray quiescent phases. They built up on time scales of days to weeks during the
periods of the enhanced infall of the matter on the magnetosphere of a neutron star
(before both Type I and Type II bursts). They disperse on time scales of weeks to
months, when the enhanced supply of the matter ceases.
The apparent exceptions to this picture are the five persistent systems (see section
3.3) that apparently maintain permanent accretion discs.
6. Decretion Discs around Be Stars
The term "excretion disc" was introduced by B. Paczy´nski during 1979 IAU General
Assembly (Paczy´nski, 1980). Simple, particle trajectories models of such discs were
discussed by Paczy´nski and Rudak (1980). St¸epi´nski (1980) considered the discs around
Be stars and proposed that variability of these stars is related to the changes in the sizes
of the discs. The concept of the presence of two discs (one excreting and one accreting) in
a Be/X-ray binary was first mentioned by Giovannelli and Zi´o lkowski (1990). Lee et al.
(1991) discussed the possible existence of the outflowing viscous discs around Be stars.
The next step was made when different authors (Okazaki, 1991; Papaloizu et al., 1992;
Telting et al., 1994; Hanuschik et al., 1995) indicated that the one-armed instability
developing in the outflowing discs may explain the, so called, V/R variability, observed
in Be stars. In recent years, the outflowing viscous discs were used to describe, in details,
the circumstellar matter around Be stars known earlier as "an envelope of a Be star"
(Okazaki, 1997; Porter, 1999; Negueruela and Okazaki, 2000b; Okazaki, 2000; Negueruela
at al., 2001; Okazaki and Negueruela, 2001). In meantime, the term "excretion" was
replaced by an apparently more correct (politically) term "decretion". The modelling
with the help of the viscous decretion discs appeared to be by far more succesful in
describing the circumstellar matter, than earlier descriptions in terms of "equatorial
winds", "expanding envelopes" or "ejected shells". In particular, the viscous disc models
were able to explain the very low outflow velocities (the observed upper limits are, at
most, few km/sec) and, also, to explain the V/R variability. The viscous decretion discs
are very similar to the, well known, viscous accretion discs, except for the changed
sign of the rate of the mass flow. Some aspects of these modellings (supply of the
matter with the sufficient angular momentum to the inner edge of the disc, interaction
of the stellar radiation with the matter of the disc) are not fully solved yet, but the
general picture is quite convincing. The viscosity in the decretion discs is usually assumed
(similarly as for accretion discs) in the form of α-viscosity. The discs are almost Keplerian
(rotationally supported) which explains the very low values of the radial component
of velocity. Nearly Keplerian discs (both inflowing and outflowing) were, since a long
time, known to undergo a global one-armed oscillation instability (Kato, 1983). This
instability (progressing density waves) provide a very successful explanation of V/R
variability, observed both in isolated Be stars and in members of Be/X-ray systems.
This phenomenon manifests itself in the form of quasi-cyclical changes of the ratio of
the strengths of the V(iolet) peak to the R(ed) peak in the double profile emission
lines. This variability (best seen for the Hα line) includes phases, when only one peak is
visible. The time scales of the quasi-cycles range from months to years or decades. The
theoretical line profiles calculated for the discs with the asymmetric matter distribution
(due to progressing density waves) were found to be in a good agreement with the
observed profiles (Hummel and Vrancken, 1995; Okazaki, 1996; Hummel and Hanuschik,
1997). Also the theoretical time scales calculated for the one-armed oscillation instability
agreed with the observed time scales of V/R variability (Negueruela et al., 2001).
7. The Interaction Between the Decretion Disc and the Orbiting Neutron
Star
Until recently, it was thought that this interaction works only in one direction. It is
obvious that the disc can strongly influence the behaviour of the neutron star. If it sends
large enough flux of the matter towards the neutron star, then the accretion gate will
open and the neutron star will undergo an X-ray outbursts. It seemed that the neutron
star could not significantly influence the "envelope of the Be star" (as it was termed
then). This point of view changed recently. Artymowicz and Lubow (1994) investigated
the truncation mechanisms for the accretion discs and found that the discs must be much
smaller in the systems with large eccentricities because of the tidal/resonant truncation.
The same mechanism works also for the decretion discs. The orbiting neutron star
exerts a negative tidal torque on the matter of the decretion disc (removes the angular
momentum). This torque diminishes the action of the viscous torque, supporting the
disc, and causes that outside of the certain radius the disc cannot be supported any
more (tidal truncation). This effect was investigated, for several Be/X-ray binaries, by
Negueruela and Okazaki (2000a,b), and Okazaki and Negueruela (2001). They found,
that the disc is usually truncated at 4:1 or 3:1 resonances (i.e. at radii for which the
Keplerian period in the disc is 4 or 3 times shorter than the orbital period). This
mechanism is known as tidal/resonant truncation. Due to its action, in the systems
with small eccentricities, the disc is always smaller than the Roche lobe around the Be
star and the flow of the matter towards the neutron star is effectively blocked. However,
in the systems with large eccentricities, the disc may extend beyond the Roche lobe,
during the periastron passages. The accretion on the neutron star will be then possible
during these passages. Therefore, the systems in the first group (like 4U 0115+63 or V
0332+53), typically, should not produce Type I bursts, while the systems in the second
group (like A 0538−66, A 0535+26 or GRO J1008−57) should produce Type I bursts
during most of their periastron passages. The observations strongly support this picture.
The tidal/resonant truncation has some further consequences. Due to truncation,
the matter accumulates in the outer rings of the disc. When the outer part of the disc
becomes sufficiently optically thick, it may warp under the influence of the radiation of
the Be star (Porter, 1998). The observations (the variability of the shape of the emission
lines) tell us that the sufficiently warped disc start precessing on a time scale of weeks
to months. The combined effects of the global one-armed oscillations and of the warping
lead to asymmetric (elongated) configuration of the disc. Such configuration may permit
an overcoming of the truncation. The, relatively, high density matter accumulated in the
outer part of the disc will overflow the critical Roche lobe and fall on the neutron star
leading to the Type II outburst. If the distorted disc happens to be elongated towards the
periastron, it may supply the matter to the neutron star during the several periastron
passages, leading to the temporary Type I outbursts in the systems, which normally do
not show the bursts of this type. The Type II bursts are probably always associated
with the major disturbances of the decretion disc, which frequently lead to the total
disruption of the disc and the disc-less phase of the Be star. Part of the matter of the
disc is accreted by the neutron star (during the Type II burst), most of it is, probably,
reaccreted by the Be star. The decretion disc is then rebuilt on a time scale of, typically,
a few years. The observations support the phenomenological picture presented above.
In addition to the observational properties described earlier we may mention here the
changes of the strength of the Hα line observed over more than 10 years period, before
and after the major Type II burst in A 1118−61 (Coe et al., 1994). We may also add
the correlation found recently for 1E 0236.6+6100/LS I+61o235 system. The size of the
decretion disc in this system (as measured by the strength of the Hα line) was known to
vary cyclically with the periodicity of ∼ 4.3 years. The system was also known to exhibit
radio outbursts (in addition to X-ray maxima) during the periastron passages. Zamanov
and Marti (2000) found a good correlation between the strength of the radio outbursts
and the strength of the Hα line as they change over the ∼ 1584 days modulation period.
8. The Comparison of Be Stars in Be/X-ray Binaries with Isolated Be Stars
The isolated Be stars and Be stars in Be/X-ray binaries are, in many respects, quite
similar. Both types have decretion discs that evolve dynamically on time scale of years
to decades. In both cases, the dynamical evolution includes one-armed oscillations, V/R
variability and disc-less phases. However, there are also substantial differences concern-
ing both Be stars themselves and their decretion discs.
8.1. Be stars
The distribution of Be stars among different spectral subtypes was compared for both
groups of Be stars by Negueruela (1998). The difference is striking. For isolated Be
stars, the distribution starts at B0, rises sharply peaking at B2, declines towards B4
and then remains at approximately constant level, extending until at least A0. For Be
stars in Be/X-ray binaries, the distribution starts at O8, peaks at B0 and ends at B2.
The lack of the later spectral type (less massive) Be stars in binary systems may be
explained with the help of the results of the evolutionary calculations by Van Bever and
Vanbeveren (1997). They showed, that in scenarios involving the non-conservative mass
transfer, binaries composed of a Be star and a neutron star are produced only with early
type Be components. The less massive systems that would lead to binaries containing
late type Be components are disrupted during the supernova explosion (the wide orbit
systems are, generally, easy to disrupt).
8.2. Decretion discs
The decretion discs around isolated Be stars are, in many respects, similar to those in
Be/X-ray binaries, as was noted earlier. The main difference is the lack of the orbiting
neutron star and, therefore, the lack of the tidal/resonant truncation mechanism. As
a result, the decretion discs in binaries are smaller and denser than those around the
isolated Be stars (Reig et al., 1997; Negueruela and Okazaki, 2000b, Zamanov et al.,
2000). As might be expected, the discs in close binaries are smaller and the discs in
wide orbit binaries - larger. This is clearly seen from the strong positive correlation
between the maximal observed strength of the Hα line and the orbital period of the
binary, noted by Reig et al. (1997). In the few widest systems, the disc is probably only
weakly perturbed by the neutron star. As a result, there are no X-ray outbursts. The
low level accretion and the low level X-ray emission are maintained continuously (and
the only variability is an occasional enhancement).
9. The Pulse Period - Orbital Period Correlation for Be/X-ray Binaries (the
Corbet Diagram)
Corbet (1984) was the first to notice that there exists a strong correlation between the
spin period of a neutron star in a Be/X-ray binary and the orbital period of the system.
The updated version of his diagram (containing 25 systems) is shown in Fig. 1. The
spread is larger than on the diagram that I have shown here nine years ago (there were
only 9 points on the diagram then), but the reality of the correlation leaves no doubt.
This correlation is fully to be expected, if we remember that the neutron star tries to
adjust its spin period to the equilibrium value. This value increases as the flux of the
matter supplied to the vicinity of the neutron star decreases (Peq ∝ M −3/7). In wide
systems, with long orbital periods, the accretion is fed from the outer, low density, parts
of the large (very extended) decretion disc and so the expected mass fluxes are rather
low. This leads to the long spin periods.
Fig. 1. The correlation between the spin period (in seconds) and the orbital period (in days) for
Be/X-ray binaries (Corbet diagram). The diagram contains 25 systems for which both periods
are presently known (the systems AX J0051-733 and 4U 2206+54 with the doubtful values of
the orbital period are omitted).
10. The Summary
Let us summarize the main points of our considerations.
• Be/X-ray binaries (called also hard X-ray transients) form the most numerous class
of massive X-ray binaries in the Galaxy (63 systems presently known as opposed to 20
supergiant systems).
• They are composed of Be stars and neutron stars (X-ray pulsars) on wide (Porb ∼
17 − 263 d), usually eccentric (e ∼ 0.1 − 0.9) orbits.
• X-Ray emission (with a few exceptions) has distinctly transient nature with rather
short active phases (a flaring behaviour). There are two types of flares, which are classi-
fied as Type I outbursts (smaller and, more or less regularly, repeating during periastron
passages over some periods of time) and Type II outbursts (larger and irregular).
• The optical components are not significantly evolved and they are substantially
smaller than their Roche lobes.
• The X-ray pulsars in Be/X-ray binaries rotate at equilibrium periods (with the
possible exception of the slowest pulsars). The range of the observed spin periods is very
wide - from 0.03 sec to 1404 sec.
• The systems contain two quasi-Keplerian ( vr /vorb <∼ 10−2) discs: a decretion
disc around Be star and an accretion disc around the neutron star.
• Both discs are temporary: decretion disc disperses and refills on time scales ∼ years
(dynamical evolution of the disc, formerly known as the "activity of a Be star"), while
accretion disc disperses and refills on time scales ∼ weeks to months (which is related
to the orbital motion on an eccentric orbit and, on some occasions, also to the major
instabilities of the other disc).
• Accretion disc might be absent over a longer period of time (∼ years), if the other
disc is very weak or absent (this phenomenon manifests itself as the missing Type I
bursts in the systems which normally display these bursts).
• The X-ray emission of Be/X-ray binaries is controlled by the centrifugal gate
mechanism, which, in turn, is operated both by the periastron passages (Type I bursts)
and by the dynamical evolution of the decretion disc (both types of bursts).
• Decretion discs are unstable to one-armed oscillation instability (which is probably
responsible for the V/R variability seen in the Be stars).
• Decretion discs in Be/X-ray binaries are truncated by the orbiting neutron stars
(the tidal/resonant truncation).
• Due to truncation, in the systems with small eccentricities, the flow of the matter
towards the neutron star is effectively blocked during the whole orbital cycle (no Type
I bursts). However, in the systems with large eccentricities, the flow is possible during
the periastron passages (regular Type I bursts).
• Due to truncation, the matter accumulates in the outer rings of the decretion disc.
This leads to the warping of the disc and then to its major instability (frequently to the
total disruption). This mechanism is probably responsible for the Type II bursts (and
for occasional Type I bursts in the systems which normally do not show them).
• Be/X-ray binaries are excellent laboratories for investigation of both the evolu-
tionary processes in the two discs and of the evolution of the neutron star rotation.
Acknowledgements
This work was partially supported by the State Committee for Scientific Research
grant No 2 P03C 006 19p01.
References
Artymowicz, P. and Lubow, S.H.: 1994, Astrophys. J. 421, 651.
Bildsten, L., Chakrabarty, D., Chiu, J. Finger, M.H., Koh, D.T., Nelson, R.W., Prince,
T.A., Rubin, B.C., Scott, D.M., Stollberg, M., Vaughan, B.A., Wilson, C.A. and
Wilson, R.B.: 1997, Astrophys. J. Suppl. 113, 367.
Clark, J.S., Lyuty, V.M., Zaitseva, G.V., Larionov, V.M., Larionova, L.V., Finger, M.,
Tarasova, A.E., Roche, P. and Coe, M.J.: 1999, Mon. Not. R. Astr. Soc. 302, 167.
Coe, M.J., Roche, P., Everall, C., Fishman, G.J., Hagedon, K.S., Finger, M., Wilson,
R.B., Buckley, D.A.H., Shrader, C., Fabregat, J., Polcaro, V.F., Giovannelli, F. and
Villada, M.: 1994, Astron. Astrophys. 289, 784.
Corbet, R.H.D.: 1984, Astron. Astrophys. 141, 91.
Corbet, R.H.D. and Marshall, F.E.: 2000, IAUC 7402.
Corbet, R.H.D., Marshall, F.E. and Markwardt, C.B.: 2001, IAUC 7562.
Corbet, R.H.D. and Peele, A.G.: 2000, Astrophys. J. Lett. 530, L33.
Corbet, R., Remillard, R. and Peele, A.: 2000, IAUC 7446.
Davidson, K. and Ostriker, J.P.: 1973, Astrophys. J. 179, 585.
Davies, R.E., Fabian, A.C. and Pringle, J.E.: 1979, Mon. Not. R. Astr. Soc. 186, 779.
Delgado-Marti, H., Levine, A.M., Pfahl, E. and Rappaport, S.A.: 2001, Astrophys. J.
546, 455.
Finger, M.H., Wilson, R.B. and Harmon, B.A.: 1996, Astrophys. J. 459, 288.
Finger, M.H., Bildsten, L., Chakrabarty, D., Prince, T.A., Scott, D.M., Wilson, C.A.,
Wilson, R.B. and Nan Zhang, S.: 1999, Astrophys. J. 517, 449.
Giovannelli, F. and Sabau Graziati, L.:1992, Space Sci. Rev. 59, 1.
Giovannelli, F. and Zi´o lkowski, J.: 1990, Acta Astron. 40, 95.
Haberl, F. and Sasaki, M.: 2000, astro-ph 0005226.
Hanuschik, R.W., Hummel, W., Dietle, O. and Sutorios, E.: 1995, Astron. Astrophys.
300, 163.
Harmanec, P., Habuda, P., Stefl, S., Hadrava, P., Korcakova, D., Koubsky, P., Krticka,
J., Kubat, J., Skoda, P., Slechta, M. and Wolf, M.: 2000, astro-ph 0011516.
Hummel, W. and Vrancken, M.: 1995, Astron. Astrophys. 302, 751.
Hummel, W. and Hanuschik, R.W.: 1997, Astron. Astrophys. 320, 852.
Illarionov, A.F. and Sunyaev, R.A.: 1975, Astron. Astrophys. 39, 195.
In't Zand, J.J.M., Swank, J., Corbet, R.H.D. and Markwardt, C.B.: 2001, astro-ph
0110695.
Israel, G.L., Stella, L., Campana, S., Covino, S., Ricci, D. and Oosterbroek, T.: 1998,
IAUC 6999.
Joss, P. and Rappaport, S.A.: 1984, Ann. Rev. Astron. Astrophys. 22, 537.
Kato, S.: 1983, Publ. Astron. Soc. Japan 35, 249.
Lamb, F.K.: 1977, in Eight Texas Symposium on Relativistic Astrophysics, Annals New
York Academy Sci. 302, 482.
Lee, U., Saio, H. and Osaki, Y.: 1991, Mon. Not. R. Astr. Soc. 250, 432.
Levine, A. and Corbet, R.: 2000, IAUC 7523.
Liu, Q.Z., van Paradijs, J. and van den Heuvel, E.P.J.: 2000, Astron. Astrophys. Suppl.
Ser. 147, 25.
Marshall, F.E., Takeshima, T. and in't Zand, J.: 2000, IAUC 7363.
Mereghetti, S.: 2001, in Frontier Objects in Astrophysics and Particle Physics, F. Gio-
vannelli and G. Mannocchi, eds., Italian Physical Society, Editrice Compositori,
Bologna, Italy, 73, 239.
Negueruela, I.: 1998, astro-ph 9807158.
Negueruela, I., Marco, A., Speziali, R. and Israel, G.L.: 2000a, IAUC 7541.
Negueruela, I. and Okazaki, A.T.: 2000a, astro-ph 0011406.
Negueruela, I. and Okazaki, A.T.: 2000b, astro-ph 0011407.
Negueruela, I., Reig, P., Finger, M.H. and Roche, P.: 2000b, astro-ph 0002272.
Negueruela, I., Okazaki, A.T., Fabregat, J., Coe, M.J., Munari, U. and Tomov, T.: 2001,
astro-ph 0101208.
Okazaki, A.T.; 1991, Publ. Astron. Soc. Japan 43, 75.
Okazaki, A.T.; 1996, Publ. Astron. Soc. Japan 48, 305.
Okazaki, A.T.; 2000, astro-ph 0010517.
Okazaki, A.T.; 1997, in The Interactions of Stars with Their Environment, L.V. Toth,
M. Kun and L. Szabados (eds.), Konkoly Observatory, Budapest, p. 407.
Okazaki, A.T. and Negueruela, I.: 2001, astro-ph 0108037.
Paczy´nski, B.: 1980, Highlights of Astronomy 5, 27.
Paczy´nski, B. and Rudak, B.: 1980, Acta Astron. 30, 237.
Papaloizu, J.C., Savonije, G.J. and Henrichs, H.F.: 1992, Astron. Astrophys. Lett. 265,
L45.
Porter, J.M.: 1998, Astron. Astrophys. 336, 966.
Porter, J.M.: 1999, Astron. Astrophys. 348, 512.
Reig, P., Fabregat, J. and Coe, M.J.: 1997, Astron. Astrophys. 322, 193.
Reig, P., Negueruela, I., Buckley, D.A.H., Coe, M.J., Fabregat, J. and Haigh, N.J.: 2000,
astro-ph 0011305.
Sasaki, M., Haberl, F., Keller, S. and Pietsch, W.: 2001, astro-ph 0102361.
Stella, L., White, N.E. and Rosner, R.: 1986, Astrophys. J. 308, 669.
St¸epi´nski, T.: 1980 Acta Astron. 30, 413.
Telting, J.H., Heemskerk, M.H.M., Henrichs, H.F. and Savonije, G.J.: 1994, Astron.
Astrophys. 288, 588.
Torii, K., Kohmura, T., Yokogawa, J. and Koyama, K.: 2000, IAUC 7441.
Ueno, M., Yokogawa, J., Imanishi, K. and Koyama, K.: 2000, IAUC 7442.
Van Bever, J. and Vanbeveren, D.: 1997, Astron. Astrophys. 322, 116.
Van den Heuvel, E.P.J.: 1977, in Eight Texas Symposium on Relativistic Astrophysics,
Annals New York Academy Sci. 302, 482.
Van Paradijs, J.: 1995, in X-Ray Binaries, W.H.G. Lewin, J. van Paradijs and E.P.J.
van den Heuvel, (eds.), Cambridge University Press, p. 536.
Yokogawa, J., Paul, B., Ozaki, M., Nagase, F., Chakrabarty, D. And Takeshima, T.:
2000a, Astrophys. J. 539, 191.
Yokogawa, J., Torii, K., Imanishi, K. and Koyama, K.: 2000b, Publ. Astron. Soc. Japan
52, L37.
Zamanov, R. and Marti, J.: 2000, astro-ph 0005201.
Zamanov, R., Marti, J. and Marziani, P.: 2001, astro-ph 0110114.
Zamanov, R.K., Reig, P., Marti, J., Coe, M.J., Fabregat, J., Tomov, N.A. and Valchev,
T.: 2000, astro-ph 0012371.
Zi´o lkowski, J.: 1980, in Close Binary Stars: Observations and Interpretations, M.J.
Plavec, D.M. Popper and R.K. Ulrich (eds.), D. Reidel Publ. Co., Dordrecht, Hol-
land, p. 335.
Zi´o lkowski, J.: 1985, Acta Astron. 35, 185.
Zi´o lkowski, J.: 1992, Be/X-Ray Binaries, review presented at Vulcano Workshop 1992
(unpublished).
Zi´o lkowski, J.: 2001, in Exploring the Gamma-Ray Universe, A. Gimenez, V. Reglero
and C. Winkler (eds.), (Proceedings of the Fourth Integral Workshop), ESA SP-459,
p. 169.
|
astro-ph/0504416 | 2 | 0504 | 2007-09-25T06:16:05 | Remarks on the formulation of the cosmological constant/dark energy problems | [
"astro-ph",
"gr-qc",
"hep-th"
] | Associated with the cosmic acceleration are the old and new cosmological constant problems, recently put into the more general context of the dark energy problem. In broad terms, the old problem is related to an unexpected order of magnitude of this component while the new problem is related to this magnitude being of the same order of the matter energy density during the present epoch of cosmic evolution. Current plans to measure the equation of state certainly constitute an important approach; however, as we discuss, this approach is faced with serious feasibility challenges and is limited in the type of conclusive answers it could provide. Therefore, is it really too early to seek actively for new tests and approaches to these problems? In view of the difficulty of this endeavor, we argue here that a good place to start is by questioning some of the assumptions underlying the formulation of these problems and finding new ways to put this questioning to the test. First, we calculate how much fine tuning the cosmic coincidence problem represents. Next, we discuss the potential of some cosmological probes such as gravitational lensing to identify novel tests to probe dark energy questions and assumptions. Then, motivated by some theorems in General Relativity, we discuss if the full identification of the cosmological constant with vacuum energy is unquestionable. Also, we point out the relevance of experiments at the interface of astrophysics and quantum field theory, such as the Casimir effect in gravitational and cosmological contexts. We conclude that challenging some of the assumptions underlying the cosmological constant problems and putting them to the test may prove useful and necessary to make progress on these questions. (Abridged) | astro-ph | astro-ph | Remarks on the formulation of the cosmological constant/dark energy problems
7
0
0
2
p
e
S
5
2
2
v
6
1
4
4
0
5
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
Mustapha Ishak1 [*],
1 Department of Physics, The University of Texas at Dal las, Richardson, TX 75083, USA
(Dated: August 12, 2013)
Associated with the cosmic acceleration are the old and new cosmological constant problems,
recently put into the more general context of the dark energy problem. In broad terms, the old
problem is related to an unexpected order of magnitude of this component while the new problem
is related to this magnitude being of the same order of the matter energy density during the present
epoch of cosmic evolution. Current plans to measure the equation of state or density parameters
certainly constitute an important approach; however, as we discuss, this approach is faced with seri-
ous feasibility challenges and is limited in the type of conclusive answers it could provide. Therefore,
is it really too early to seek actively for new tests and approaches to these problems? In view of the
difficulty of this endeavor, we argue in this work that a good place to start is by questioning some
of the assumptions underlying the formulation of these problems and finding new ways to put this
questioning to the test. First, we calculate how much fine tuning the cosmic coincidence problem
represents. Next, we discuss the potential of some cosmological probes such as weak gravitational
lensing to identify novel tests to probe dark energy questions and assumptions and provide an exam-
ple of consistency tests. Then, motivated by some theorems in General Relativity, we discuss if the
full identification of the cosmological constant with vacuum energy is unquestionable. We discuss
some implications of the simplest solution for the principles of General Relativity. Also, we point
out the relevance of experiments at the interface of astrophysics and quantum field theory, such as
the Casimir effect in gravitational and cosmological contexts. We conclude that challenging some of
the assumptions underlying the cosmological constant problems and putting them to the test may
prove useful and necessary to make progress on these questions.
PACS numbers: 98.80.Jk,04.20.Cv
I.
INTRODUCTION
Several different types of astrophysical observations, e.g.
[1, 2, 3, 4, 5, 6, 7], have established the evidence that
the expansion of the universe entered a phase of acceleration. Associated with this acceleration is a cosmological
constant, or another dark energy component, that contributes significantly to the total energy density of the universe.
2
Cosmic acceleration and dark energy constitute one of the most important and challenging of current problems in
cosmology and other areas of physics. There are many comprehensive reviews of the cosmological constant or dark
energy, including the observational evidence for it and the problems associated with it, and we refer the reader to
some of them [8, 9, 10, 11, 12, 13, 14, 15].
Three questions related to the cosmic acceleration are encountered in the published literature, two of which are
found in different formulations or expressions. These are:
i) What is causing the cosmic acceleration? Is it a cosmological constant or a dynamical dark energy component
[16]? Is this associated with the stress-energy momentum sector or with the curvature sector of the Einstein field
equations (EFE)? Is this an indication of new physics at cosmological scales?
ii) The old cosmological constant problem: If the acceleration is caused by vacuum energy, then why is the value
measured from astrophysics so small compared to values obtained from quantum field theory calculations (this is the
puzzling smallness formulation, see for example [12])? Another formulation is: Why do all the contributions to the
effective cosmological constant term cancel each other up to a very large number of decimal places (this is the fantastic
cancelation formulation, see for example [8])? We contrast the two formulations in the next section.
iii) The new cosmological constant problem: Why is the acceleration happening during the present epoch of the
cosmic evolution? (Any earlier would have prevented structures from forming in the universe.) This is also formulated
as the cosmic coincidence: i.e. why is the dark energy density of the same order of magnitude as the matter density
during the present time?
In this paper, we argue that questioning the formulation of these problems and challanging the underlying as-
sumptions may proove useful and necessary in order to make progress on the questions. We start by discussing the
cosmic coincidence problem and then make some clarifications using some fraction calculations. Then, we calculate
some prospective constraints on the cosmological parameters of dark energy/cosmological constant and discuss the
inferred possible answers. Constraining the equation of state parameters is certainly an important approach to per-
sue, however, as we show, the level of precision required is very challanging, and yet the approach is limited in the
kind of answers it could provide. This points out the need for new approaches to dark energy/cosmological constant
problems. In particular, questioning the assumptions underlying these questions may be found useful in order to look
for new types of tests or approaches to these problems. We provide an example where cosmological probes of the
cosmic expansion and the growth rate of large-scale struture-formation can be used in order to distinguish between
cosmic acceleration due to dark energy and cosmic acceleration due to some modification to General Relativity at
cosmological scales. Next, we question the full identification of the cosmological constant with vacuum energy in
light of some theorems on the most general curvature tensor in the Einstein Field equations. Then, we discuss how
a simple intrinsic constant curvature of spacetime would fit within important interpretations of General Relativity’s
principles. Finally, we point out the relevance of experiments at the interface of particle physics and astrophysics,
such as the Casimir effect in astrophysical and cosmological contexts. A discussion and conclusion are provided in
the last section.
3
II. PRELIMINARIES
A. Notation
We recall here only some preliminary equations and definitions necessary for the clarity of the paper. We refer the
reader to review papers, see e.g. [8, 9, 10, 11, 12, 13, 14] and text books, see e.g. [17, 18, 19, 20]. The Einstein Field
Equations (EFE) with a cosmological constant Λ read
where κ ≡ 8πG and
Gαβ + Λgαβ = κTαβ
Gαβ ≡ Rαβ −
1
2
gαβ R
(1)
(2)
is the Einstein tensor, Rαβ , R and gαβ are the Ricci tensor, the Ricci curvature scalar and the metric tensor respectively.
For a perfect fluid, the energy-momentum tensor is given by
Tαβ = (ρ + p)uαuβ + pgαβ
(3)
where uα is the four velocity vector and ρ and p are the energy-density and isotropic pressure relative to uα . With
global isotropy, when Tαβ = 0, the EFE (1) admit de Sitter space (for Λ > 0) as a solution, for which a line element
is given by (A5) in the appendix.
Motivated by current observations, e.g.
[1, 2, 3, 4, 5, 6, 7], and for the sake of simplicity, let us consider the
concordance spatially flat universe with a positive cosmological constant Λ. The EFE (1) with a dust source can be
solved explicitly to give
Λ (cid:17)1/3 h sinh(
a(t) = (cid:16) 3C
√3Λ
2
t)i2/3
(4)
where C ≡ 8πρa3/3 is a constant; and the spacetime line element is given by
ds2 = −dt2 + a2 (t)(dr2 + r2 dΩ2 ).
4
(5)
where dΩ2 = (dθ2 + sin2 θdφ2 ). At early stages, the universe is matter dominated and a(t) = ( 9C
4 t2 )1/3 . At late stages,
the universe is dominated by the cosmological constant and enters a de Sitter phase with a(t) = ( 3C
4Λ )1/3 exp pΛ/3t.
We plot the curvature evolution and profile of these spacetimes in the appendix.
Now, to introduce the vacuum energy density, let us consider a scalar field with Lagrangian density
and energy momentum tensor
Lφ = −
1
2
gαβ ∂αφ∂β φ − V (φ)
T φ
αβ = ∂αφ∂β φ −
1
2
(g δγ ∂δ φ∂γ φ)gαβ − V (φ)gαβ
(6)
(7)
The lowest energy density of the field configuration is when the kinetic (or gradient) term vanishes and the potential
is at the minimum V (φmin ). The energy momentum (7) reduces to
This form of T vac
αβ is also the only Lorentz-invariant form for the vacuum energy.
T vac
αβ = −V (φmin )gαβ = −ρvacgαβ
(8)
B. The old cosmological constant problem
The common identification of the cosmological constant with the vacuum energy density is based on the mathemat-
ical equivalence of the vacuum energy momentum tensor (8) and the cosmological constant term on the LHS of the
EFE (1). Also, writing a lagrangian density that includes gravitational terms and terms from quantum field theory
leads to the temptation to combine some of these terms; however, one should bear in mind that we don’t have such
a unified theory yet.
Now, if one considers that a geometrical cosmological constant term is an integral part of the EFE then the old
cosmological constant problem can be expressed as (see for example [8, 13]): Why do all the contributions from
vacuum energy density fantastically cancel with the geometrical Λ term? This formulation of the problem seems to
be less often recalled in some of the recent literature. To see this quantitatively, one can combine equations (1) and
(8) to write
Λef f ective
8πG
=
Λ
8πG
+ ρvac
(9)
where from effective quantum field theory
5
1
2 X ¯hω .
Now, if one considers only quantum field fluctuations cut off at particle energies of 100 GeV (this means, we consider
ρvac =
(10)
only the well known physics of the standard model), one can write (with ¯h = c = 1) ρvac ∼ (100GeV )4 . However,
from astrophysical observations (Λef f ective /8πG) is found to be comparable to the critical density, i.e. ∼ 10−48 GeV 4
which means that the two terms in the RHS of (9) must cancel to 56 decimal places.
On the other hand, from the identification of the cosmological constant with the vacuum energy, the old Λ problem
becomes: Why is the vacuum energy measured from astrophysics (∼ 10−48 GeV 4 ) so small compared to the value of
the vacuum energy estimated from quantum field theory (∼ (100GeV )4 )? Of course, the situation is made even worse
by taking the GUT or Planck scales.
This full identification may bring some limitations of its own to the old cosmological constant problem and is
questioned later in this work.
III. HOW MUCH FINE-TUNING DOES THE COSMIC COINCIDENCE REPRESENT? (THE NEW
”PROBLEM”)
The cosmic coincidence problem is usually stated as why is the cosmic acceleration recent in the cosmic history
of the universe (why “now”?) It is also discussed in terms of a fine tuning problem: i.e. why is the matter energy
density of the same order of magnitude as the dark energy density at the present epoch? Related to this question is
the fact that if the dark energy density was much larger then what is measured it would have dominated over the
matter energy density much earlier and prevented structures from forming in the universe. It is important to discuss
the cosmic coincidence because it is an argument that is used to motivate the search for dynamical components in
order to be able to explain the coincidence. We will try to quantify some of these statements here.
The matter energy density is related to the redshift by
ρm (z ) = (1 + z )3ρtoday
m (z = 0),
and in a ΛCDM universe
t(a) = Z a
0
where a = 1
1+z is normalized to 1 today.
da′
a′H (a′ )
= Z a
0
da′
a′H0√ΩΛ + Ωma′−3
(11)
(12)
Now, from the concordance model Ωm ≈ 0.3 and ΩΛ ≈ 0.7 today and it follows that
ρΛ =
ΩΛ
Ωm
ρtoday
m
≈ 2.33ρtoday
m .
6
(13)
At the transition from deceleration to acceleration the dark energy density is half the matter energy density and
Ωm (cid:19)1/3
1 + ztrans = (cid:18) 2ΩΛ
(see for a first measurement of this [1]: 1 + ztrans = 1.46 ± 0.13).
It is a clarifying exercise to evaluate the age of the universe t at this transition. Using (12) gives a transition age
≈ 1.67
(14)
of ≈ 7.14 billion years (we use H0 = 72km/s/M pc [3]), showing that the transition is not that recent in the time
history of the universe. Moreover, ρm and ρΛ have been within the same order of magnitude starting roughly from
1 + z ≈ 2.85 (with ρm ≈ 9.94ρΛ) and that corresponds to an age of the universe of ≈ 3.38 billion years. i.e. ρm and
ρΛ have been within the same order of magnitude for at least the last 10.32 billions years of the time history of the
universe.
In order to evaluate how much fine-tuning the cosmic coincidence represents, one could evaluate some informative
fractions using the transition time or the period during which dark energy has been of the same order of magnitude
as the matter energy density, compared to the whole history of the universe. These fractions provide the percent
chances for an observer randomly put in the time history of the universe to have the matter energy density and the
dark energy density being of the same order of magnitude. We will compare energy densities using a logarithmic scale
of time, so the results will be similar to those using a logarithmic redshift scale.
On a logarithmic scale, things can appear very significant on plots; however, the fractions are actually of a few
percent:
and
ln ttoday − ln ttransition
ln ttoday − ln ti
≈ 2%,
ln ttoday − ln tsame−order
ln ttoday − ln ti
≈ 3%
(15)
(16)
ln ttoday − ln tsame−order
ln ttoday − ln tP lanck ≈ 1%
We used here ti = 1 sec as we considered observational constraints from the Big Bang nucleosynthesis and CMB
(17)
results to trace back in time the universe up to 1 second at the e+ e− annihilation. We also use the Planck time
tP lanck = 10−44 sec and find that the fraction is ≈ 1%.
insignificant:
(Of course on a linear scale of time the coincidence is
7
and
ttoday − ttransition
ttoday − ti
≈ 48%
ttoday − tsame−order
ttoday − ti
≈ 75%.)
(18)
(19)
In summary, even on a logarithmic scale the fractions remain of the order of a few percent. This is significant but
not totally unusual in physics and astrophysics.
Consequently, while it remains perhaps a motivated problem to seek models that could naturally explain some of
these small fractions, this fine tuning should not constitute a barrier to reject models that address successfully the
other cosmological constant problems.
IV. OBSERVATIONAL CONSTRAINTS ON THE EQUATION OF STATE PARAMETERS AND THE
COSMOLOGICAL CONSTANT/DARK ENERGY QUESTIONS
An important approach to the cosmological constant questions is to parameterize the dark energy using its equation
of state, and then to use different cosmological data sets in order to determine these parameters. One can thus infer
answers to the cosmological constant/dark energy questions from the values found for these parameters. As we will
delineate further, some results will be more conclusive than others. It is essential to combine multiple complementary
cosmological probes and techniques in order to break degeneracies between the cosmological parameters including the
dark energy parameter space. A powerful combination for constraining the dark energy parameters is the Cosmic
Microwave Background Radiation (CMB) plus distance measurements from the type Ia Supernovae (SNIe) and Weak
Gravitational Lensing (WL) also called cosmic shear, e.g. see [50, 51, 52]. These probes provide orthogonal constraints
that, when combined, reduce significantly the uncertainties on the individual parameters. Also, constraints from
clusters of galaxies and baryons oscillations were shown to be good probes of the dark energy equation of state, see
e.g. [21, 22]. Interestingly, WL allows one to do tomography by separating the source-galaxies into redshift bins in
order to obtain further improvements on the parameter constraints, see e.g. [23, 24, 25]. In this analysis, we consider
future constraints from the CMB+SNIe+WL combination plus WL tomography with various numbers of redshift
bins.
8
TABLE I: Cosmological model. We use fiducial values for the parameters from recent results from the WMAP (3-years data
results) + Large scale Structure + Supernova data as listed in table II with the respective references (we use for the equation
of state the cosmological constant parameter values). We assume a spatially flat universe with Ωm + ΩΛ = 1. This fixes Ωm
and H0 as functions of the basic parameters. We do not include massive neutrinos, or primordial isocurvature perturbations.
Symbol
Description
Fiducial value Probe-1 Probe-2 Probe-3
Ωm h2
ΩΛ
physical matter density
0.1259
fraction of the critical density in a dark energy component
0.745
w0 ,w1 (or wa )
equation of state parameters
σ lin
8
amplitude of linear fluctuations
-1, 0
0.712
ns (k0 = 0.05h/Mpc) spectral index of the primordial scalar power spectrum at k0
0.946
αs
zp
ζs
ζr
Ωb h2
τ
T /S
running of the primordial scalar power spectrum
-0.06
the characteristic redshift of source galaxies for lensing
0.76, 1.12
absolute calibration parameter for WL power spectrum [42, 43] 0.0
relative calibration parameter for WL power spectrum [42, 43] 0.0
physical baryon density
optical depth to reionization
scalar-tensor fluctuation ratio
0.0448
0.08
0.2
WL
WL
WL
WL
WL
WL
WL
WL
WL
CMB
SNIe
CMB
SNIe
CMB
SNIe
CMB
CMB
CMB
CMB
CMB
CMB
A. methodology
In this analysis, we consider a cosmological model with 13 parameters that take the fiducial values in table I based
on recent 3-years results from the Wilkinson Microwave Anisotropy Probe (WMAP) [2, 40, 41] combined with large
scale structure and supernova data as given in table II below along with the respective references (except for the
dark energy equation of state, where we use the cosmological constant values). We consider in this analysis, the two
standard parameterizations for the dark energy equation of state w = P /ρ given by
• w(z ) = w0 + w1 z if z < 1 and w(z ) = w0 + w1 otherwise, e.g. [1, 35], and
• w(a) = w0 + wa (1 − a), e.g. [36, 37].
The dark energy density as a function of redshift is thus given by
Q(z ) ≡ ρde (z )/ρde (0) = exp 3 Z z
0
1 + w(z ′ ))dz ′
1 + z ′
.
(20)
TABLE II: Data sets used in [41] to derive the constraints on equation of state parameter w using CMB, large scale structure
and supernova data (table 9 in [41]). See also the NASA’s data center for Cosmic Microwave Background (CMB) research,
9
LAMBDA at http://lambda.gsfc.nasa.gov/ .
data set/experiment
references
WMAP 3-years results
(Spergel, et al., 2006) [41]
2dF Galaxy Redshift Survey
(Cole, et al., 2005) [26]
BOOMERanG + ACBAR
(Montroy, et al., 2005, Kuo, et.al., 2004) [28, 29]
CBI + VSA
(Readhead, et al., 2004, Dickinson, et.al., 2004) [30, 31]
Sloan Digital Sky Survey
(Tegmark, et al., 2004; Eisenstein, et.al., 2005) [27, 32]
Supernova ”Gold Sample”
(Riess, et.al., 2004) [34]
Supernova Legacy Survey (SNLS) (Astier, et al., 2006) [33]
Despite its simplicity, statistical inference theory using Fisher matrices was proven to be a very efficient tool to
calculate constraints that will be obtained from future planned or proposed experiments. As an example of its
efficiency, one could look at the good predictions on the parameter uncertainties forecasted, for instance, by [23, 38]
for WMAP and compare them to the real WMAP results, obtained years later [2]. We describe here only the basic
ideas of the methods that we use in the current paper but provide references [45] where the full detail can be found.
We use the standard approach to calculate the statistical error on a given parameter pα by using the combination:
σ2 (pα ) ≈ [(FCM B + FW L + FSN e + Π)−1 ]αα ,
(21)
where FCM B , FW L and FSN e are the Fisher matrices from CMB, weak lensing, and supernovae respectively, and Π
is the prior matrix. As discussed in some detail in [23, 25], we calculate FW L using
Fαβ =
∂Pκ
∂ pα
∂Pκ
∂ pβ ;
ℓmax
Xℓ=ℓmin
where the lensing convergence power spectrum is given by [46, 47, 48]:
m Z χH
H 4
0 Ω2
0
P3D (cid:18)
1
(∆Pκ )2
g 2 (χ)
a2 (χ)
P κ
l =
9
4
(22)
(23)
, χ(cid:19) dχ.
l
sinK (χ)
where P3D is the 3D nonlinear power spectrum of the matter density fluctuation, δ ; a(χ) is the scale factor; and
sinK χ = K −1/2 sin(K 1/2χ) is the comoving angular diameter distance to χ (for the spatially flat universe used in this
10
analysis, this reduces to χ). The weighting function g (χ) is the source-averaged distance ratio given by
g (χ) = Z χH
sinK (χ′ − χ)
sinK (χ′ )
χ
where n(χ(z )) is the source redshift distribution normalized by R dz n(z ) = 1. The uncertainty in the observed lensing
spectrum is given by: [47, 48]
n(χ′ )
dχ′ ,
(24)
is the intrinsic ellipticity of galaxies. For this analysis, we consider a lensing survey with 10%
∆Pκ (ℓ) = s
2 (cid:11)¯n ! ,
(2ℓ + 1)fsky Pκ (ℓ) + (cid:10)γint
2
where fsky = Θ2π/129600 is the fraction of the sky covered by the gravitational lensing survey of dimension Θ in
int(cid:11)1/2
degrees, and (cid:10)γ 2
sky coverage, a median redshift of 1, an average galaxy number density of ¯n = 30 gal/arcmin2 , and intrinsic ellipticities
int (cid:11)1/2
(cid:10)γ 2
consider a deeper survey (space-based like) with fsky = 0.10, a median redshift of roughly 1.5, ¯n = 100 gal/arcmin2 ,
int (cid:11)1/2
and (cid:10)γ 2
≈ 0.25. We calculate constraints from 10 tomographic bins using this survey. For both surveys, we use
ℓmax = 3000 to keep the assumption of a Gaussian shear field valid.
= 0.4. We calculate the constraints on the parameters using five tomographic bins for this survey. We also
(25)
Weak lensing has been recognized as a very powerful probe of dark energy parameters, however, several systematic
effects have been identified so far, see [50] and references therein for an overview. In this analysis, we included the
effect of the shear calibration bias [43, 49, 53, 54, 55, 56] on our results by marginalizing over its parameters. Because
of this bias, the shear is systematically over or under-estimated by a multiplicative factor, and results in an overall
rescaling of the shear power spectrum. Following the parameterization discussed in [42], we used the absolute power
calibration parameter ζs and the relative calibration parameter ζr between two redshift bins. Another systematic
effect that we considered in the analysis is the incomplete knowledge of the source redshift distribution [44, 57]. A
remedy to this poor knowledge of the redshift distribution using spectroscopic redshift has been explored recently
in [44]. In the present analysis we marginalize over the redshift bias by including the characteristic redshift of the
distribution as a systematic parameter zs and we assume a reasonable prior of 0.05 on this parameter.
In order to calculate the supernova Fisher matrix, FSN e , we use (see, e.g. [58, 59])
Fαβ =
N
Xi=1
where DL ≡ H0dL /c is the dimensionless luminosity distance to a supernova, given in a spatially flat model by
DL (z ) = (1 + z ) Z z
0
1
σm (DL,i )2
∂DL,i
∂ pβ .
∂DL,i
∂ pα
1
p(1 − ΩΛ )(1 + z ′)3 + ΩΛQ(z ′ )
dz ′ ,
(26)
(27)
11
where Q(z ) is as defined in Eq.(20). We recall that the SN Ia apparent magnitude as a function of redshift is given
by m(z ) = 5 log10 (DL (z )) + M, where M depends as the absolute magnitude of type Ia supernovae as well as on
the Hubble parameter H0 . As usual, we treat M as a nuisance parameter. We use two sets of 2000 SNe Ia uniformly
distributed with zmax = 0.8 and zmax = 1.5. It is important to briefly note here that there are systematic uncertainties
associated with supernova searches: these include luminosity evolution, gravitational lensing and dust; see, e.g. [60]
and references therein.
In order to partly include the effect of these systematics and the effect of the supernova
peculiar velocity uncertainty [62], we follow [60, 61] and use the following quadrature for the effective magnitude
uncertainty
m = sσ2
ln(10)cz (cid:19)2
m + (cid:18) 5σv
σef f
where, σv = 500km/sec is the peculiar velocity, and δm is a floor uncertainty in each bin [60, 61]. The quadrature
+ N{per bin} δ2
m
(28)
relation (28) assures that there is an uncertainty floor set by the systematic limit δm so that the overall uncertainty
per bin cannot be reduced to arbitrarily low values by adding more supernovae.
Finally, for the CMB and lensing tomography, we use the generalized form of the Fisher matrix above (e.g. see
[23]) as
ℓmax
∂ pβ (cid:19) ,
(ℓ + 1/2)fskyTr (cid:18)C−1
∂Cℓ
Xℓmin
ℓ
ℓ + N X Y
ℓ = C X Y
where Cℓ is the covariance matrix of the multipole moments of the observables C X Y
ℓ
∂Cℓ
∂ pα
Fαβ =
C−1
ℓ
(29)
with N X Y
ℓ
being
the power spectrum of the noise in the measurement. Here X Y takes the values κκ for lensing tomography spectra
and cross-spectra, and T T , T E , and EE for CMB spectra. We consider constraints from 1-year data from the Planck
satellite. Our findigs are discussed in the next section.
B. Summary of results and implications for the cosmological constant/dark energy questions
Our results, summarized in table III, show that when WL with multiple-bins tomography is added to the CMB+SN
combination, the uncertainties on the equation of state parameter w0 reduce by roughly a factor of 3 and the uncer-
tainties on w1 and wa reduce by roughly factors of 3 or 4 at least. Also, we note that the uncertainties on the set
{w0 , w1 } are different from those of the set {w0 , wa } showing that the uncertainties are parameterization dependent.
The deeper lensing survey, which allows one to implement 10-bins tomography, provides an additional factor of 2
improvement on w1 and wa compared with the 5-bins tomography results. We find that in order to bring the uncer-
tainties on both parameters to the order of several percent, very ambitious surveys are required. Current observations
12
TABLE III: Dark energy parameter constraints from various combinations of CMB, Weak Gravitational Lensing and Supernova
future surveys. The constraints are (1σ uncertainties) on the two dark energy parameterizations given in section IV A. The
uncertainties are calculated using combinations of 1-year data from Planck, 2000 uniformly distributed supernovae with zmax =
0.8, 1.5, a lensing survey with 5-bins tomography, and a very deep lensing survey with 10-bins tomography. The results are
presented for the dark energy parameters {w0 , w1} and {w0 , wa}. The systematic effects discussed in section IV A are included
in the calculations.
Simulated Experiment/Survey
Parameterization-1 parameterization-2
σ (w0 )
σ (w1 )
σ (w0 )
σ (wa )
1-year data from PLANCK + 2000 SN with zmax = 0.8
0.11
0.26
0.13
0.47
1-year data from PLANCK + 2000 SN with zmax = 0.8
0.039
0.092
0.033
0.12
+ WL survey with 5-bins tomography, zmed = 1.0 and fsky = 0.10
1-year data from PLANCK + 2000 SN with zmax = 1.5
0.08
0.20
0.10
0.35
1-year data from PLANCK + 2000 SN with zmax = 1.5
0.022
0.043
0.024
0.057
+ WL deep survey with 10-bins tomography, zmed = 1.5 and fsky = 0.10
are only able to constrain, with some significance, models with a constant equation of state, i.e. one single parameter
with no redshift dependence. For example, the WMAP team recently combined constraints from currently available
CMB, large scale structure, and supernova data sets (see table II) and obtained w = −0.926+0.051
−0.057 (table 9 in [41]).
Our table III shows the 1-sigma uncertainties on the equations of state parameters, and in order to consider the
2 and 3-sigma constraints, one has to multiply the values found by factors of two and three respectively. Therefore,
one can see that even when ambitious surveys are considered the remaining uncertainty is still too large to constrain
significantly multiple-parameters dark energy models. However, It will be possible to exclude some proposed models
with significant deviations from the cosmological constant parameters, namely w0 = −1, and w1 = 0. These include
for example trackers models [63] and some SUGRA inspired models [64] with for example w0 = −0.8 and w1 = 0.3.
The most decisive answer will be if the data can show conclusively that dark energy is not a cosmological constant. A
very suggestive but less decisive answer will be to show that the dark energy parameters are those of a cosmological
constant to a very high level of precision (a few percent, perhaps). But of course, the degeneracy in this case will
remain and other tests, beyond the equation of state approach, will be needed.
In this case, other tests are also
necessary because the problems of the cosmological constant are just re-affirmed. Finally, an important question that
need to be addressed in all cases is whether the equation of state obtained is a true or forced equation of state as
we explore further below. We discuss in the next sections some possible directions of such tests and illustrate one
promising test using cosmological probes.
13
V. RECONSIDERING AND TESTING SOME OF THE ASSUMPTIONS ABOUT THE OLD
COSMOLOGICAL CONSTANT AND DARK ENERGY PROBLEMS
In this section we propose some examples on how one could question some of the assumptions made about the
cosmological constant/dark energy problems and, in some cases, how to put this questioning to the test.
A. Cosmological tests beyond the equation of state approach: The expansion history versus the growth rate
of large scale structure
The approach of the equation of state is certainly an important one. However, an important question that remains
after some dark energy parameters are obtained from analyzing observational data is as follows. Is this an effective
equation of state of some dark energy component in the Einstein’s equations, or is this just a forced equation of state
obtained from fitting dark energy models on the top of some modified gravity at cosmological scales? New tests are
necessary in order to address this question.
Indeed, of great importance are innovative ways of using current and future astrophysical observations that could
distinguish between acceleration models beyond the equation of state.
In other words, for the same degenerate
effective equation of state, these novel tests could distinguish between dynamical dark energy models, a geometrical
cosmological constant, and acceleration due to some modification to the gravity sector, and thus will allow one to test
some of the important basic assumptions. Some cosmological probes such as weak gravitational lensing (for reviews,
see [50, 51, 52, 56] and references therein) and clusters of galaxies (see for example [21, 22] and references therein)
are very rich tools and very promising for identifying this type of test because they provide more than one way to
constrain dark energy or cosmic acceleration. Both probes can capture the effect of dark energy on the expansion
history and also its effect on the growth rate of large-scale structure (the rate at which clusters and super clusters
of galaxies form over the history of the universe). Interestingly, this can be used to identify consistency checks to
test the dark energy beyond the equation of state degeneracy. In particular, this could allow one to test dark energy
14
models based on new particles and fields versus cosmic acceleration due to some modification in the curvature sector
of the EFE as suggested in some recent studies, see e.g. [65, 74, 75], recent review [15] and references therein. Most
importantly, these kinds of consistency checks could be used to test some of the assumptions discussed in this paper,
namely on the origin of the intrinsic constant curvature of spacetime. Another test beyond the equation of state and
based on the dark energy potential was discussed in [76].
In this paper, we explore an example to demonstrate that tests that go beyond the equation of state degeneracy
are possible. An important point for this test is that cosmic acceleration affects cosmology in two ways: 1) It affects
the expansion history of the universe by speeding it up, 2) It affects the growth rate of large scale structure in the
universe by suppressing it. The idea explored is that, for dark energy models, these two effects must be consistent
one with another because their respective functions are mathematically related by General Relativity equations. The
presence of significant inconsistencies between the expansion Hubble function and the growth rate function could be
the signature of some modified gravity at cosmological scales as we will demonstrate.
In order to illustrate how the test works, we will need to use a viable modified gravity model. We choose a
model proposed by Dvali, Gabadadze and Porrati (DGP) [65] where the cosmic acceleration is due to the effect of an
extra large dimension modifying gravity at cosmological scales. This DGP model is motivated by higher dimensional
physics and is not ruled out by current astrophysical observations [66, 69, 79]. We have no particular interest in the
phenomenology and precise testing of the viability of this model. We are only interested to use it as an example in
order to illustrate the test considered. We refer the interested reader to some studies dedicated to the DGP model
phenomenology [69, 70, 71, 81].
We provide here a very brief description of this model but again refer the reader to [65, 67] for a full description.
The action for this five-dimensional theory is [65, 67]
(4) Z d4xp−g(4)R(4) + Smatter ,
(5) Z d4x dyp−g(5)R(5) +
M 2
M 3
where the subscripts 4 and 5 denote quantities on the brane and in the bulk, respectively; M(5) is the five dimensional
S(5) =
(30)
1
2
1
2
reduced Planck mass; M(4) = 2.4 × 1018GeV is the four dimensional effective reduced Planck mass; R and g are
the Ricci scalar and the determinant of the metric, respectively. The first and second terms on the right hand side
describe the bulk and the brane, respectively, while Smatter is the action for matter confined to the brane. The two
(5)/2 and M 2
different prefactors M 3
(4)/2 in front of the bulk and brane actions give rise to a characteristic length scale
[77], rc = M 2
(4)/2M 3
(5) . If M(5) is much less than M(4) , then the brane terms in the action above will dominate over
the bulk terms on scales much smaller than rc , and gravity will appear four dimensional. On scales larger than rc ,
the full five dimensional physics will be recovered, and the gravitational force law will revert to its five dimensional
1/r3 form. This is usually discussed in terms of gravity leakage into an extra dimension. Ref. [66] shows that tuning
M(5) to about 10 − 100MeV, implying rc ∼ H −1
0 , is consistent with cosmological data. They have also been discussed
in [79]. Low redshift cosmology in DGP brane worlds was studied in [67, 77]. Following [67], one could define the
15
effective energy density ρrc ≡
3
c ) so that Friedmann’s first equation becomes
(32πGr2
8πG
3 (cid:0)pρ + ρrc + √ρrc (cid:1)2
c H −2
where Ωrc ≡ 1
4 r−2
0 . We focus here on a flat universe (k = 0) containing only nonrelativistic matter, such as
(cid:1)2
baryons and cold dark matter, in which Ωrc = (cid:0) 1−Ωm
2
dimension, on large length scales, becomes a substitute for dark energy.
. In this model, the gravitational “leakage” into the fifth
k
a2 =
H 2
DGP +
(31)
.
Now, for the standard general relativistic (GR) cosmological model, FLRW, with zero spatial curvature (k=0), and
a dark energy component (DE), the expansion history is expressed by the Hubble function and is given by
HGR+DE (z ) = H0pΩm (1 + z )3 + (1 − Ωm )Q(z ).
And the growth rate of large scale structure G(a=1/(1+z)) is given by integration of the differential equation, [72, 73],
1 + X (a) (cid:21) G′GR+DE
G′′GR+DE + (cid:20) 7
w(a)
2 −
a
where ′ ≡ d/da, G = D/a is the normalized growth rate, Q(a) is as given by equation 20, and
Ωm
(1 − Ωm )a3Q(a)
1 − w(a)
1 + X (a)
GGR+DE
a2
X (a) =
= 0,
(33)
(34)
(32)
3
2
+
3
2
.
On the other part, for the spatially flat DGP model, the expansion Hubble function is given by
(1 − Ωm ) + r 1
HDGP (z ) = H0(cid:16) 1
(1 − Ωm )2 + Ωm (1 + z )3(cid:17).
2
4
and the suppression of the growth rate function is given by, [79, 80]
where
δDGP + 2HDGP
δDGP − 4πGρ(cid:0)1 +
1
3β (cid:1)δDGP = 0
(35)
(36)
HDGP
β = 1 − 2rcHDGP (cid:16)1 +
DGP (cid:17)
3H 2
We illustrate in Fig.1, Hubble expansion functions and growth rate functions for dark energy and DGP models
(37)
including the cosmological constant model, ΛCDM. The figure shows that compared to a ΛCDM model with the same
16
24
23.5
23
22.5
22
M
-
m
LCDM: w0=-1, w1=0(cid:10)
SUGRA: w0=-0.8, w1=0.3(cid:10)
DGP model with Omegam=0.20
DGP model with Omegam=0.27
21.5
0.5
0.6
0.7
0.8
1
1.1
1.2
1.3
0.9
z
a
/
δ
h
t
w
o
r
G
1
0.95
0.9
0.85
0.8
0.75
0.7
0.65
0.6
LCDM: w0=-1, w’=0(cid:10)
SUGRA: w0=-0.8, w’=0.3(cid:10)
DGP model with Omegam=0.20
DGP model with Omegam=0.27
0.2
0.3
0.4
0.5
0.7
0.8
0.9
1
0.6
a
FIG. 1: Supernova Hubble diagrams (Left) and Growth rate functions for dark energy and DGP models. Note that the ΛCDM
model (red solid line) and the Ωm = 0.20 DGP model (blue dotted) have degenerate Hubble diagrams, but different growth
rates. The degeneracy of the Hubble diagrams is even stronger for SUGRA (green dashed) and Ωm = 0.27 DGP (black double
dotted) models. Interestingly, the growth rate in the Ωm = 0.27 DGP model is suppressed with respect to that in the ΛCDM
model, which has the same Ωm .
matter density, a DGP model has a distinct suppression of the growth rate. Also, Fig.1b displays how the degenerate
models of Fig.1a show distinct growth rate functions.
The basic idea for the test explored is that equations (32) and (33) must be mathematically consistent one with
another via General Relativity. Similarly, equations (35) and (36) must be consistent one with another via DGP
theory. A consistency cross-check of these two functions constitute a test for the fundamental underlying theory.
In order to apply the test discussed below, we assume the availability of a sample of NSN e = 2000 type Ia
supernova, evenly distributed in redshift between zM in = 0 and zM ax = 1.7, with a magnitude uncertainty per
supernova of σm = 0.2. As discussed in section IV A, we also include systematic effects using the quadrature (28).
We also assume the availability of measurements of the growth rate in twenty bins evenly spaced in the scale factor a,
between amin = 0.25 and amax = 1 and an uncertainty of σG (a)/G(a) = 0.02. In order to break the usual degeneracy
between Ωm and the equation of state parameters, we add to the SN Ia data and to the growth rate measurements
a constraint from the CMB shift parameter. The CMB shift parameter for a spatially flat universe is given by
m R zC M B
R = Ω1/2
0
reports that the value of R from current data is Robs = 1.716 ± 0.062.
In order to demonstrate the working of the consistency check we proceed as follows: We assume that the true
dz
H/H0 (z) , where zCM B = 1089 is the redshift of the surface of last scattering, given in [2]. Ref. [82]
(cid:10)
(cid:10)
(cid:10)
(cid:10)
17
cosmology is described by a DGP model and simulate the expansion and growth data using a fiducial DGP model.
Then we ask what contradictions arise when the data are instead analyzed based on the assumption of a dark energy
model (as mentioned earlier, we are not particularly interested in the DGP cosmology here but we want to use it as
an example to illustrate the procedure.)
Because we will generate the data using the DGP model, the consistency relation from General Relativity between
the expansion history and the growth rate of large scale structure will be broken. The dark energy equation of state
wexp (z ) which best fits measurements of the expansion will not be consistent with the equation of state wgrowth (z )
which best fits measurements of the growth.
The methods and steps we use are as follows:
i) We use a fiducial DGP model with Ωm = 0.27 and simulate the data for the expansion and the growth rate.
We generate supernova magnitudes, a growth rate function, and the CMB shift parameter
ii) We use a χ2 minimization method in order to find the best fit dark energy model to the expansion. We obtain
the best fit model to the supernova magnitudes and the CMB shift parameter with a first dark energy parameter
space {Ωde , w0 , w1 }. The χ2 minimization method was discussed in detail in [35] and was shown to give similar
results to those from Monte Carlo Markov Chain method although the χ2 minimization can have a better handle
on degeneracies [35])
iii) We use the χ2 minimization in order to find the best fit dark energy model to measurements of the growth
rate of large scale structure. We determine the best fit model to the growth rate function and the CMB shift
parameter and obtain a second dark energy parameter space {Ω′de , w′0 , w′1}
iv) Next, we use a standard Fisher matrix approach to calculate the confidence regions (or χ2 contours) around the
two best fit dark energy models (this standard procedure is discussed in detail in references [45])
v) Then, we compare the allowed regions in the {Ωde , w0 , w1 } parameter space between the two data combinations
in order to look for inconsistencies
Our results, in figure 2, show that the two allowed regions in the dark energy parameter space are significantly
different. This signals the expected inconsistency between the expansion history and the growth rate of large scale
structure. The source of the inconsistency is from point (i) where the data was generated using a DGP model,
i.e. from our hypothesis that the true cosmology is that of a modified gravity DGP model. Thus, the inconsistency
18
1
w
1
0.8
0.6
0.4
0.2
0
-0.2
-0.4
-0.6
-0.8
SN + CMB Shift (1,2 sigma)
GF + CMB Shift (1,2 sigma)
Omega_de
0.9
0.88
0.86
0.84
0.82
0.8
0.78
0.76
–1.8
–1.6
–1.4
–1.2
w_0
–1
–0.8
–0.6
–0.6 –0.4 –0.2
0
0.4
0.2
w_1
0.8
0.6
1
-1.8
-1.6
-1.4
-1.2
w0
-1
-0.8
-0.6
-0.4
FIG. 2: LEFT(2D): Best fit equations of state: Solid contours are for fits to SN Ia simulated data and the CMB shift parameter
R, while dashed contours are for fits to the growth-rate simulated data and the CMB shift parameter. RIGHT(3D): Best fit
Dark Energy parameter spaces (Ωde , w0 , w1 ). The ellipsoid to the right of the 3D-figure is for fit to SN Ia simulated data and
the CMB shift parameter, while the ellipsoid to the left (3D-figure) is for fits to the growth rate simulated data and the CMB
shift parameter data. For both figures, the significant difference (inconsistency) between the parameter spaces found using the
two combinations is due to the DGP model assumed by hypothesis and used to simulate the data. The inconsistency is thus
an indication that cosmic acceleration in this case is due to modified gravity at cosmological scales rather then a Dark Energy
component.
constitute an observational detection of the assumed underlying modified gravity model. And finding two significantly
different equations of state implies that these are not true but forced ones.
The test is based on the comparison of measurements of the expansion history and measurements of the growth
rate of large scale structure and shows that we can go beyond the equation of state analysis. We demonstrated here
the working of consistency tests based on this comparison and provided a preliminary implemention in [83]. Other
works that explored the same idea include [84, 85]. Future work is needed in order to make these tests more robust
and generic: e.g. to consider other dark energy models (with couplings, unusual sound speeds), other modified gravity
models, and comparison with systematic effects of the probes. Another interesting approach was discussed in reference
[86] where the authors considered signatures of quintessence models and their extension to scalar-tensor gravity on
weak gravitational lensing observables. They found that some models can let an imprint of ten percent on lensing
observables. The important point from these examples and others is that cosmological observations that can probe
the growth of cosmological perturbations are promising tools to learn about the acceleration of the universe beyond
the effective equation of state degeneracy.
19
B. The Weyl-Lovelock theorem as an argument against the full identification of the cosmological constant
with vacuum energy
Based on the following theorems, one could argue against the full identification of the cosmological constant with
vacuum energy. Indeed, Cartan [87], Weyl [88] and Vermeil [89] proved different theorems showing that the only
tensor of valency two, Aαβ , that is:
a) constructed from the metric tensor gαβ and its first two partial derivatives, gαβ ,γ and gαβ ,γ δ ,
b) divergence free, i.e. Aαβ
;β = 0,
c) symmetric, i.e. Aαβ = Aβα ,
d) linear in the second derivatives of gαβ ,
is
Aαβ = c1Gαβ + c2gαβ
(38)
where c1 and c2 are constants and Gαβ is the Einstein tensor (2). Lovelock [90, 91] showed that conditions c) and d)
are superfluous when the spacetime dimension is 4. (Note that when Aαβ is put in the EFE, c1 is absorbed in the κ
factor.)
In Refs.
[88] and [87], it was first proven that the most general curvature tensor Aαβ is a linear combination of
Rαβ , Rgαβ and gαβ , i.e. of the form
Aαβ = aRαβ + bRgαβ + cgαβ ,
(39)
where a, b, and c are constants. Then values a = 1 and b = − 1
2 are derived from the divergence free condition
(conservation law).
Consequently, one is tempted to take the standpoint that unless one is guided by some physical laws or measure-
ments, setting the constant c (i.e. Λ) in equation (39) to any particular value, including zero is unjustifiable.
Imposing a priori a particular value on Λ (for instance zero) is perhaps making the same mistake Einstein did by
putting the particular value
ΛE instein =
4
9C 2
(40)
(for a closed static universe) where C is as defined previously, after equation (4).
Thus, this suggests that a geometrical constant Λ-term in the EFE (1) is part of the equations on its own right
with no reference to any energy momentum tensor. This could be used as an argument not in favor of the exact
identification of the geometrical cosmological constant with vacuum energy.
20
C. What are the implications of the simplest solution in view of principles of General Relativity?
As discussed in the previous sections, some of the assumptions underlying the cosmological constant problems are
not unquestionable and it is important to find ways to challenge them and put them to the test.
Needless to recall, the simplest solution can arise from, first, abandoning the assumed identification of the cosmo-
logical constant with vacuum energy (based on the theorems discussed in the previous sub-section V B), and second,
putting on the side the cosmic coincidence (see section III for a discussion). The remaining question is then why
the huge vacuum energy densities (see section II B) from quantum field theory estimations do not contribute to the
energy budget in the universe. This question can be legitimately replaced by, how does vacuum energy contribute to
the EFE? For example, is it correct to try to add contributions from vacuum energy density to the EFE using some
ultra-violet cutoff energy? We discuss this point more in section V D, see also [93, 94, 95, 96].
Further, did we really exhaust all possibility of exact cancelation mechanisms for vacuum energy? [8, 9, 10, 11, 12,
13, 14].
It is perhaps worth mentioning that a rather negative vacuum energy/cosmological constant was expected within
some candidates for a unified theory such as String Theory [98], as there is no attractive way to derive a stable vacuum
with a positive cosmological constant [98]. So, in addition to the magnitude problems, could this sign problem be a
further indication to revise the full identification above?
In this simple solution, what is measured currently is simply an intrinsic curvature of the spacetime, and the value
of Λ is just a constant measured from experiments, as is Newton’s gravitational constant G.
At this point, we would like to discuss the following subtle point. The usual aesthetic interpretation of one of General
Relativity’s principles is that the mass-energy content of the universe creates curvature of the spacetime (assuming a
zero Weyl tensor as in standard FLRW cosmology). One could then ask the following question: if spacetime is to have
a curvature in absence of mass-energy sources and this curvature is not due to vacuum energy then what is generating
this curvature? There are two possible answers to the question: If one wants to preserve the interpretation above,
then one needs to explain the source of this curvature. However, it is also correct to take the other standpoint and
consider that the Einstein field equations are a set of differential equations containing a cosmological constant and
21
governing the laws of General Relativity, and that in absence of sources the trivial spacetime is simply de Sitter with
an intrinsic curvature.
The simple solution we discussed in this section is of the latter type and is a consequence of questioning some of the
assumptions usually made. To our best knowledge, this particular simple but subtle point about loosing the aesthetic
interpretation above has not been discussed in literature about dark energy.
With these reconsiderations in mind, it is important to think about identifying new astrophysical tests or experi-
ments at the interface of particle and gravitational physics, as we discuss in the next sub-section.
D. Vacuum energy and Casimir effect in gravitational and cosmological contexts
Another possible successful approach to the cosmological constant problems is to think of a situation or an ex-
periment where the validity of the cosmological constant-vacuum energy identification can be put to the test. A
geometrical cosmological constant has no quantum properties while vacuum energy has both gravitational and quan-
tum properties. Also, is it possible to learn more on how vacuum energy contributes to the cosmological constant?
Some of these questions started to be addressed in the literature as we cite further.
It is perhaps relevant at this point to recall the Casimir effect [99] which is a purely quantum field theory phenomenon
(see [100] for a recent comprehensive review, and references therein.) The Casimir effect results from a change in the
zero-point oscillations spectrum of a quantized field when the quantization domain is restricted or when the topology
of the space is non-trivial. For example, a Casimir force appears as the result of the alteration of the vacuum energy
by some boundaries. In its simplest form, predicted by Casimir [99], two neutral plane parallel conducting plates
placed in a vacuum at a distance a from one another will experience an attractive force FC = − π2 ¯hc
240a4 S where S ≫ a
is the plate area. The Casimir effect has been now extensively measured with a few percent precision [100].
Cosmologically, the Casimir effect is significant when the topology of the model of the universe is non-trivial
(different from an infinite Euclidean topology), see e.g. [101]. The effect has been discussed in models with non-trivial
topology, notably the simple case of a closed FRW universe with a 3-torus topology, see e.g. [102]. Also, the Casimir
energy has been used from compact extra dimensions [94, 95, 96] to discuss the cosmological constant problem, and
with models with supersymmetric large extra dimensions to propose cancelation mechanisms for the cosmological
constant problem, see e.g. [97].
In a more relevant context for our discussion, one would like to study, via the Casimir effect, the gravitational
properties of the vacuum energy. For example, Refs. [103, 104] calculated correction terms to the Casimir force due
22
to the weak gravitational field. Such corrections represent the effect of gravitational curvature on quantum vacuum
fluctuations. The authors of Ref. [104] evaluated the order of the force acting on a Casimir apparatus redshifting in
a weak gravitational field and concluded that, although some issues with signal modulation need to be solved, testing
such force should be feasible and within reach of present technological resources.
Now, related to our question on how the vacuum energy may fit within the EFE, it has been argued in some papers,
see for example [94, 95, 96], that as the measured Casimir effect is related to vacuum energy differences, the vacuum
energy may not contribute to the cosmological dynamics via some fixed cutoff energy but rather via energy differences
as in Casimir energy. This Casimir energy can be produced from some compact extra dimensions [94, 95, 96] or
non-trivial topology of the spacetime [101].
We could state that if this is the case, then as we have not yet detected any non-trivial topology for a wide rang of
models [105], this could imply the vanishing of the vacuum energy contribution at cosmological scales. On the other
hand, not all non-trivial topologies have been ruled out and one could push the idea further.
Therefore, questioning and testing how vacuum energy contributes to the cosmological constant using, for example,
the Casimir effect in a cosmological context may prove helpful to the dark energy questions.
VI. CONCLUDING REMARKS
We discussed different formulations of the cosmological constant/dark energy problems and some of the assumptions
underlying them. We argued for the usefulness of clarifying and questioning some of these assumptions and identifying
new strategies in order to put them to the test.
We used some fraction calculations in order to evaluate how much of a fine-tuning is involved in the cosmic
coincidence. We found that these fractions are of the order of a few percent even in the worst cases. This is significant
but not totally unusual in physics and astrophysics. Therefore, on one hand, it remains perhaps a motivated and
interesting problem to seek models that could naturally explain these numbers. On the other hand, it was important to
clarify that this fine tuning should not constitute a barrier that rejects a successful solution for the other cosmological
constant problems.
Current and future plans are focused on constraining the equation of state of dark energy using cosmological
probes. This is certainly an important approach and some progress has been made, however as we showed in section
IV, constraining a variable equation of state will require very sophisticated and challanging future experiments.
Furtheremore, this approach is limited in the kind of decisive answers it could provide on the nature of dark energy.
23
Indeed, unless we are lucky enough to find a dark energy that has an equation of state significantly different from that
of a cosmological constant, new kinds of tests or experiments will be necessary in order to provide conclusive answers
to the dark energy problem. For example, finding that dark energy parameters are those of a cosmological constant
to a few percent precision will be very suggestive but will require tests different from the equation of state in order to
rule out decisively dynamical dark energy models. Now, even if we are ready to accept some high level of precision to
be satisfactory (or if we reach fundamental limitations of our experiments), finding dark energy parameters that are
characteristic of a cosmological constant will only confirm the cosmological constant problems with no further clues.
Further, once an equation of state is determined from cosmological observations, one is always left with the following
question: Is this an effective equation of state of some dark energy component in the energy momentum tensor or is
this a forced equation of state obtained by fitting dark energy models on the top of some modified model of gravity?
Therefore it is important to encourage other directions and strategies for approaching the dark energy problems.
In particular, we discussed the relevance of questioning and challenging some of the assumptions underlying the
formulation of the cosmological constant problems in order to look for new types of tests.
Next, we showed that comparing cosmological observations of the expansion history and cosmological observations
of the growth rate of large-scale structure can distinguish between cosmic acceleration due to some dark energy
models and cosmic acceleration due to some modification to gravity physics at cosmological scales. The basic idea is
that the effect of cosmic acceleration on the expansion function and its effect on the growth rate function must be
mathematically consistent one with another because of the underlying gravity theory (General Relativity). As shown
in section V-A, the failure in the consistency relation can be used as a test to distinguish between cosmic acceleration
due to dark energy models and acceleration due to modified gravity at cosmological scales. This consistency test shows
the potential of some cosmological probes such as, supernova searches, gravitational lensing and clusters of galaxies
to go beyond the equation of state approach in order to address the cosmological constant/dark energy questions.
Next, motivated by some theorems on the most general curvature tensor in the Einstein field equations, we argued
that the identification of the cosmological constant with the vacuum energy is not unquestionable and might bring
some limitations of its own because it changes the formulation of the old cosmological constant problem. Recall that
as a result of this questioning, dark energy can be identified as a simple geometrical cosmological constant. In the
absence of sources, the trivial spacetime is then de Sitter with an intrinsic constant curvature. However, then two
questions arise.
i) An important interpretation of one of principles of General Relativity is that the mass-energy
content of the universe creates curvature of the spacetime. One could then ask the following question: if spacetime is
24
to have a curvature in the absence of mass-energy sources, and this curvature is not due to vacuum energy, then what
is generating this curvature? There are two possible answers to this question: If we want to preserve the interpretation
above then we do need to explain this curvature. However, it is fully correct to take the other standpoint and consider
that the Einstein field equations are a set of differential equations containing a cosmological constant and governing the
laws of gravity (General Relativity), and that in the absence of sources, spacetime has an intrinsic constant curvature.
This last possibility requires one to sacrifice the important interpretation mentioned above. ii) The second question
is why would the huge vacuum energy densities evaluated from quantum field theory calculations not contribute to
the measured effective cosmological constant? As we discussed, this question could be re-addressed in the context of
how the vacuum energy may contribute to the Einstein field equations. In particular, is the usual method of using a
given ultraviolet cutoff energy as a source of gravity questionable? For example, other propositions have been made
in literature [94, 95, 96] where vacuum energy will contribute via energy differences as experienced with the Casimir
effect.
Finally, we pointed out the possible role of the Casimir effect used in gravitational and cosmological contexts for
testing some of the assumptions and questions discussed. This is a purely quantum field theory phenomenon and
could be used to look for clues on how the vacuum energy may fit within the Einstein field equations. Other kinds of
experiments at the interface between quantum field theory and general relativistic principles have been also discussed
in [106, 107, 108, 109] and might be of similar interest.
We conclude that challenging some of the assumptions underlying the formulation of the cosmological constant/dark
energy problems and putting them to the test may prove useful and necessary to make progress on these questions.
Acknowledgments
The author thanks Latham Boyle, Simon DeDeo, G.F.R. Ellis, Chris Hirata, Pat McDonald, David Spergel, Paul
Steinhardt, and Amol Upadhye for useful comments. The author thanks James Richardson for reading the manuscript.
This is not a review paper and we acknowledge that the list of references cited here is incomplete. We tried to
provide only some examples from the literature when necessary.
Partial support from the Natural Sciences and Engineering Research Council of Canada (NSERC) and NASA Theory
Award NNG04GK55G at Princeton University is acknowledged. The author acknowledges the partial support from
the Hoblitzelle Foundation and a Clark award at the University of Texas at Dallas.
FRW dust+Lambda
FRW dust
2e-171
FRW dust+Lambda
FRW dust
25
b
a
d
c
R
d
c
b
a
R
=
K
e
;
b
a
d
c
R
e
;
d
c
b
a
R
=
K
0
0
5e+27
1e+28
t
1.5e+28
2e+28
1.6e+28 1.8e+28 2e+28 2.2e+28 2.4e+28 2.6e+28 2.8e+28 3e+28
t
FIG. 3: a) Plot of the curvature invariant K = Rαβ
γδ Rγδ
αβ . b) Plot of the differential invariant DiRiem = Rαβ
γδ
;η Rγδ
αβ ;η .
The curvature decreases during a matter dominated universe to reach a constant curvature Lambda-dominated universe.
Length units are used with Λ = 10−56 cm−2 . We also display on the left vertical lines for cttransition = 0.67 × 1028 cm and
ctρΛ=ρm = 0.90 × 1028 cm.
APPENDIX A: EXAMPLES OF SPACETIME CURVATURE INCLUDING A COSMOLOGICAL
CONSTANT
In order to plot the evolution of spacetime curvature with a cosmological constant, we consider curvature invariants
constructed from the Riemann tensor, Rαβγ δ . These scalars allow a coordinate independent study of some geometrical
features of a spacetime. They can also be linked to physical quantities via the EFE. For the special spacetimes we
consider here the invariants are all related via algebraic relations [110, 111], and for the sake of simplicity we just
choose here the Kretchman scalar,
and the differential invariant
K = Rαβ
γ δ Rγ δ
αβ
DiRiem = Rαβ
γ δ
;η Rγ δ
αβ ;η
and
to trace the evolution of the curvature of spacetime. For the metric (5), the invariants read
√3Λ
√3Λ
t(cid:1)(cid:17)2
t(cid:1)(cid:17)4
K = Λ2 h 5
3 (cid:16) coth (cid:0)
− 2(cid:16) coth (cid:0)
2
2
√3Λ
2 t(cid:1)(cid:17)2
DiRiem = −9Λ3 (cid:16) coth (cid:0)
2 t(cid:1)(cid:17)4 .
√3Λ
(cid:16) sinh (cid:0)
+ 3i
(A1)
(A2)
(A3)
(A4)
1.005e-49
FRW dust+Lambda
FRW dust
FRW dust+Lambda
FRW dust
26
x
t
x
t
R
y
x
y
x
R
-1e-49
0
5e+27
1e+28
t
1.5e+28
2e+28
0
0
2e+27
4e+27
6e+27
8e+27
1e+28
t
FIG. 4: a) LEFT: Plot of the Riemnann components Rtxtx = Rty ty = Rtztz . It starts with a power law decrease to reach a
negative range exponential decrease during a de Sitter phase. For comparison, the no-Lambda curve shows how this component
continues with a power law decrease within a positive range. Length units are used with Λ = 10−56 cm−2 . We display vertical
for cttransition = 0.67 × 1028 cm and ctρΛ=ρm = 0.90 × 1028 cm.
b) RIGHT: Plot of the Riemann components Rxyxy = Rxzxz = Ryzyz . The profile is similar to that of the scale factor.
These components transit to an exponential increase at very large t. We also display cttransition = 0.67 × 1028 cm and
ctρΛ=ρm = 0.90 × 1028 cm.
For the matter dominated universe, these are simply given by, K = 80
37
has the usual line element
t4 and DiRiem = − 60
1
3
1
t6 . The de Sitter space
ds2 = −dt2 + exp (cid:0)2r Λ
3
and constant curvature with K = 3
8 Λ2 .
Figure 3a and Figure 3b show the profile of K and DiRiem and how the spacetime curvature decreases during the
expanding matter dominated universe to reach a constant curvature Λ-dominated universe at late times. This can
t(cid:1)(dr2 + r2 dΩ2 )
(A5)
also be seen from taking the limits of (A3) and (A4) at very large t.
The vertical lines in Figure 3a are the time at equality of dark energy density with matter energy density and the
time of transition from deceleration to acceleration. The no-Lambda curves are shown for comparison.
Furthermore, in order to trace some features of the curvature lost in the squared quantities, we recourse to plotting
directly the non-vanishing components of the Riemann tensor. Though coordinate dependent, these can be informative
[17].
Schwartzschild-de Sitter
Schwartzschild
27
1
100000
1e+10
1e+15
1e+20
1e+25
1e+30
1e+35
t
b
a
d
c
R
d
c
b
a
R
=
K
1e+50
1
1e-50
1e-100
1e-150
1e-200
FIG. 5: Plot of K = Rαβ
γδ Rγδ
αβ as a function of r . The curvature decreases as 1/r2 from the central mass m to become
dominated by the Λ term after r
3 Λ2 = 48m2
8
r2
m = 0.74 × 1017 cm.
In cartesian coordinates, these are:
= .36 × 1025 cm. Length units are used with Λ = 10−56 cm−2 and a mass
Λ (cid:17)2/3
6 (cid:16) 3C
Λ
×
Rtxtx = Rtyty = Rtztz =
√3Λ
√3Λ
t(cid:1)(cid:17)−2/3
t(cid:1)(cid:17)4/3 i
h(cid:16) sinh (cid:0)
− 2(cid:16) sinh (cid:0)
2
2
(A6)
and
Λ (cid:17)4/3
Rxyxy = Rxzxz = Ryzyz = 31/3Λ(cid:16) C
×
√3Λ
√3Λ
t(cid:1)(cid:17)2/3 i
t(cid:1)(cid:17)8/3
+ (cid:16) sinh (cid:0)
h(cid:16) sinh (cid:0)
2
2
The time evolution of the Rtxtx components is shown in Figure 4a where it starts with a power law decrease to
(A7)
reach a negative range exponential decrease during a de Sitter phase. For comparison, the no-λ curve shows how this
component continues with power law decrease within a positive range.
Figure 4b shows that the profile of the Rxyxy component is similar to that of the scale factor. This component
transits to an exponential increase at very large t.
Also, we consider another informative example, the Schwarzschild-de Sitter spacetime with
Λr2
2m
ds2 = −(cid:16)1 −
3 (cid:1)dt2 +
r −
Λr2
2m
(cid:16)1 −
3 (cid:1)−1
dr2 + r2 dΩ2 .
r −
At large r, it tends to the de Sitter space limit. The explicit de Sitter case is obtained by setting m = 0 while the
(A8)
explicit Schwarzschild case is obtained by setting Λ = 0. Here, for (A8)
K =
8
3
Λ2 +
48m2
r2
28
(A9)
and is plotted in Figure (5) where the curvature due to the central mass m decreases as a function of r and is overtaken
by the Λ term at very large r. The Schwarzschild curve is plotted for comparison.
[*] Electronic Address: [email protected]; Part of this work was written when the author was a research associate at
Princeton University.
[1] A. G. Riess, et. al., Astron.J. 116 1009 (1998); P. M. Garnavich et. al., Astrophys.J. 509 74 (1998); A. V. Filippenko
and A. G. Riess, Phys.Rept. 307 31 (1998); S. Perlmutter, et. al., Astrophys.J. 517 565 (1999); S. Perlmutter, et. al.,
Bull.Am.Astron.Soc. 29 1351 (1997); A. G. Riess, et. al., Astrophys.J. 536 62 (2000); A. G. Riess, et. al., Astrophys.J.
560 49 (2001); J. L. Tonry, et. al., Astrophys.J. 594 1 (2003); R. A. Knop, et. al., Astrophys. J. 598 102 (2003); B. J.
Barris, et. al., Astrophys.J. 602 571 (2004); A. G. Riess, et. al., Astrophys. J. 607 665 (2004); C. L. Bennett, et. al.,
Astrophys.J.Suppl. 148 1 (2003).
[2] D.N. Spergel et al., Astrophys. J. Supp., 148, 175 (2003); L. Page L. et al., Astrophys.J.Suppl. 148 (2003) 233.
[3] B. Netterfield et al, Ap. J., 571, 604 (2002); P. de Bernardis et al, Nature, 404, 955 (2000).
[4] M. Tegmark et al., Astrophys.J. 606 702 (2004); N. Afshordi, Y.-S. Loh, and M. A. Strauss, Phys. Rev. D 69 083524
(2004).
[5] R. Carlberg et al, Astrophys. J. 478, 462 (1997); N. Bahcall et al, Astrophys. J. 541, 1 (2000).
[6] W.L. Freedman et al., Astrophys. J. , 553, 47 (2001).
[7] P. Fosalba, E. Gaztanaga, and F. Castander, Astrophys.J. 597 L89 (2003); P. Fosalba and E Gaztanaga,
Mon.Not.Roy.Astron.Soc. 350 L37 (2004); R. Scranton, et. al., preprint astro-ph/0307249 (2003); N. Afshordi, Y.-S.
Loh, and M. A. Strauss, Phys. Rev. D 69 083524 (2004); S. Boughn and R. Crittenden, Nature, 427 6969 (2004); N.
Padmanabhan et al., Phys. Rev. D. 70, 103501 (2004).
[8] S. Weinberg Rev. Mod. Phys., 61, 1 (1989).
[9] S.M.Carroll, W.H. Press and E.L. Turner, Ann. Rev. Astron. Astrophys., 30, 499 (1992).
[10] M.S. Turner, Phys. Rep., 333, 619 (2000).
[11] Varun Sahni, Alexei Starobinsky Int.J.Mod.Phys. D9, 373 (2000).
[12] S.M. Carroll, Living Reviews in Relativity, 4, 1 (2001).
[13] T. Padmanabhan, Phys. Rep., 380, 335 (2003).
[14] P.J.E. Peebles and B. Ratra, Rev. Mod. Phys. 75, 559 (2003).
29
[15] E. J. Copeland, M. Sami, S. Tsujikawa Int.J.Mod.Phys. D15 (2006) 1753-1936
[16] P.J.E Peebles and B.Ratra, Astrophys.J.Lett. 325 L17 (1988); B.Ratra and P.J.E. Peebles, Phys.Rev.D 37 3406 (1988);
C. Wetterich, Ncl. Phys. B 302, 668 (1988); I. Zlatev, L. Wang, and P.J. Steinhardt, Phys. Rev. Lett. 82 896 (1999);
P.J.Steinhardt, L. Wang, and I.Zlatev, Phys. Rev. D 59 123504 (1999).
[17] C.W. Misner, K.S. Thorne and J.A. Wheeler, Gravitation (Freeman, San Francisco 1973).
[18] S. Weinberg, Gravitation and cosmology: principles and applications of the general theory of relativity, (Wiley, New York,
1972).
[19] R.A. D’Inverno, Introducing Einstein’s Relativity (Oxford University Press, New York, 1996).
[20] W. Rindler, Relativity: Special, General, and Cosmological Second Edition (Oxford University Press, 2006).
[21] J.J. Mohr, astro-ph/0408484 . Observing Dark Energy, ASP Conference Series, Vol. 339, Proceedings of a meeting held
18-20 March 2004 in Tucson, Arizona. Edited by Sidney C. Wolff and Tod R. Lauer. San Francisco: Astronomical Society
of the Pacific, 2005., p.140
[22] S. Wang et al., Phys. Rev. D, 70, 123008 (2004).
[23] Hu W., Phys. Rev. D 65, 023003 (2001)
[24] Takada M., Jain B., 2004, MNRAS, 348, 897
[25] M. Ishak, MNRAS, 363, 469-478 (2005).
[26] Cole, et.al., MNRAS, 362, 505 (2005).
[27] D. Eisenstein, et.al., ApJ, 633, 560 (2005).
[28] Montroy, et.al., ApJ, 647, 799 (2006).
[29] Kuo, et.al., 2004, ApJ, 600, 32 (2004).
[30] Readhead, et.al., ApJ, 609, 498 (2004).
[31] Dickinson, et.al., MNRAS, 353, 732 (2004).
[32] M. Tegmark, et.al., ApJ, 606, 702 (2004).
[33] Astier, et.al., AA, 447, 31 (2006).
[34] A. Riess, et.al., ApJ, 607, 665 (2004).
[35] A. Upadhye, M. Ishak, and P.J. Steinhardt, Phys.Rev. D72, 063501(2005).
[36] M. Chevallier, D. Polarski, and A. Starobinsky, Int. J. Mod. Phys. D 10, 213 (2001).
[37] E. Linder, Phys. Rev. Lett., 90, 091301 (2003).
[38] Y. Wang, D. N. Spergel, M. A. Strauss, Astrophys.J. 510, 20 (1999)
[45] Press et al., Numerical Recipes in C, Cambridge University Press (1992); S. Dodelson, Modern Cosmology, Academic
Press (2003); Tegmark, Taylor, and Heavens, Astrophys.J. 480, 22 (1997).
[40] C.L. Bennett, et al., Astrophys. J. Supp., 148, 1 (2003)
[41] D.N. Spergel , et al., Astrophys.J.Suppl. 170, 377 (2007).
[42] M. Ishak, C. M. Hirata, P. McDonald, U. Seljak, Phys. Rev. D 69, 083514 (2004).
[43] C. Hirata, U. Seljak, Mon. Not. R. Astron. Soc., 343, 459 (2003).
[44] Ishak M., Hirata C., Phys. Rev. D, 71, 023002 (2005)
[45] Press et al., Numerical Recipes in C, Cambridge University Press (1992); S. Dodelson, Modern Cosmology, Academic
Press (2003); Tegmark, Taylor, and Heavens, Astrophys.J. 480, 22 (1997).
30
[46] B. Jain, U. Seljak, Astrophys. J. , 484, 560 (1997).
[47] N. Kaiser, Astrophys. J. , 498, 26 (1998).
[48] N. Kaiser, Astrophys. J. , 388, 272 (1992).
[49] N. Kaiser, Astrophys. J. , 537, 555 (2000).
[50] A. Refregier, Ann. Rev. Astron. Astrophys.41, 645 (2003)
[51] M. Bartelmann and P. Schneider, Phys. Rep.340, 291 (2001)
[52] Y. Mellier, Ann. Rev. Astron. Astrophys.37, 127 (1999)
[53] T. Erben et al., a, 366, 717 (2001).
[54] D.J. Bacon, et al., Mon. Not. R. Astron. Soc., 325, 1065 (2001).
[55] G. Bernstein, M. Jarvis, Astron. J.123, 583 (2002).
[56] L. Van Waerbeke and Y. Mellier, astro-ph/0305089 . Lecture given at the Aussois winter school, (2003).
[57] Wittman et al., Nature, 405, 143 (2000).
[58] M. Tegmark et al., astro-ph/9805117 (1998).
[59] D. Huterer, M. S. Turner, Phys. Rev. D, 64, 123527 (2001).
[60] G. Aldering G., et al., for the SNAP pro ject collaboration, astro-ph/0405232 (2003).
[61] Kim et al., 2003, Mon. Not. R. Astron. Soc., 347, 909 (2004).
[62] J.L. Tonry, et al, Astrophys. J. , 594, 1 (2003).
[63] I. Zlatev, L. Wang, P.J. Steinhardt, Phys. Rev. Lett. 82, 896 (1999).
[64] P. Brax, J. Martin, Phys. Lett. B, 468, 40 (1999).
[65] G.R. Dvali, G. Gabadadzi, and M. Porrati, Phys. Lett. B 485, 208 (2000).
[66] C. Deffayet, et. al., Phys. Rev. D 66, 024019 (2002).
[67] C. Deffayet, G. Dvali, G. Gabadadze, Phys.Rev. D 65 044023 (2002).
[79] A. Lue, R. Scoccimarro, G. Starkman,. Phys. Rev. D 69, 124015 (2004).
[69] A. Lue, Phys. Rept. 423,1 (2006).
[70] I. Sawicki, S. M. Carroll, preprint astro-ph/0510364 .
[71] M. Fairbairn, A. Goobar, preprint astro-ph/0511029 .
31
[72] C.P. Ma, R. R. Caldwell, P. Bode, L. Wang, Astrophys. J. , 521, L1 (1999)
[73] E. Linder, A. Jenkins, MNRAS, 346, 573 (2003).
[74] S. Carroll et al., Phys.Rev. D71 063513 (2005).
[75] S. Capozziello, V.F. Cardone, A. Troisi, Phys.Rev. D 71 043503 (2005).
[76] J. Simon, L. Verde, R. Jimenez, Phys.Rev. D71 123001(2005).
[77] C. Deffayet, Phys. Lett. B 502, 199-208 (2001).
[78] C. Deffayet, Phys. Rev. D 66, 103504 (2002).
[79] A. Lue, R. Scoccimarro, G. Starkman,. Phys. Rev. D 69, 124015 (2004)
[80] K. Koyama, R. Maartens JCAP 601, 16 (2006)
[81] C. Deffayet, S. J. Landau, J. Raux, M. Zaldarriaga, P. Astier, Phys. Rev. D 66, 024019 (2002).
[82] Y. Wang and M. Tegmark, Phys. Rev. Lett. 92 241302 (2004).
[83] M. Ishak, A. Upadhye, D.N. Spergel, Phys. Rev. D 74, 043513 (2006) (astro-ph/0507184 (2005)).
[84] L. Knox, Y. Song, and J. A. Tyson Phys. Rev. D 74, 023512 (2006) (astro-ph/0503644 (2005)).
[85] E. V. Linder, Phys.Rev. D72 043529 (2005). (astro-ph/0507263 (2005)).
[86] C. Schimd, J.P. Uzan, and A. Riazuelo, Phys.Rev. D 71 083512 (2005).
[87] E. Cartan, J. Math. Pure Appl., 1, 141 (1922).
[88] H. Weyl, Space, time, matter (Dover, New York, 1922).
[89] H. Vermeil, Nachr. Ges. Wiss. Gottingen, 334 (1917).
[90] D. Lovelock, J. Math. Phys., 12, 498 (1971).
[91] D. Lovelock, J. Math. Phys., 13, 874 (1972).
[92] S.M. Carroll astro-ph/0107571 .
[93] T. Padmanadhan, Class.Quant.Grav. 22, L107-L110 (2005)
[94] K. A. Milton, Grav.Cosmol. 8, 65 (2002). ibid 9, 66 (2003).
[95] K. A. Milton, R. Kantowski, C. Kao, Y. Wang Mod. Phys. Lett. A 16, 2281 (2001).
[96] P. O. Mazur, E. Mottola, gr-qc/0405111 . Contributed talk at the Sixth Workshop on ’Quantum Field Theory Under
The Influence Of External Conditions’ held at the Univ. of Oklahoma, September 15-19, 2003, to be published in the
Proceedings by Rinton Press (2004)
[97] C. P. Burgess, hep-th/0411140, THE NEW COSMOLOGY: Conference on Strings and Cosmology; The Mitchell Sympo-
sium on Observational Cosmology. AIP Conference Proceedings, Volume 743, pp. 417-449 (2004).
[98] E. Witten, hep-ph/0002297. Sources and Detection of Dark Matter and Dark Energy in the Universe. Fourth International
Symposium, held February 23-25, 2000, at Marina del Rey, California, USA. Edited by David B. Cline. Published by
Springer-Verlag, Berlin, New York, ISBN 3-540-41216-6, p.27
32
[99] H. Casimir Proc. K. Ned. Akad. Wet. 51, 793 (1948).
[100] M. Bordag, U. Mohideen, V.M. Mostepanenko Phys. Rep.353, 1 (2001).
[101] M. Lachieze-Rey, J.P. Luminet, Phys. Rep.C254, 135 (1995).
[102] Y.B. Zeldovich, A.A Starobinsky Pisma v Astronomicheskii Zhurnal, 10, 323 (1984).
[103] R. R. Caldwell astro-ph/0209312
[104] E. Calloni, L. Di Fiore, G. Esposito, L. Milano, L. Rosa, Phys. Lett. A 297, 328 (2002).
[105] N. J. Cornish, D. N. Spergel, G. D. Starkman, E. Komatsu Phys. Rev. Lett. 92, 201302 (2004).
[106] S. Reynaud, A. Lambrecht, C. Genet, M.T Jaekel C. R. Acad. Sci. Paris, 2-IV (2001) p.1287-1298. (quant-ph/0105053)
(2001).
[107] M.T. Jaekel, S, Reynaud Rept. Prog. Phys. 60, 863 (1997).
[108] L. Viola, R. Onofrio Phys. Rev. D 55, 455 (1997).
[109] G. Papini, gr-qc/0110056 . Proceedings of the 17th Course of the International School of Cosmology and Gravitation
”Advances in the interplay between quantum and gravity physics” edited by V. De Sabbata and A. Zheltukhin, Kluwer
Academic Publishers, Dordrecht
[110] J. Carot and da Costa, Class. Quantum Grav. 10, 461 (1993).
[111] K. Santosuosso, D. Pollney, N. Pelavas, P. Musgrave and K. Lake, Computer Physics Communications 115, 381 (1998).
|
astro-ph/0208479 | 1 | 0208 | 2002-08-27T17:30:14 | Discovery of a peculiar DQ white dwarf | [
"astro-ph"
] | We report the discovery of a new carbon rich white dwarf that was identified during a proper motion survey for cool white dwarfs based on photographic material used for the construction of the Guide Star Catalog II. Its large proper motion (0.48 arcsec/yr) and faint apparent magnitude (V = 18.7) suggest a nearby object of low luminosity. A low-resolution spectrum taken with the William Herschel Telescope clearly shows strong C2 Deslandres-d'Azambuja and Swan bands, which identify the star as a DQ white dwarf. The strength of the Deslandres-d'Azambuja bands and the depression of the continuum in the Swan-band region are signs of enhanced carbon abundance for the given Teff. Comparison of our spectrophotometric data to published synthetic spectra suggests 6000 K < Teff < 8000 K although further analysis with specialized synthetic models appear necessary to derive both Teff and chemical composition. Finally, the range of spatial velocity estimated for this object makes it a likely member of the halo or thick disk population. | astro-ph | astro-ph | Astronomy & Astrophysics manuscript no.
(will be inserted by hand later)
2
0
0
2
g
u
A
7
2
1
v
9
7
4
8
0
2
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
Discovery of a peculiar DQ white dwarf ⋆,⋆⋆
D. Carollo,1 S.T. Hodgkin,2 A. Spagna,1 R.L. Smart,1 M.G. Lattanzi,1 B.J. McLean,3 D.J. Pinfield4
1 INAF, Osservatorio Astronomico di Torino, I-10025 Pino Torinese, Italy
2 Cambridge Astronomical Survey Unit, Institute of Astronomy, Madingley Road, Cambridge, CB3 0HA, UK
3 Space Telescope Science Institute (STScI), Baltimore, MD 21218, USA
4 Astrophysics Research Institute, Liverpool John Moores University, Birkenhead, CH41 1LD, UK
Received 23 January 2002/Accepted 2 August 2002
Abstract. We report the discovery of a new carbon rich white dwarf that was identified during a proper motion
survey for cool white dwarfs based on photographic material used for the construction of the Guide Star Catalog
II. Its large proper motion (µ ≃ 0.48 arcsec/yr) and faint apparent magnitude (V ≃ 18.7) suggest a nearby object
of low luminosity. A low-resolution spectrum taken with the William Herschel Telescope clearly shows strong
C2 Deslandres-d'Azambuja and Swan bands, which identify the star as a DQ white dwarf. The strength of the
Deslandres-d'Azambuja bands and the depression of the continuum in the Swan-band region are signs of enhanced
carbon abundance for the given Teff . Comparison of our spectrophotometric data to published synthetic spectra
suggests 6000 K < Teff < 8000 K, although further analysis with specialized synthetic models appear necessary to
derive both Teff and chemical composition. Finally, the range of spatial velocity estimated for this object makes
it a likely member of the halo or thick disk population.
Key words. White dwarfs -- Stars: carbon -- Stars: kinematics -- Stars: individual(GSC2U J131147.2+292348) --
Astrometry -- Techniques: spectroscopic
1. Introduction
Star GSC2U J131147.2+292348 was identified during a
proper motion survey for cool halo white dwarfs (WDs)
based on photographic material used for the construction
of the Second Guide Star Catalogue (GSC-II) (see, e.g.,
Lasker et al. 1995, McLean et al. 2000). The object is lo-
cated near the North Galactic Pole (NGP) at l ≃ 61◦,
b ≃ 85◦, is fast moving (µ ≃ 0.48 arcsec yr−1), and faint
(V ≃ 18.7), as expected for a low luminosity object in the
solar neighborhood. An accurate check on the SIMBAD
database revealed that the star is not in the NLTT cat-
Send offprint requests to: Daniela Carollo
e-mail: [email protected]
⋆ Based on observations made with the William Herschel
Telescope operated on the island of La Palma by the Isaac
Newton Group in the Spanish Observatorio del Roque de los
Muchachos of the Instituto de Astrofisica de Canarias.
⋆⋆ Based on observations made with the Italian Telescopio
Nazionale Galileo (TNG) operated on the island of La Palma
by the Centro Galileo Galilei of the INAF (Istituto Nazionale
di Astrofisica) at the Spanish Observatorio del Roque de los
Muchachos of the Instituto de Astrofisica de Canarias.
alogue (Luyten 1979) but, quite surprisingly, is listed as
a quasar candidate (object OMHR 58793) by Moreau &
Reboul (1995), who measured an UV excess but did not
detect any proper motion.
2. Observations and Data Analysis
2.1. Astrometry and photometry
Our material consists of Schmidt plates from the Northern
photographic surveys (POSS-I, Quick V and POSS-II)
carried out at the Palomar Observatory (see Table 1). All
plates were digitized at STScI utilizing modified PDS-type
scanning machines with 25 µm square pixels (1.7 ′′/pixel)
for the first epoch plates, and 15 µm pixels (1 ′′/pixel) for
the second epoch plates (Laidler et al. 1996). These digital
copies of the plates were initially analyzed by means of the
standard software pipeline used for the construction of the
GSC-II. The pipeline performs object detection and com-
putes parameters and features for each identified object.
Further, the software provides classification, position, and
magnitude for each object by means of astrometric and
photometric calibrations which utilized the Tycho2 (Høg
2
Carollo et al.: Discovery of a peculiar DQ star
Table 1. GSC2 plate material used for the astrometry and the photographic photometry of the new DQ white dwarf.
Field
Center (J2000)
Survey
13:04:14.7 +29:48:37
XJ443 POSS-II
13:04:15.2 +29:48:42
XP443 POSS-II
XI443 POSS-II
13:04:20.7 +29:44:17
N322 Quick V 13:06:56.6 +29:13:25
13:06:55.5 +29:13:25
13:06:56.1 +29:13:24
POSS-I
POSS-I
XE322
XO322
Epoch
1995.234
1993.288
1991.299
1983.294
1955.288
1955.288
Pixel
15 µm
15 µm
15 µm
25 µm
25 µm
25 µm
Color Emulsion + Filter
BJ
RF
IN
V12
E
O
IIIaJ + GG385
IIIaF + RG610
IV-N + RG9
IIaD+Wratten 12
103a-E + red plexiglass
103a-O unfiltered
tion refers to the epoch of the most recent plate (XJ443),
while the accurate proper motion was computed by
combining the image locations of the star as measured on
the 6 different plates of Table 1, which span ∼ 40 years.
The photographic magnitudes are given in the natural
photometric system of the POSS-II and Quick-V plates
as defined by the emulsion-filter combinations in Table 1.
In particular, the transformation between the photo-
graphic and Johnson V is V12 = V − 0.15(B − V ) ac-
cording to Russell et al. (1990). Also, recently acquired
NIR images 1 provided the J, H, Ks magnitudes in Table
2. Finally, Moreau & Reboul (1995) published the values
U ≃ 19.15 and V ≃ 19.10. Note that their visual mag-
nitude is fairly consistent with our V12, considering the
above color transformation and the errors of the photo-
graphic photometry.
2.2. Spectroscopy
Spectroscopy of GSC2U J131147.2+292348 was obtained
on the night of 2001 January 29 using the intermediate
dispersion spectrographic and imaging system (ISIS) on
the 4.2-m William Herschel Telescope on the island of La
Palma. The 5700 A dichroic was used to split the light
and feed to the blue and red arms of the spectrograph.
We used the R158B grating on the blue arm, which
gave a nominal dispersion of 1.62 A/px and useful wave-
length coverage from 3200 to 5700 A. (The dichroic cuts
in at wavelengths > 5700 A, and at short wavelengths,
the sensitivity falls off with the quantum efficiency of the
detector). On the red arm, we used the R158R grating to
give a nominal dispersion of 2.9 A/px covering from 5500
to 8000 A. A blocking filter (GG495) was also used on the
red arm to cut out second order blue light. A 30-minute
exposure was made using a 1-arcseconds slit. Subsequent
exposures were taken of the spectrophotometric standards
Feige 67 and Feige 34 to enable flux calibration of the pri-
mary target. We took arc lamp exposures to enable wave-
length calibration and tungsten lamp exposures for the
pixel-to-pixel sensitivity variation and enable flat fielding.
The data were reduced within the IRAF environment,
following standard procedures. No attempt was made to
correct for extinction, both standards and targets were
measured with an airmass ≤ 1.1. Observations were made
with a slit width of 1.02 arcseconds, which corresponds to
1 Taken with the NICS camera on the 3.6-m TNG telescope
on La Palma
Fig. 1. First epoch (POSS-I, XE322) and second epoch
(POSS-II, XP443) plates in the direction of the newly dis-
covered WD, the encircled star near the field center. The
large relative motion of the object is evident.
et al. 2000) and the GSPC-2 (Bucciarelli et al. 2001) as
reference catalogs. Accuracies better than 0.1-0.2 arcsec
in position and 0.15-0.2 mag in magnitude are generally
attained.
Star GSC2U J131147.2+292348 was part of the sam-
ple of WD candidates discovered after screening the high
proper motion stars found in survey field 443 (Table 1).
These were selected on the basis of their relative proper
motions as derived by applying the procedure described in
Spagna et al. (1996) to just the POSS-II plates, spanning
∼ 4 years. The finding charts in Figure 1 show the high
proper motion of this object.
The
of GSC2U
J131147.2+292348 are given in Table 2. The posi-
photometry
astrometry
and
Table 2. Astrometry and photometry of GSC2U
J131147.2+292348. The position was determined from
plate XJ443 (epoch 1995.234, equinox J2000), while all
of the available plates were used for the proper motions.
The error of the photographic photometry is better than
0.2 mag (1σ).
α (h m s)
(J2000)
δ (d m s)
(J2000)
µα cos δ
(arcsec/yr)
13 11 47.21 +29 23 48.0 −0.382 ± 0.002
µδ
(arcsec/yr)
0.286 ± 0.005
BJ
19.6
V12
18.7
RF
18.1
IN
17.5
J
H
Ks
17.48 ± 0.05
17.13 ± 0.10
17.08 ± 0.12
Carollo et al.: Discovery of a peculiar DQ star
3
4 detector pixels in the blue, i.e. a dispersion of 6.5 A per
resolution element. For the red arm, the pixel scale is 0.36
arcseconds per pixel, leading to a resolution element of size
3 pixels, i.e. a resolution of 8.2 A. The blue and red arm
spectra have been gaussian smoothed at these resolutions.
Good agreement between the red and blue arm spectra
was found in the overlap region, with fluxes agreeing to
better than 10% in the range 5600-5700 A.
3. On the nature of GSC2U J131147.2+292348
of
spectrum
flux-calibrated
GSC2U
The
J131147.2+292348 is shown in Figure 2. The signal-
to-noise is around 10 for the whole spectrum, increasing
slightly to the red. This noise level is clearly visible in the
spectrum, and limits our ability to detect weak features.
The crosses in Figure 2 represent the fluxes at differ-
ent effective wavelengths as derived from the BJ , V12,
RF , and IN photographic magnitudes in Tab. 2. The ul-
traviolet flux was derived from the photographic U magni-
tude of Moreau et al. (1995). The agreement appears rea-
sonably consistent with the 10% and 20% accuracy levels
of the flux-calibrated spectroscopy and the photographic
photometry, respectively.
The spectrum appears dominated by strong absorp-
tion bands due to C2 molecules. The four Swan bands
with bandheads at λ = 4382, 4737, 5165, and 5636 A are
clearly identified, along with the less common Swan band
at 6191 A. In addition, strong Deslandres and d'Azambuja
(D-d'A) absorption bands are also present in the blue part
of the spectrum at 3600, 3852, and 4102 A. These bands
have been observed in the spectra of WDs with carbon rich
atmospheres (DQ WDs) and temperatures2 above 6500 K.
Finally, the spectrum in Figure 2 shows an evident depres-
sion of the continuum in the Swan band region between
4500 and 6200 A.
The spectral energy distribution (SED) of DQ stars
changes with Teff and carbon abundance as shown by
the model atmosphere spectra presented in Koester et
al. (1982) and Wegner & Yackovich (1984). Figure 5 of
Wegner & Yackovich gives and indication on what to ex-
pect for different combinations of Teff and C:He abun-
dance. Swan bands are generally present, while D-d'A
bands start to become visible in models with C:He>∼ 10−6
at Teff ≃ 6600 K and with C:He >∼ 10−2 at Teff ≃ 10 000
K.
A SED with C2 bands similar in strength to those ob-
served in our spectrum requires a much enhanced C:He
ratio for the given Teff. This can be seen by comparing the
models in Figure 5 of Wegner & Yackovich with those in
their Figures 2 and 3. At temperatures between 6000 K
and 7000 K, deep absorption bands are produced with
C:He ≈ 10−4. At Teff = 8000 K, carbon abundance has
to increase to a rather extreme value, C:He=0.9, for the
2 This temperature seems to be the lower limit for DQ stars,
and might be associated with the transition of C2 into C2H
molecules (Bergeron et al. 2001).
simultaneous presence of strong D-d'A and Swan bands in
the synthetic SED (bottom panel of Figure 3 of Wegner &
Yackovich). This model bears the most resemblance with
the spectrum of our WD, however, it does not show any
evidence of the continuum depression seen in the observed
spectrum. Theoretical evidence that such depression of the
continuum emission could occur is provided in Koester et
al. (1982). Their Figure 1 displays theoretical C2 spec-
tra at Teff = 8000 K and increasingly higher C:He ratios.
The effect is to boost band strengths, thus depressing the
continuum in the Swan-band region.
Although the models with Teff = 8000 K just exam-
ined seem consistent with the appearance of the C2 band
systems observed in the spectrum of our WD, the relative
flux at blue wavelengths (below ∼ 4100 A) is probably
too high compared to the observed SED in Figure 2. In
this regard, an attempt to find a black body compatible
with the observed spectrum at λ > 7000 A, the NIR fluxes
from our JHKs magnitudes, and with the blue peaks in
the D-d'A region, resulted in a black-body temperature of
∼ 6000 K. (Note that in this case the depressed continuum
occurs in the region of maximum black-body emission.)
From the discussion above, it is evident that much is
still to be learned about the properties of this new DQ
star, and the reliable determination of its temperature
and chemical composition must await more detailed at-
mosphere models. Also, improved spectral coverage in the
UV, below 3500 A, would probably be of help in better
constraining model calculations.
Finally, an approximate photometric parallax for
GSCU J131147.2+292348 was estimated from the abso-
lute magnitudes of theoretical models of non-DA stars.
From the values in Tables 2 and 4 of Bergeron et al.
(1995) for pure helium atmosphere WDs and averaging
the distance moduli computed for the IJHKs bands (which
are not affected by the strong C2 absorption bands) we
estimate the distances d ≈ 70, 80, and 90 parsecs for
Teff = 6000 K, 7000 K, and 8000 K, respectively.
This distance interval corresponds to a range of tan-
gential velocity Vtan = 4.74 · µ d ≃ 160-200 km s−1
and galactic components3 with respect to the LSR from
(U, V ) ≃ (−148.1, +9.6) to (U, V ) ≃ (−193.3, +10.8)
km s−1, for d = 70 pc and 90 pc, respectively. These rela-
tively high values are not consistent (3σ) with the velocity
distribution of the thin disk, while they are consistent with
the kinematics of the halo or thick disk stellar population4.
3 Assuming a solar motion of U⊙ = 10.00 km s−1 and V⊙ =
5.25 km s−1, as from Dehnen & Binney (1998). The (U,V) com-
ponents are computed from µα cos δ, µδ only. However, given
the high galactic latitude of this star (b ≃ 85◦), the unknown
Vr component would contribute less than 4% and 7.6% to the
U and V values, respectively.
4 Here, we
have
the
adopted
ellipsoids
(σU , σV , σW ; va) = (34, 21, 18; 6) km s−1 and (61, 58, 39; 36)
km s−1 for the thin and thick disk respectively (Table
10.4 of Binney & Merrifield 1998). The halo ellipsoid
(σU , σV , σW ; va) = (160, 89, 94; 217) km s−1
from
Casertano, Ratnatunga & Bahcall (1990). Note that these
velocity
is
4
Carollo et al.: Discovery of a peculiar DQ star
Fig. 2. The WHT optical spectrum of GSC2U J131147.2+292348. Vertical marks indicate the locations of the strong
C2 Deslandres-d'Azambuja and Swan bands, and of the telluric O2. The crosses refer to the fluxes (with ±20% error
bars) derived from the BJ , V12, RF , IN photographic photometry of Table 2 and the U mag from Moreau and Reboul
(1995).
4. Conclusions
We have discovered a new carbon rich white dwarf (DQ),
which shows very strong C2 Deslandres-d'Azambuja and
Swan bands. To the best of our knowledge, no other object
is known today which such a strong simultaneous evidence
of the two molecular band systems associated with C2.
Comparisons to published synthetic spectra suggest
6000 < Teff < 8000 K, while a black-body fit to the
observed fluxes at λ > 7000 A, and to the peaks be-
low ∼ 4100 A supports the possibility that TBB ∼
6000 K. Therefore, it is evident that the reliable determi-
nation of temperature and chemical composition of GSCU
J131147.2+292348 must await more detailed atmosphere
model calculations. Anyhow, it is likely that the carbon
abundance in the atmosphere of this WD is significantly
enhanced compared to other known DQ stars of similar
temperature.
A photometric distance of 70-90 parsecs has been esti-
mated, which implies a relatively large spatial velocity and
makes this new DQ white dwarf a likely member of the
halo or thick disk population. Of course, a direct determi-
nation of the distance will be the only way to derive model
independent absolute magnitude and kinematics for this
object.
Acknowledgements. We are indebted to the referee, U. Heber,
for his valuable comments and suggestions that were essential
for the proper interpretation of our observations. The constant
support of our GSC2 collaborators B. Bucciarelli, J. Garcia,
V. Laidler, C. Loomis, and R. Morbidelli is acknowledged.
And thanks go also to A. Boden and R. Cutri who repro-
cessed their 2MASS frames to look for this object. The GSC II
is a joint project of the Space Telescope Science Institute
and the Osservatorio Astronomico di Torino. Space Telescope
kinematics parameters are still not well established. In partic-
ular the estimated (U,V) components would result consistent
with the halo kinematics, but only marginally with the thick
disk parameters, recently derived by Chiba and Beers (2000).
Science Institute is operated by AURA for NASA under con-
tract NAS5-26555. Current participation of the Osservatorio
Astronomico di Torino is supported by the Italian National
Institute for Astrophysics (INAF). Partial financial support to
this research comes from the Italian CNAA and the Italian
Ministry of Research (MIUR) through the COFIN-2001 pro-
gram. STH and DJP acknowledge the financial support of
the Particle Physics and Astronomy Reasearch Council of the
United Kingdom. This research has made use of the SIMBAD
database, operated at CDS, Strasbourg (France).
References
Bergeron, P., Wesemael, F., Beauchamp, A. 1995, PASP, 107,
1047
Bergeron, P., Leggett, S.K., Ruiz, M.T. 2001, ApJ SS, 133, 413
Binney, J. & Merrifield, M. 1998, Galactic Astronomy,
Princeton Univ. Press
Bucciarelli, B., Garc´ıa Yus, J., Casalegno, R., Postman, M.,
Lasker, B. M., Sturch, C., Lattanzi, M.G., McLean, B.J.,
et al. 2001, A&A, 368, 335
Casertano, S., Ratnatunga, K.U., Bahcall, J.N. 1990, ApJ, 357,
435
Chiba, M., Beers, T.C. 2000, AJ, 119, 2843
Dehnen, W., Binney, J.J. 1998, MNRAS, 298, 387
Høg, E., Fabricius, C., Makarov, V.V., Urban, S., et al. 2000,
A&A 355, 27
Koester, D., Weidemann, V. & Zeidler-K.T., E.M. 1982, A&A
116, 147
Laidler, V.G., Sturch, C.R., Greene, G.R., Lasker, B.M., et al.
1996, BAAS, 188, 54.21
Lasker B.M., McLean B.J., Jenkner H., Lattanzi, M.G., Spagna
A. 1995, Future Possibilities for Astrometry in Space,
Cambridge (England, UK), Jun 19-21, 1995, Perryman
M.A.C., van Leeuwen F. & Guyenne T.-D. eds., ESA SP-
379, 137-141
Luyten, W.J. 1979, NLTT, Minneapolis, Univ. of Minnesota
McLean B.J., Greene G.R., Lattanzi M.G., Pirenne B., 2000,
ADASS IX, Kohala Coast (HI, USA), Oct 3-6, 1999,
Manset N., Veillet C. & Crabtree D. eds., ASP Conf. Ser.,
216, 145-148
Moreau, O., Reboul H. 1995, A&A SS, 111, 169
Carollo et al.: Discovery of a peculiar DQ star
5
Russell, J.L., Lasker, B.M., Sturch, C.R. & Jenkner H. 1990,
AJ, 99, 2059
Spagna, A., Lattanzi, M. G., Lasker, B. M., McLean, B. J.,
Massone, G., Lanteri, L. 1996, A&A, 311, 758
Wegner, G. & Yackovich F.H. 1984, ApJ, 284, 257
|
astro-ph/0305161 | 1 | 0305 | 2003-05-09T18:49:54 | Peculiar spectral and power spectral behaviour of the LMXB GX 13+1 | [
"astro-ph"
] | We present results of an analysis of all 480 ks of Rossi X-ray Timing Explorer Proportional Counter Array data obtained from 17 May 1998 to 11 October 1998 on the luminous low mass X-ray binary GX 13+1. We analysed the spectral properties in colour-colour diagrams (CDs) and hardness-intensity diagrams (HIDs) and fitted the power spectra with a multi-Lorentzian model. GX 13+1 traces out a curved track in the CDs on a time scale of hours, which is very reminiscent of a standard atoll track containing an island, and lower and upper banana branch. However, both count rate and power spectral properties vary along this track in a very unusual way, not seen in any other atoll or Z source. The count rate, which varied by a factor of ~1.6, along a given track first decreases and then increases, causing the motion through the HIDs to be in the opposite sense to that in the CD, contrary to all other Z and atoll sources. Along a CD track, the very low frequency noise uniquely decreases in amplitude from ~5 to ~2% (rms). The high frequency noise amplitude decreases from ~4% to less than 1% and its characteristic frequency decreases from ~10 to \~5 Hz. The 57-69 Hz quasi-periodic oscillation (QPO) found earlier is also detected, and no kHz QPOs are found. In addition the entire track shows secular motion on a time scale of about a week. The average count rate as well as the amplitude of the very low frequency noise correlate with this secular motion. We discuss a possible explanation for the peculiar properties of GX 13+1 in terms of an unusual orientation or strength of a relativistic jet. | astro-ph | astro-ph |
Astronomy&Astrophysicsmanuscript no.
(DOI: will be inserted by hand later)
November 16, 2018
Peculiar spectral and power spectral behaviour
of the LMXB GX 13+1
R. S. Schnerr1, T. Reerink1, M. van der Klis1, J. Homan1
,
2, M. M´endez1
,
3, R. P. Fender1, and E. Kuulkers4
1 Astronomical Institute "Anton Pannekoek", University of Amsterdam, and Center for High Energy Astrophysics, Kruislaan
403, 1098 SJ Amsterdam, The Netherlands
2 INAF - Osservatorio Astronomico di Brera, via Bianchi 46, 23807 Merate (LC), Italy
3 SRON, National Institute for Space Research, Sorbonnelaan 2, 3584 CA Utrecht, The Netherlands
4 ESA-ESTEC, Science Operations & Data Systems Division, SCI-SDG, Keplerlaan 1, 2201 AZ Noordwijk, The Netherlands
Received 10 Dec 2002 / Accepted 06 May 2003
Abstract. We present results of an analysis of all 480 ks of Rossi X-ray Timing Explorer Proportional Counter Array data
obtained from 17 May 1998 to 11 October 1998 on the luminous low mass X-ray binary GX 13+1. We analysed the spectral
properties in colour-colour diagrams (CDs) and hardness-intensity diagrams (HIDs) and fitted the power spectra with a multi-
Lorentzian model. GX 13+1 traces out a curved track in the CDs on a time scale of hours, which is very reminiscent of a
standard atoll track containing an island, and lower and upper banana branch. However, both count rate and power spectral
properties vary along this track in a very unusual way, not seen in any other atoll or Z source. The count rate, which varied by a
factor of ∼1.6, along a given track first decreases and then increases, causing the motion through the HIDs to be in the opposite
sense to that in the CD, contrary to all other Z and atoll sources. Along a CD track, the very low frequency noise uniquely
decreases in amplitude from ∼5 to ∼2% (rms). The high frequency noise amplitude decreases from ∼4% to less than 1% and
its characteristic frequency decreases from ∼10 to ∼5 Hz. The 57-69 Hz quasi-periodic oscillation (QPO) found earlier is also
detected, and no kHz QPOs are found. In addition the entire track shows secular motion on a time scale of about a week. The
average count rate as well as the amplitude of the very low frequency noise correlate with this secular motion. We discuss a
possible explanation for the peculiar properties of GX 13+1 in terms of an unusual orientation or strength of a relativistic jet.
Key words. accretion, accretion disks -- stars: individual: GX 13+1 -- stars: neutron -- binaries: close -- X-rays: binaries
1. Introduction
Low mass X-ray binaries (LMXBs) containing a neutron star
can be divided into two subclasses, atoll and Z sources,
based on their correlated X-ray timing and spectral properties
(Hasinger & van der Klis 1989). In the colour-colour diagram
(CD) LMXBs tend to trace out a well-defined, fully connected
1-dimensional track on a time scale of hours to days for Z
sources and days to weeks for atoll sources.
Z sources trace out a Z-shaped track in the CD, where the
three branches of the Z from top left to bottom right are called
the horizontal branch (HB), the normal branch (NB) and the
flaring branch (FB). The (sharp) turns in the track separat-
ing these branches are called vertices. The power spectral fea-
tures found in Z sources are very low frequency noise (VLFN),
which has an approximately power law shape and is observed
up to a few Hz, a band limited noise (BLN) component with
a characteristic frequency of 2 to 20 Hz and various differ-
ent quasi periodic oscillations (QPOs) with a Lorentzian shape
to:
Send
[email protected]
requests
offprint
R.
S.
Schnerr,
e-mail:
and frequencies that range from a few Hz to more than a kHz
(van der Klis 1995; van der Klis 2000).
These power spectral features occur correlated to the posi-
tion of the source in the Z track (see Fig. 1). QPOs with fre-
quencies of 13-55 Hz are found on the HB (HBOs), and with
frequencies of 4-7 Hz on the NB (NBOs); in some sources the
latter continue onto the FB where the frequency increases up
to ∼20 Hz. BLN with a fractional root-mean-square amplitude
(rms) of up to ∼8% is found along the HB and disappears as
the source moves onto the NB. There is a strong tendency of
the amplitude of the VLFN to increase from less then 1% to
∼10% along the whole track from the HB to the FB. The flux
increases from left to right along the HB, stays approximately
constant or slightly decreases along the NB, and usually in-
creases again on the FB. The mass accretion rate ( M) is thought
to increase along the HB via the NB to the FB on time scales
of hours to days (Hasinger & van der Klis 1989), although on
longer time scales the relation with M may be more compli-
cated (e.g. Kuulkers et al. 1994; Homan et al. 2002).
Atoll sources trace out a track that generally has a C or
U-shape, consisting of an island branch, which often takes the
shape of one or more isolated patches because of observational
2
R. S. Schnerr et al.: Peculiar behaviour of the LMXB GX 13+1
HB
NB
FB
IS
LB
UB
HB
NB
FB
LFN
HBO
H
F
N
NBO
IS
HFN
Q
P
O
?
LB
UB
HFN
VLFN
VLFN
Fig. 1. Typical CDs (top) of Z (left) and atoll (right) sources, and power spectra (bottom) corresponding to their different branches
in the CD. The power spectra of the HB and the island state (IS) were shifted upwards by a factor of 104 and those of the NB and
lower banana (LB) by a factor of 102. The figure was taken from Wijnands (2001).
windowing (usually in the upper left region of the CD), and a
banana branch (extending from the lower banana (LB) -- lower
left -- to the upper banana (UB) -- upper right). VLFN, BLN and
QPOs similar to those in Z sources also occur in atoll sources
(see Fig. 1; Wijnands 2001). Along an atoll-source track from
the island to the lower, middle and then to the upper banana,
the rms amplitude of the VLFN increases from ∼2 to 5%, the
characteristic frequency of the BLN, which sometimes consists
of several components (van Straaten et al. 2002), first increases
and then decreases within a range of ∼2 to ∼70 Hz, and its
rms decreases from ∼20% to less than 1%. The count rate in-
creases by a factor of ∼2-10 (Hasinger & van der Klis 1989)
going from the island state to the banana state (except for the
transient sources, see below). The frequencies of the various
QPOs increase in the same sense along the track in Z and atoll
sources.
GX 13+1 is a LMXB containing a neutron star, since
X-ray bursts have been found from this source (Fleischman
1985; Matsuba et al. 1995). Together with GX 3+1, GX 9+1
and GX 9+9, GX 13+1 forms the subclass of persistently
bright atoll-sources (the "GX atoll sources") which were ob-
served to be in the banana state by Hasinger & van der Klis
(1989). In GX 13+1 and GX 3+1, two branched structures
have been observed in the CD and hardness-intensity dia-
gram (HID) (Stella et al. 1985; Lewin et al. 1987; Schulz et al.
1989; Homan et al. 1998; Muno et al. 2002). In luminos-
ity,
these sources are intermediate between the very lu-
minous Z sources and the weaker remaining atoll sources
(Ford et al. 2000). Contrary to the other atoll sources and the
Z sources, these GX atoll sources have so far not shown
kHz QPOs. Although GX 13+1 has been classified as an
atoll source based on upper-banana branch phenomenology
(Hasinger & van der Klis 1989), among the atoll sources it is
the source that shows X-ray properties closest to those of the
Z sources (VLFN amplitude: Hasinger & van der Klis 1989,
CD track: Muno et al. 2002, ∼65 Hz QPOs: Homan et al.
1998). The radio luminosity of GX 13+1 is similar to that
of the Z sources, whereas the other atoll sources are weaker
(Garcia et al. 1988; Fender & Hendry 2000). From infra-red
spectroscopy Bandyopadhyay et al. (1999) conclude GX 13+1
might be a Z source, based on the nature of its companion sug-
gesting a long orbital period.
High resolution spectroscopic observations of GX 13+1
with ASCA (Ueda et al. 2001) and XMM-Newton (Sidoli et al.
R. S. Schnerr et al.: Peculiar behaviour of the LMXB GX 13+1
3
2002), have revealed the presence of narrow absorption lines,
similar to those in the strong galactic jet sources GRS
1915+105 and GRO J1655-40 and the dippers MXB 1659-29
and X 1624-49.
For some Z sources it is known that the position of the
track they trace out in the CD drifts on a time scale of weeks,
with shifts in the colours of up to ∼10% (Kuulkers et al. 1994;
Kuulkers et al. 1996); this effect is called secular motion. It is
also seen in the atoll source 4U 1636-53 (Prins & van der Klis
1997; Di Salvo et al. 2003), but in the CD it is not as strong
as in some Z-sources. In GX 13+1, Muno et al. (2002) also
detected secular motion (see also Homan et al. 1998). These
shifts are usually larger in the HIDs compared to the CDs due
to the shift of the track being strongly correlated with X-ray
luminosity (Lx). The power spectral properties of all LMXBs
seem to depend primarily on the position (S, Hasinger et al.
1990) of the source in the atoll or Z track and not on Lx. Lx
shows a strong correlation (usually positive, sometimes nega-
tive) with this S on short time scales (hours to days), but this
correlation can be completely absent on long time scales. It has
long been assumed that this is because S is a measure of the
accretion rate M, with Lx not proportional to M due to uncer-
tain causes such as, e.g., beaming, but more recently a scenario
has been suggested in which the inner radius of the accretion
disc (rinner) determines the power spectral properties, while at
the same time Lx is determined by the total mass accretion plus
perhaps nuclear burning (van der Klis 2001). The decorrelation
of Lx from the power spectral properties occurs in this scenario
through energy release that does not originate from accretion
through the inner disc, but does respond to inner disc count
rate changes in a time-averaged sense.
Recently, Muno et al. (2002) and Gierli´nski & Done (2002)
suggested that the distinction between atoll and Z sources
might be an artifact of incomplete sampling, because the atoll
sources that cover a wide enough range in count rate (the tran-
sient sources Aql X-1, 4U 1608-52 and 4U 1705-44, each with
a flux range of more than a factor of 10, see also Barret & Olive
2002) also exhibit roughly Z-shaped tracks with at the low-
est luminosities a 'horizontal branch'. However, the differ-
ences in flux ranges and time scales on which these tracks
are traced out in the CDs (as also noted by Muno et al.
2002) as well as their topology and power spectral behaviour
(van Straaten et al. 2003), argue against any simple unification
of these two classes. Being a persistently bright source with a
flux range of a factor of ∼2, GX 13+1 is not among the sources
that has been observed to show such an atoll source 'horizontal
branch'.
In this paper we discuss the X-ray colour, flux and tim-
ing properties of GX 13+1 and their mutual dependence. We
find highly unusual power spectral -- colour-colour diagram be-
haviour, which does not fit in with either Z or atoll behaviour
as well as very unusual secular motion. In Sect. 4.4 we suggest
a possible explanation for the unusual properties we find in GX
13+1.
2. Observations
We used all publicly available Rossi X-ray Timing Explorer
(RXTE, Bradt et al. 1993) Proportional Counter Array (PCA,
Jahoda et al. 1996) data of GX 13+1 in the third RXTE gain
epoch, a total of 44 observations (listed in Table 1) ranging
from May 17 to October 10, 1998 and lasting about 480 ksec
in total. Individual observations typically last ∼10 ksec, in-
terrupted several times by South Atlantic Anomaly passages
and Earth occultations. The uninterrupted parts of observations
are normally ∼3 ksec in duration. In 33 observations the five
Proportional Counter Units (PCUs) 0 -- 4 were all active, but in
11 observations PCU 3 and/or 4 were inactive for part or for the
entire observation. The 2-60 keV count rate for 5 PCUs varied
between ∼3300 -- 5500 cnts s−1. For all observations Standard
2 data are available which have a time resolution of 16 s in 129
energy channels (covering the 2-60 keV PCA range), as well as
high time resolution Single Bit and Event data with a time res-
olution of 1/4096 s or better, also covering the 2-60 keV PCA
range.
Standard 2 data of PCUs 0, 1 and 2 (which were always ac-
tive) were used to create CDs and HIDs with 16 s data points.
The data were background corrected using the Bright Source
Model for the RXTE PCA1. The soft colour (SC) was defined
as the count rate ratio between PCA Standard 2 spectral chan-
nels 7-14 and 3-6 (3.6-6.5 and 2.2-3.6 keV respectively) and
the hard colour (HC) as the ratio between channels 21-40 and
15-20 (8.7-16.0 and 6.5-8.7 keV). The intensity was taken to
be the 2-60 keV (channel 1-129) count rate.
We checked the colours for instrumental drifts caused by
changes in the response of the RXTE PCA by analysing 4 Crab
observations covering the entire observation period of the data
we used. We created CDs of the Crab observations using the
same bands as we used for the GX 13+1 data and calculated
the average colours. The shifts of Crab in the CD are below
1% in SC and below 0.5% in HC. Dead time effects result in
shifts of the data points in the CD relative to one another of less
than ∼0.1% in both HC and SC. Both the effect of instrumental
drifts and of dead time are negligible compared to the effects
reported below.
3. Analysis and results
3.1. The colour-colour and hardness-intensity
diagrams
On cursory examination, the CD of our data appears to exhibit
the lower part of an atoll track or perhaps the NB/FB part of a Z
track, traced out on a time scale of hours to days (Fig. 2) with a
rather sharp vertex. This track shifts on longer time scales, re-
sulting in a CD with overlapping tracks. Similar shifts are also
observed in the HIDs. We divided the data into 15 different
tracks by adding observations together in chronological order
for at most one week, unless the track shifted before then (see
Table 1). A shift in this context was defined as motion of the
track over a colour-colour interval more or less perpendicular
to the track, significant compared to the track width. Clearly,
1 http://rxte.gsfc.nasa.gov/docs/xte/whatsnew/calibration.html
4
R. S. Schnerr et al.: Peculiar behaviour of the LMXB GX 13+1
2
6
1
5
8
9
4
7
12
3
10a
10b
10c
11a
11b
l
r
u
o
o
C
d
r
a
H
Soft Colour
Soft Colour
Fig. 2. The colour-colour diagrams of tracks 1 to 12 (black points); the tracks that show a vertex are plotted in the left plot and
those that do not in the right. For comparison the colour-colour diagram of track 6 is plotted (grey points) in every frame. Notice
the shifts of the track, mainly in the hard colour. For each point we used a 16 s average spectrum to calculate the SC (3.6-6.5
keV/2.2-3.6 keV) and HC (8.7-16.0 keV/6.5-8.7 keV). Typical errors are 0.04 (top left of the CD) up to 0.06 (top right of the
CD) in the SC and around 0.013 in the HC.
In Fig. 2 we show the CDs of all 15 tracks in black su-
perimposed on track 6 (in grey) for reference. As mentioned
above GX 13+1 shows a pattern very reminiscent of a stan-
dard atoll track, with, however, a rather sharp vertex. The
tracks are shifted relative to each other; the vertex locations
are presented in Fig. 4 and Table 1. The shifts are consistent
(χ2/dof = 8.3/7) with being on a diagonal line in the CD.
a shift parallel to the tracks can be overlooked in this way. For
that reason data from our 15 tracks could still be to some extent
mixed up. However, it is not likely that this happened very of-
ten. The reason for this is, that, as shown below, the shifts tend
to occur along one particular diagonal in the CD. The tracks do
not usually run parallel to that diagonal.
Tracks 3, 10a,b,c and 11a,b were excluded from further
analysis, because they do not show a vertex. Because of this
lack of a vertex we cannot with confidence compare the posi-
tions of these tracks in the CD with those of the others.
For each of the remaining 9 tracks we defined a parameter
Sa that indicates the position along the track, in a similar way
as e.g. Dieters & van der Klis (2000) for the Z source Sco X-1.
This was done by approximating the track with a spline onto
which the data points are projected along a vector defined by
their errors in HC and SC (see Fig. 3). Different from e.g Sco
X-1, GX 13+1 only shows one vertex in the CDs instead of two.
We chose the value of Sa at this vertex to be equal to 2, and
fixed the arc length along the track by setting Sa =1 at a point
on the top branch at a fixed HC and SC interval (∆HC=0.3,
∆SC=0.2) above and to the right of the vertex. The locations
of the vertices (the middle of the sharp turn) were estimated by
eye. To quantify the shift of each track in the CD we defined
a parameter Sshift as the distance of the track's vertex to the
point HC=0.46 and SC=3.07 (the vertex of track 1): Sshift ≡
p(∆HC 2 + ∆SC 2).
R. S. Schnerr et al.: Peculiar behaviour of the LMXB GX 13+1
5
aS =1
C
H
S =2a
SC
Fig. 3. The CD of track 6 and the spline used to calculate Sa.
The locations on the curve where Sa = 1 and Sa = 2 are
indicated; the arrow indicates the direction of increasing Sa.
9
5
4
2
6
7
12
1
8
Fig. 4. The vertex locations of the tracks with their track num-
ber and the best linear fit to the points. The vertices are con-
sistent with being on a line of increasing hard and soft colour
(χ2/dof = 8.3/7), spanning a 7% increase in SC and a 43%
increase in HC.
Representative HIDs of track 1 are shown in Fig. 5, together
with the CD. The colour-colour points have different symbols
based on their location in the CD to simplify tracking the differ-
ences in the shape of the track between the diagrams. Contrary
to what is seen in other sources the track is flipped in the HIDs
compared to the CD.
3.2. Dependence of the count rate on Sa and Sshift
In Fig. 6 we show the 2-60 keV count rate plotted ver-
sus Sa for each track. Contrary to what is normally seen
in atoll sources on these timescales (Hasinger & van der Klis
1989; Di Salvo et al. 2001, 2003), the count rate depends non-
monotonically on Sa, with usually a minimum somewhere in
the range Sa=2 -- 3.4. Such a minimum is seen at the NB-FB ver-
tex (Sa=2) in some Z sources (see, e.g, Dieters & van der Klis
2000; Jonker et al. 2000; Homan et al. 2002), but there the min-
imum always occurs at the vertex, while in GX 13+1 its po-
sition drifts between Sa values of 2 and 3 depending on ob-
servation. In Fig. 7 we have plotted the average count rate for
Sa = 2.1 ± 0.1 versus Sshift. From Figs. 6 and 7 a correla-
tion between count rate and Sshift is observable: the two tracks
with the lowest Sshift, tracks 1 and 8, have lower count rates
for a given Sa than the other tracks, except for the lowest three
Sa intervals of track 1. Tracks 9 and 5 which have the highest
Sshift show the highest count rates above Sa ∼ 2.5.
3.3. The power spectra
Power spectra with a Nyquist frequency of 2048 Hz were cal-
culated using 16 s segments of high time resolution data and
all energy channels, i.e., covering the 2-60 keV PCA band.
Each of the 9 tracks separately was divided into (typically 8)
Sa intervals and the power spectra corresponding to the se-
lected colour-colour points were averaged. A constant Poisson
level determined by using the 1024 to 2048 Hz range (where no
kHz QPOs were detected) was subtracted. Three representative
power spectra from track 7 are presented in power×frequency
vs. frequency representation in Fig. 8. The VLFN, BLN and a
QPO are clearly visible.
The power spectra were fitted with 2 -- 4 Lorentzians. We
chose to fit the VLFN with two zero-centered Lorentzians in-
stead of a power law, because it has a "bumpy" structure, which
can not be fitted well with a single power law (to obtain a good
fit one has to add additional components to fit the bumps in
any case) and is too broad a feature to be well fitted by a sin-
gle Lorentzian. We will call these two components the low and
high VLFN. This approach has the additional advantage that it
avoids the common problem that the VLFN power law attempts
to also fit power at high frequencies, which is probably unre-
lated to the VLFN, and can bias the power law to ill represent
the data at low frequencies. Two more Lorentzians fit the BLN
and a QPO, when present. The Lorentzians are reported in the
νmax, Q, r representation used by Belloni et al. (2002), where
νmax is the frequency at which the Lorentzian reaches its max-
imum in a power×frequency vs. frequency power spectral rep-
resentation, Q the quality factor (for which we use the standard
6
R. S. Schnerr et al.: Peculiar behaviour of the LMXB GX 13+1
Fig. 5. The colour-colour diagram and hardness-intensity diagrams of track 1. The colour-colour points have different symbols
based on their location in the CD to simplify tracking the differences in the shape of the track. The count rates are for 5 PCUs.
Typical error bars are shown for two points at the ends of the track in the different diagrams.
definition of centroid frequency divided by full width at half
maximum) and r the fractional rms amplitude. We will regard
νmax to be the characteristic frequency of a Lorentzian com-
ponent. The rms powers were calculated in the 0 -- ∞ Hz range,
except for the two components fitting the VLFN where we used
the 0.125 -- ∞ Hz range because for these components the un-
measured power below our lowest Fourier frequency provides
a significant contribution (which is not the case for the other
components). The rms amplitudes were calculated assuming
an average background of 140 cnts s−1 (2-60 keV, 5 PCUs);
the ∼20 cnts s−1 deviations from this average are negligible in
view of the high source count rates.
In all of the tracks we have found VLFN and BLN and in
5 of the 9 tracks we detected the 57-69 Hz QPO first reported
by Homan et al. (1998). For all Sa intervals we found VLFN;
the BLN usually disappeared in the highest Sa intervals and
QPOs were only detected in the lowest Sa intervals. The power
spectral components and their dependence on Sa and Sshift is
discussed in more detail in Sect. 3.4.
3.4. Dependence of the power spectral properties
on Sa and Sshift
3.4.1. The very low frequency noise
As described in Sect. 3.3, the VLFN was fitted with two
zero-centered (Q = 0) Lorentzians. The power of both these
Lorentzians was added to calculate the total rms amplitude of
the VLFN. For all 9 tracks the rms of the total VLFN generally
decreases with Sa (see Fig. 9), from ∼4.5 -- 5.5% to ∼3 -- 3.5%.
An anti-correlation with Sshift is visible: tracks 5 and 9, which
have a high Sshift, have a VLFN rms that is significantly lower
than that of the other tracks. To quantify this we fitted a straight
line to the VLFN amplitude - Sa relation, and determined the
value of the VLFN amplitude at Sa = 2 from this fit. In Fig.
10 we show this amplitude as a function of Sshift. As Sshift
increases from 0 to ∼0.3 the VLFN rms at Sa = 2 decreases
from ∼4.8% to ∼3%.
The absolute rms of the VLFN as a function of Sa of all
tracks combined is shown in Fig. 11. Up to Sa = 2 the absolute
rms decreases from ∼50 to 30 cnts s−1, and for Sa > 2 it
stays approximately constant at ∼30 cnts s−1. So for Sa <
2 the decreasing absolute rms is correlated to the decreasing
count rate; however, for Sa > 2 the count rate in some tracks
increases, but this does not affect the absolute rms, which stays
approximately constant.
The characteristic frequency (νmax) of the low and high
VLFN does not show a clear correlation with Sa, but varies
between 0.1 -- 0.3 Hz for the low and between 0.5 -- 1.5 Hz for
the high VLFN (with a few exceptions up to 2 Hz when no
BLN was fitted, and for low Sa in track 9 where the BLN also
shows an extremely high νmax).
3.4.2. The band limited noise
The BLN characteristic frequency, shown in Fig. 12, and rms,
shown in Fig. 13, both decrease with Sa. For most tracks the
frequencies decrease from ∼10 to ∼4 -- 5 Hz, except for track 9
which has a BLN frequency of ∼30 Hz for its lowest two Sa
intervals (Sa ∼1.5 -- 1.7). The highest BLN rms is ∼4 -- 6% and
is found for low Sa; when Sa increases the BLN rms decreases
until it becomes undetectable around Sa ∼2.5 -- 3. Typical 95%
confidence upper limits for the BLN for the Sa intervals where
it was not detected are ∼2%, with νmax = 5 Hz and Q = 0.5.
For tracks 9 and 4 the rms increases again with Sa above Sa ∼
3.
The characteristics of the BLN do not show a clear corre-
lation with Sshift, although the BLN in track 9 has a higher
rms for a given Sa than in the other tracks and also shows ex-
tremely high frequencies for low Sa (see above). The Q of the
BLN is also not significantly correlated with Sa. It varies be-
tween 0 and 1, with the exception of the last interval of track 2
(Sa = 3.13 ± 0.23, Q = 2.6+22.1
−1.7 ) and Sa interval 7 of track
9 (Sa = 2.94 ± 0.05, Q = 4+4
−2), where the BLN also shows a
very high frequency of 12.1+0.7
−0.8 Hz; in these two cases (only)
this component is sufficiently narrow to be called a QPO.
R. S. Schnerr et al.: Peculiar behaviour of the LMXB GX 13+1
7
3.4.3. Quasi-periodic oscillations
QPOs were detected with frequencies ranging from ∼60 -- 75 Hz
and Q values of ∼2 -- 5 (detailed properties are shown in Table
2). They are only detected at low Sa values, with the highest
rms at the lowest Sa, and in too small a range of Sa to say
anything about the dependence of their frequency and Q on
Sa. The fact that QPOs were detected in tracks 1 and 8 and not
in tracks 5 and 9 could be related to the high Sshift values of
tracks 5 and 9, but this could also be due to the fact that these
tracks do not reach the upper-left most part of the track, where
the QPOs are found. The frequencies of the detected QPOs are
all consistent with being in the range of those previously found
by Homan et al. (1998).
We have searched for kHz QPOs up to 2048 Hz, but none
were detected above the 3σ confidence level. The strongest nar-
row features (Q ∼ 10) above 300 Hz were between 2 and
3σ and resulted in 3σ upper limits of up to 3% rms. A broad,
3.1σ, high frequency feature (2.6 ± 0.5 % rms, Q = 0.9+0.9
−0.6,
νmax = 858+225
−137 Hz) is found in track 7 (Sa = 1.50 ± 0.08),
but when the number of trials is taken into account this is not a
significant detection.
8
R. S. Schnerr et al.: Peculiar behaviour of the LMXB GX 13+1
=1
=12
=4
=6
=9
=8
=2
=7
=5
Fig. 6. The count rate vs Sa for all the 9 tracks, distributed over two plots for clarity. The track number and corresponding symbol
are shown in the figure. For all tracks the count rate decreases up to Sa=2 -- 3.4 and in some tracks it increases again for Sa >2 -- 3.4.
Fig. 7. The count rate at Sa = 2 for all 9 tracks plotted vs.
Sshift. The two tracks with the lowest Sshift (tracks 1 and 8)
show the lowest count rates.
R. S. Schnerr et al.: Peculiar behaviour of the LMXB GX 13+1
9
low VLFN
high VLFN
QPO
BLN
Fig. 8. Representative power spectra of GX 13+1. Shown are power spectra of track 7 with an averaged Sa of 1.50 ± 0.08 (left;
χ2/dof = 125/145), 2.01 ± 0.06 (middle; χ2/dof = 135/148) and 2.73±0.12 (right; χ2/dof = 138/151). All power spectral
features become weaker with increasing Sa. The solid line represents the fit and the dashed and dotted lines the Lorentzian
components. The VLFN is fitted by the long dashed line (low VLFN) and the dotted line (high VLFN), the BLN by the dash-
single dot line and the QPO by the dash-triple dot line. In the left power spectrum some excess power is present in the 200-2000
Hz range, but no significant kHz QPO is detected.
Fig. 9. VLFN fractional rms amplitude vs Sa for all 9 tracks; the amplitude plotted is that of the combined low and high VLFN,
integrated from 0.125 to ∞ Hz. For all tracks the amplitude of the VLFN generally decreases with Sa, and tracks 5 (five-pointed
stars, right frame) and 9 (six-pointed stars, left frame), which have the highest Sshift, have the lowest amplitude for a given Sa.
Symbols are the same as in Fig. 6.
10
R. S. Schnerr et al.: Peculiar behaviour of the LMXB GX 13+1
Fig. 10. The VLFN amplitude at Sa=2 for all 9 tracks (as deter-
mined from a linear fit to a plot of the VLFN rms vs. Sa), as a
function of Sshift. The amplitude decreases from ∼4.8% to ∼3%
as Sshift increases from 0 to ∼0.3.
Fig. 11. The absolute amplitude of the VLFN as a function of Sa
for all 9 tracks combined. Up to Sa=2 the absolute rms ampli-
tude decreases from ∼50 to 30 cnts s−1, and for Sa >2 it stays
approximately constant at ∼30 cnts s−1.
R. S. Schnerr et al.: Peculiar behaviour of the LMXB GX 13+1
11
Fig. 12. The characteristic frequency of the BLN vs Sa for all 9 tracks. The characteristic frequency generally decreases with
Sa. Track 9 shows an extremely high frequency for its lowest two Sa intervals. Symbols are the same as in Fig. 6; in some Sa
intervals no BLN was detected.
Fig. 13. The fractional rms amplitude of the BLN vs Sa for all 9 tracks. The amplitude of the BLN generally decreases with Sa
for all tracks, but in track 4 (diamonds, left frame) and 9 (six-pointed stars, left frame) it increases again for the highest two Sa
intervals. 3σ Upper limits were determined with Q = 0.5 and νmax = 5 Hz, and are represented by the points with only a a
negative error bar extending to the bottom of the plot. Symbols are the same as in Fig. 6
12
4. Discussion
R. S. Schnerr et al.: Peculiar behaviour of the LMXB GX 13+1
4.1. Extreme secular motion in CD and HIDs
We have studied the secular motion both in the CD and in the
HIDs of GX 13+1. The largest shifts are observed in the direc-
tion of HC (∼40%) and not in SC (∼7%). In other atoll and Z
sources which show secular motion, shifts of less than 3% in
HC and ∼4% in SC (4U 1636-53, Di Salvo et al. 2003) and
6% in HC and 8% in SC (GX 5-1, Kuulkers et al. 1994) are
observed; in Cyg X -- 2 the shifts in both HC and SC are up to
∼10% (Kuulkers et al. 1996). This is not strongly dependent
on the choice of bands. The shifts in the CD of GX 13+1 are
remarkably large, particularly in HC, even when compared to
Z sources, where the shifts are larger than in atoll sources. As
is also seen in other sources (references above), the shift of the
track shows a correlation with average count rate resulting in
shifted tracks in the HIDs. We find that in GX 13+1 the am-
plitude of the VLFN also shows a correlation with this shift;
it decreases when the track is shifted towards higher HC and
SC. Such a correlation is not observed in the Z source GX 5-
1, but in the atoll source 4U 1636-53 the part of the track that
shifted to higher SC and intensity also has a lower VLFN rms.
The VLFN rms in Cyg X -- 2 shows the same dependence on
the shift of the track, but the overall intensity level of a track
decreases with this shift (Kuulkers et al. 1999). In GX 13+1,
track 9, which has the largest shift also has unusual BLN prop-
erties (Sect. 3.4.2); in 4U 1636-53 an extra, broad component
(Q = 0.8) is detected next to the BLN in the high intensity part
of the track at ∼ 1.1% rms, but not in the low intensity part
(with 1σ upper limits of ∼0.6%).
4.2. The flipping effect
In the HIDs with the HC or SC on the vertical, and count rate
on the horizontal axis the arc-shape found in the CD is re-
versed, i.e., the source moves through the arc in the opposite
sense in the HIDs as compared to the CD. This is different from
what is usually observed in atoll sources (Wijnands et al. 1999;
van Straaten et al. 2000; Muno et al. 2002; Di Salvo et al.
2003), where, on short timescales, the same basic pattern is
traced out in both the CD and the HIDs and the sense of mo-
tion is always the same. On longer timescales secular shifts in
the HIDs are sometimes observed (van Straaten et al. 2003). In
most Z sources the observed basic pattern in the CD and HIDs
is also similar (Schulz et al. 1989; Kuulkers & van der Klis
1996; Jonker et al. 2002; Homan et al. 2002), although for
some sources the count rate decreases along (a part of) the FB,
causing this branch to be in the opposite direction in the HIDs
compared to the CD (Kuulkers et al. 1996). The cause for this
unusual behaviour of GX 13+1 in the HIDs is that the count rate
depends on Sa in a different way than in most other sources (see
Sect. 3.2). A possible explanation for this will be discussed in
Sect. 4.4.
4.3. GX 13+1 in the Z/atoll scheme
The tracks that GX 13+1 traces out in the CD are, at first sight,
very reminiscent of a classic island/banana in an atoll source, or
a normal-branch/flaring-branch in a Z source. However, from
the results presented in Sect. 3 we find that the power spectral
properties of this source as a function of Sa resemble neither
those of a classic atoll, nor of a Z source.
In a given track the count rate of GX 13+1 decreases along
the track up to Sa=2 -- 3 after which (in some tracks) it increases
again with Sa, while in atoll sources on short timescales the
count rate generally increases monotonically with Sa along
the whole track (see Sect. 3.2; for references on atoll source
behaviour see van Straaten et al. 2002; Di Salvo et al. 2003;
Hasinger & van der Klis 1989). This is also different from what
is observed in Z sources, where the count rate can slightly
decrease along the NB, but increases immediately after the
NB/FB vertex so that the minimum in count rate is at the ver-
tex (see Sect. 3.2; for references on Z source behaviour see
Dieters & van der Klis 2000; Jonker et al. 2000; O'Neill et al.
2001; Jonker et al. 2002; Homan et al. 2002). The VLFN of
GX 13+1 is very strong (usually ∼4 -- 5.5% rms for Sa ≤2)
compared to atoll sources in the island state or lower banana
branch and Z sources in the NB (the VLFN frequency range
is similar). The VLFN rms of GX 13+1 generally decreases
with Sa. This is the opposite of what is normally seen in both
atoll and Z sources, where the VLFN increases with Sa along
the whole track (see, however, Wijnands et al. 1999, on the
atoll source 4U 1820-30). The BLN is quite weak (∼2 -- 5%)
compared to that in atoll sources in the island state and lower
banana and its rms decreases with Sa as is usual in atoll and
Z sources, but in tracks 4 and 9 it increases again in the high-
est Sa intervals. The characteristic frequency of the BLN also
decreases with Sa, which, in either atoll sources in the island
state and lower banana branch or Z sources, has so far only
been observed in the atoll source 4U 1820-30, where a BLN-
like component that evolves into or is replaced by a ∼7 Hz
QPO (Wijnands et al. 1999) is seen. In the Z source GX 349+2,
a similar noise component is seen of which the centroid fre-
quency decreases with Sa, but the νmax of this component
shows a more complicated dependence on Sa (O'Neill et al.
2002). Generally, in atoll sources the νmax of the BLN in-
creases with Sa up to the upper banana, where it starts to de-
crease again, whereas in Z sources it increases monotonically.
We find the QPO properties of GX 13+1 to be unlike ei-
ther Z or atoll sources. At the lowest Sa intervals the 57-69 Hz
QPO found by Homan et al. (1998) is detected. In atoll sources
a low frequency QPO is usually detected in the island state with
a frequency of 10-15 Hz (see, however, Wijnands et al. 1999)
,increasing up to 40-47 Hz towards the upper banana, where it
becomes undetectable (although a second low frequency QPO
is sometimes also seen at similar frequencies, but for higher
Sa, see van Straaten et al. 2002). If the QPO would remain de-
tectable, one would expect to find such a 57-69 Hz QPO at
higher Sa, in the middle banana branch. The HBO in Z sources
has a frequency up to ∼55 Hz in the upper NB; if the QPO in
GX13+1 is interpreted as this QPO one would also expect to
find a ∼6 Hz NBO on this branch; in the Z source GX 349+2
R. S. Schnerr et al.: Peculiar behaviour of the LMXB GX 13+1
13
no NBO is found, but a very broad noise component is seen in-
stead (Kuulkers & van der Klis 1998; O'Neill et al. 2001), and
of this GX 13+1 shows no evidence either. The fact that no kHz
QPOs are detected anywhere along the branch (see Sect. 3.4.3)
is uncharacteristic for both atoll and Z sources.
Clearly, GX 13+1 is a peculiar source in which the proper-
ties correlate to Sa in a very unusual way when compared both
to atoll and Z sources. In the final section we explore a possible
explanation for this unusual behaviour.
4.4. A jet scenario
If the behaviour of GX 13+1 is interpreted within the frame-
work of atoll source behaviour, then clearly for high Sa the
source is in the upper banana. Based on the CD, at low Sa one
would expect the source to be in the lower left banana and is-
land state. However, the fact that for low Sa 57-69 Hz QPOs,
and weak BLN with a characteristic frequency that decreases
with Sa are found, suggests that, contrary to what one would
expect based on the CD, the "true" state of the source there
is the middle part of the banana branch, because that is where
other atoll sources can be expected to show this kind of power
spectral behaviour. This is supported by the fact that so far no
kHz QPOs have been found in this source, consistent with their
absence in other atoll sources in the middle and upper parts of
the banana branch.
In GX 13+1 the count rate decreases as Sa increases up to
Sa=2 -- 3, and only for higher Sa increases, which is what in a
regular atoll source one would expect along the whole track
(see Sect. 3.2). One interpretation for this unusual behaviour
would be that there is something adding extra flux at low Sa,
which becomes weaker towards higher Sa resulting in a count
rate that first decreases, and then increases as the standard atoll
upper banana behaviour takes over. From the two branched
structure in the CDs one can conclude that this extra flux would
have to be hard, because for low Sa the source spectrum gets
harder when the count rate increases. This scenario would also
explain why the BLN is so weak and shows a decreasing νmax
along the track, and why no kHz QPOs are detected at low Sa:
these properties are normal for the middle and upper parts of
the banana branch, which is, by hypothesis, the true state of
the source. In most atoll sources the low frequency QPO disap-
pears at the same Sa, or a slightly lower value, as the kHz QPO
(van Straaten et al. 2002; Di Salvo et al. 2003); the fact that in
GX 13+1 we still see the low frequency QPO while the kHz
QPO is absent is unusual, but in line with the relatively high
frequency of the low frequency QPO. So the hypothesis we are
exploring is that the hard extra flux creates an anomalous sec-
ond branch in the CD reminiscent of a lower banana and island
branch, while the source remains in the "middle" banana state.
A possible source for this extra, hard flux could be a rela-
tivistic jet. In general in atoll and Z sources radio emission, as-
cribed to the presence of a relativistic jet (Fender 2002), tends
to increase towards lower Sa. If this more intense jet is associ-
ated with the extra hard flux, this could cause a second branch
in the CD, while the source is still in the upper banana state.
We further hypothesize that the unusually strong VLFN would
occur in the hard jet flux. Consequently, when the jet becomes
weaker towards higher Sa this causes the VLFN to decrease in
amplitude. This is in accordance with the fact that up to Sa=2
the absolute rms amplitude of the VLFN decreases, from which
it can be concluded that it is the flux, that also decreases in this
part of the track, that shows relatively strong VLFN that dis-
appears. For Sa >2 the absolute rms amplitude of the VLFN
stays approximately constant, while in some tracks the flux in-
creases for Sa >3. The extra flux in these tracks does lead to a
significant increase of the absolute rms amplitude.
The question of course is: why would a jet in GX 13+1
cause this peculiar behaviour and not in other atoll sources?
The radio emission of GX 13+1 is unusually strong for an atoll
source (Fender & Hendry 2000), so perhaps for a reason un-
known the jet is unusually strong in this source. A more at-
tractive possibility is that it is not the jet that is unusual, but
the system's orientation. If the jet is pointing directly towards
us Doppler boosting could significantly increase its radio flux
(e.g. Fender & Hendry 2000); a strongly relativistic jet (which
has already been directly imaged for the Z source Sco X-1,
Fomalont et al. 2001a,b) could easily Doppler boost the radio
emission by more than an order of magnitude, putting the radio
emission of GX 13+1 in accordance with measurements and
upper limits of other atoll sources (Fender & Hendry 2000).
The unusual X-ray phenomena could then for example be the
consequence of a better view of the X-ray emitting regions as-
sociated with the base of the jet, offered by the unusual ori-
entation of the disk, possibly assisted by Doppler boosting as
well. A similar possibility of Doppler-boosted jet X-ray emis-
sion causing hard flaring ("microblazar") was recently explored
for Cyg X-1 by Romero et al. (2002).
The large shifts of the track in the CD of GX 13+1 com-
pared to other atoll and Z sources might be related to this better
view of the X-ray emitting base of the jet region as well. This
would require this region to contribute different baseline levels
of hard flux at different epochs in order to produce the secular
motion (which, consistent with this, takes place mostly in the
HC), VLFN whose amplitude anticorrelates, as well as extra
flux at Sa > 3 which correlates, with this baseline flux level.
This could for example be related to secular geometry changes
(e.g. precession) modulating our view of the base of the jet, or
to variations in the jet activity itself.
One can also attempt to explain the peculiar behaviour of
GX 13+1 with a similar relativistic jet scenario assuming the
source is a Z source. The QPO observed in GX 13+1 would
then be interpreted as the HBO, occurring at the top of the NB,
leading to the conclusion that the CD shows both a NB and
a FB. The strong VLFN that decreases along the whole track
might be related to a jet similarly to the atoll scenario. This
jet could be related to the high radio over X-ray luminosity
ratio compared to other Z sources. However, this model leaves
unexplained the decreasing frequency of the BLN with Sa and
the absence of an NBO.
So, when a choice is made between an atoll or a Z clas-
sification for GX 13+1 (in either case modified by an unusual
jet geometry) the overall phenomenology favours the former
option. An alternative would be that GX 13+1 combines some
properties of atoll and Z sources (as noted in Sect. 1, GX 13+1
14
R. S. Schnerr et al.: Peculiar behaviour of the LMXB GX 13+1
has been described as the atoll source closest to the Z sources).
This possibility will be further explored in a forthcoming paper
(Homan et al. 2003).
It has been suggested that narrow absorption features, as
found in GX 13+1 (Ueda et al. 2001), are related to a high in-
clination of the source, contrary to the low inclination predicted
by the scenario discussed above. However, it is also possible
that the absorption features are not related to the inclination of
the source, but, as previously suggested by Ueda et al. (1998),
to a highly ionized plasma associated with their relativistic jets,
causing absorption from the flux from the hot inner accretion
region.
Several predictions can be made based on the atoll
source+jet hypothesis. Since the unusually strong VLFN oc-
curs in the jet flux, one expects the VLFN, at low Sa, to be the
strongest in the hard flux. Also, as the jet is the main source
of radio emission in LMXBs, we expect the radio emission to
be strongest in the upper part of the second branch in the CD
(at low Sa), decreasing as the source moves towards the vertex.
This can be tested with simultaneous radio and X-ray obser-
vations. One would have to account for a delay factor in the
radio emission of perhaps up to half an hour (Klein-Wolt et al.
2002) associated with the expansion of the synchrotron clouds.
Such observations have been done by Garcia et al. (1988), but
with a small time range (∼5 hours of overlapping X-ray and
radio observations) and no CD analysis was performed, and by
Berendsen et al. (2000), also with a short period of overlapping
X-ray and radio observations. Recent simultaneous radio and
X-ray observations also exist (to be reported in Homan et al.
2003), and preliminary results of this work that became avail-
able after the current analysis and interpretation were com-
pleted, appear to confirm our suggestion.
Acknowledgements. RS and MK acknowledge support from the
Netherlands Organisation for Scientific Research (NWO).
References
Ford, E. C., van der Klis, M., M´endez, M., et al. 2000, ApJ,
537, 368
Garcia, M. R., Grindlay, J. E., Molnar, L. A., et al. 1988, ApJ,
328, 552
Gierli´nski, M. & Done, C. 2002, MNRAS, 331, L47
Hasinger, G. & van der Klis, M. 1989, A&A, 225, 79
Hasinger, G., van der Klis, M., Ebisawa, K., Dotani, T., &
Mitsuda, K. 1990, A&A, 235, 131
Homan, J., van der Klis, M., Jonker, P. G., et al. 2002, ApJ,
568, 878
Homan, J., van der Klis, M., Wijnands, R., Vaughan, B., &
Kuulkers, E. 1998, ApJ, 499, L41
Homan, J., Wijnands, R., Fender, R., et al. 2003, in prep.
Jahoda, K., Swank, J. H., Giles, A. B., et al. 1996, in Proc.
SPIE Vol. 2808, p. 59-70, EUV, X-Ray, and Gamma-Ray
Instrumentation for Astronomy VII, Oswald H. Siegmund;
Mark A. Gummin; Eds., 59 -- 70
Jonker, P. G., van der Klis, M., Homan, J., et al. 2002, MNRAS,
333, 665
Jonker, P. G., van der Klis, M., Wijnands, R., et al. 2000, ApJ,
537, 374
Klein-Wolt, M., Fender, R. P., Pooley, G. G., et al. 2002,
MNRAS, 331, 745
Kuulkers, E. & van der Klis, M. 1996, A&A, 314, 567
-- . 1998, A&A, 332, 845
Kuulkers, E., van der Klis, M., Oosterbroek, T., et al. 1994,
A&A, 289, 795
Kuulkers, E., van der Klis, M., & Vaughan, B. 1996, A&A,
311, 197
Kuulkers, E., Wijnands, R., & van der Klis, M. 1999, MNRAS,
308, 485
Lewin, W. H. G., van Paradijs, J., Hasinger, G., et al. 1987,
MNRAS, 226, 383
Matsuba, E., Dotani, T., Mitsuda, K., et al. 1995, PASJ, 47, 575
Muno, M. P., Remillard, R. A., & Chakrabarty, D. 2002, ApJ,
568, L35
O'Neill, P. M., Kuulkers, E., Sood, R. K., & Dotani, T. 2001,
A&A, 370, 479
O'Neill, P. M., Kuulkers, E., Sood, R. K., & van der Klis, M.
Bandyopadhyay, R. M., Shahbaz, T., Charles, P. A., & Naylor,
2002, MNRAS, 336, 217
T. 1999, MNRAS, 306, 417
Barret, D. & Olive, J. 2002, ApJ, 576, 391
Belloni, T., Psaltis, D., & van der Klis, M. 2002, ApJ, 572, 392
Berendsen, S. G. H., Fender, R., Kuulkers, E., Heise, J., & van
Prins, S. & van der Klis, M. 1997, A&A, 319, 498
Romero, G. E., Kaufman Bernad`o, M. M., & Mirabel, I. F.
2002, A&A, 393, L61
Schulz, N. S., Hasinger, G., & Truemper, J. 1989, A&A, 225,
der Klis, M. 2000, MNRAS, 318, 599
48
Bradt, H. V., Rothschild, R. E., & Swank, J. H. 1993, A&AS,
Sidoli, L., Parmar, A. N., Oosterbroek, T., & Lumb, D. 2002,
97, 355
A&A, 385, 940
Di Salvo, T., M´endez, M., van der Klis, M., Ford, E., & Robba,
Stella, L., White, N. E., & Taylor, B. G. 1985, in Recent Results
N. R. 2001, ApJ, 546, 1107
Di Salvo, T., M´endez, M., & van der Klis, M. 2003, accepted
Dieters, S. W. & van der Klis, M. 2000, MNRAS, 311, 201
Fender, R. 2002, in LNP Vol. 589: Relativistic Flows in
Astrophysics, 101
Fender, R. P. & Hendry, M. A. 2000, MNRAS, 317, 1
Fleischman, J. R. 1985, A&A, 153, 106
Fomalont, E. B., Geldzahler, B. J., & Bradshaw, C. F. 2001a,
ApJ, 553, L27
-- . 2001b, ApJ, 558, 283
on Cataclysmic Variables, 125 -- 128
Ueda, Y., Asai, K., Yamaoka, K., Dotani, T., & Inoue, H. 2001,
ApJ, 556, L87
Ueda, Y., Inoue, H., Tanaka, Y., et al. 1998, ApJ, 492, 782
van der Klis, M. 1995, in X-Ray Binaries, ed. W. Lewin, J. van
Paradijs, & E. van den Heuvel (Cambridge University Press),
252 -- 307
van der Klis, M. 2000, ARA&A, 38, 717
van der Klis, M. 2001, ApJ, 561, 943
R. S. Schnerr et al.: Peculiar behaviour of the LMXB GX 13+1
15
van Straaten, S., Ford, E. C., van der Klis, M., M´endez, M., &
Kaaret, P. 2000, ApJ, 540, 1049
van Straaten, S., van der Klis, M., di Salvo, T., & Belloni, T.
2002, ApJ, 568, 912
van Straaten, S., van der Klis, M., & M´endez, M. 2003, sub-
mitted
Wijnands, R. 2001, Advances in Space Research, 28, 469
Wijnands, R., van der Klis, M., & Rijkhorst, E. 1999, ApJ, 512,
L39
16
Track
nr
1
2
3
4
5
6
7
8
9
10a
10b
10c
11a
11b
12
Vertex
SC
3.07
3.13
-
3.10
3.27
Sshift
0.00
0.092
-
0.085
0.249
HC
0.46
0.53
-
0.54
0.60
768.2±43.3
857.1±47.8
-
864.2±53.7
824.4±36.2
0.51
3.14
0.086
829.2±49.9
0.51
3.14
0.086
837.3±50.7
0.46
3.08
0.010
756.0±50.0
0.66
3.30
0.305
842.7±20.6
-
-
-
-
-
-
-
-
-
-
-
-
-
-
-
-
-
-
-
-
0.49
3.14
0.076
829.7±33.6
30051-01-01-01
30051-01-02-00
30051-01-03-01
30051-01-04-00
30051-01-05-00
30051-01-06-00
30051-01-07-01
30051-01-08-00
30050-01-01-00
30050-01-01-02
30050-01-01-04
30050-01-01-05
30050-01-01-07
30050-01-01-08
30050-01-02-01
30050-01-02-03
30051-01-09-00
30051-01-10-00
30051-01-11-00
30051-01-11-02
30051-01-12-00
30051-01-12-01
30051-01-12-02
30050-01-03-00
30050-01-03-01
30050-01-03-02
30050-01-03-03
R. S. Schnerr et al.: Peculiar behaviour of the LMXB GX 13+1
Flux at Sa = 2
(cnts s−1 PCU−1)
Observation ID
of included observations
tstart (UTC)
tend (UTC)
month-day hours:min
05-17 11:01
05-21 12:39
05-17 20:57
05-24 19:21
05-28 08:33
06-01 09:60
06-05 05:16
05-28 14:30
06-01 17:40
06-09 12:42
tobs
(ks)
28
29
2
28
25
06-13 01:26
06-17 15:29
110
06-17 15:58
06-20 10:19
69
30051-01-01-00
30051-01-03-00
30051-01-03-02
30051-01-04-01
30051-01-07-00
30051-01-07-02
30051-01-08-01
30050-01-01-01
30050-01-01-03
30050-01-01-050
30050-01-01-06
30050-01-01-080
30050-01-02-00
30050-01-02-02
30051-01-09-01
06-26 09:34
06-29 16:09
30051-01-11-01
30051-01-11-03
07-03 14:22
07-09 15:40
07-13 09:36
07-31 15:11
08-09 13:31
08-14 13:18
08-21 13:02
10-10 01:45
07-13 15:39
07-31 18:59
08-09 17:25
08-14 17:55
08-21 16:11
10-10 06:44
41
78
22
14
14
17
11
18
Table 1. The RXTE observation IDs of the observations that are included in each track, the HC and SC of their vertices, and the
distance, Sshift, over which each track is shifted relative to track 1. Average flux at the vertex (Sa = 2), start time of the first and
end time of the last included observations and the total included observation time are also indicated. Note that tracks 3, 10a-c and
11a-b do not show a vertex. All dates are in the year 1998; typical errors in HC are ∼ 0.01, in SC∼ 0.03, in Sshift ∼ 0.01 and in
flux at Sa = 2 ∼ 43.
Track
Sa
1
1
2
2
2
6
6
7
7
7
8
8
1.38±0.03
1.47±0.04
1.51±0.05
1.64±0.04
1.76±0.03
1.41±0.05
1.60±0.07
1.31±0.05
1.50±0.08
1.76±0.08
1.16±0.06
1.38±0.08
Sign.
(σ)
4.3
3.8
3.7
3.5
2.7
7.4
6.7
6.7
11.2
2.7
10.0
4.4
νmax
(Hz)
66+5
−4
66+5
−4
69±2
73+4
−3
74±3
66.2+1.2
−1.3
65.5+1.6
−1.5
63.9±0.9
64.2±0.7
64.3+5.7
−1.4
64.1+1.1
−1.0
62±3
Amplitude
(% rms)
2.3±0.3
1.9±0.3
1.8±0.3
1.5±0.2
1.3±0.3
2.7±0.2
2.05+0.15
−0.16
2.20±0.17
2.37+0.10
−0.11
1.0+0.4
−0.2
2.75+0.13
−0.14
2.2+0.2
−0.3
Q
2.0+0.8
−0.6
2.2+1.4
−0.8
4.5+2.3
−1.5
3.7+2.7
−1.5
5+6
−2
4.0+0.9
−0.7
3.7+0.9
−0.7
4.4+1.0
−0.8
3.5+0.5
−0.4
9+36
−7
3.0+0.4
−0.3
2.6+1.2
−0.7
Table 2. All detected QPOs and their properties. Typical 95%
confidence upper limits (νmax = 68 Hz, Q = 3) for tracks in
which the QPO was not detected are up to ∼2% rms.
|
astro-ph/9811428 | 1 | 9811 | 1998-11-27T05:21:09 | 3D Simulation of the Gas Dynamics in the Central Parsec of the Galaxy | [
"astro-ph"
] | It is thought that many characteristics of the gaseous features within the central parsec of our Galaxy, are associated with the accretion of ambient plasma by a central concentration of mass. Using a 3D hydrodynamical code, we have been simulating this process in order to realistically model the gaseous flows in the center of our Galaxy. In the most recent simulation, we have taken into account the multi-point-like distribution of stellar wind sources, as well as the magnetic heating and radiative cooling of these stellar winds. As expected, we find that the structure of the flow is significantly different from that due to a uniform medium. We also investigate the possibility that Sgr A* is due to a distributed mass concentration instead of the canonical point mass of a black hole. We discuss the physical state of the accreting gas and how our results suggest that Sgr A* is unlikely to be associated with a ``dark cluster''. | astro-ph | astro-ph |
3D Simulation of the Gas Dynamics in the Central Parsec
of the Galaxy
R.F. Coker1
Physics Department, The University of Arizona, Tucson, AZ 85721,
USA
Fulvio Melia2
Physics Department and Steward Observatory, The University of
Arizona, Tucson, AZ 85721, USA
Abstract.
It is thought that many characteristics of the gaseous fea-
tures within the central parsec of our Galaxy, are associated with the
accretion of ambient plasma by a central concentration of mass. Using a
3D hydrodynamical code, we have been simulating this process in order
to realistically model the gaseous flows in the center of our Galaxy. In
the most recent simulation, we have taken into account the multi-point-
like distribution of stellar wind sources, as well as the magnetic heating
and radiative cooling of these stellar winds. As expected, we find that
the structure of the flow is significantly different from that due to a uni-
form medium. We also investigate the possibility that Sgr A* is due to
a distributed mass concentration instead of the canonical point mass of
a black hole. We discuss the physical state of the accreting gas and how
our results suggest that Sgr A* is unlikely to be associated with a "dark
cluster".
1.
Introduction
The hydrodynamics of the interstellar medium (ISM) within the central parsec of
the Galaxy has long been thought to be dominated by the gravitational potential
due to a central mass concentration (for a review, see Mezger, Duschl, & Zylka,
1996). For example, the Western Arc, one of the dominant kinematic features
of the region, appears to be in circular rotation about Sgr A*, a unique compact
radio source whose lack of proper motion suggests it lies at the dynamical heart
of the Galaxy (Roberts & Goss, 1993; Backer, 1994). Stellar motions (Genzel,
et. al., 1997), gas kinematics (Herbst, et. al., 1993), and velocity dispersion
measurements (Eckart & Genzel, 1998) together suggest the presence of ∼ 2.6 ×
106M⊙ of dark mass located within ∼ .01 parsecs of Sgr A*. Note that the
1NASA GSRP Fellow.
2Presidential Young Investigator.
1
Galactic center (GC) is at a distance of ∼ 8 kpc so that 1 arcsecond ∼ 0.04
parsecs.
However, showing that the GC must contain a centralized dark mass con-
centration does not necessarily imply that it is in the form of a single compact
object nor does it imply that Sgr A* must be associated with it. Stellar kine-
matic arguments (Genzel et. al., 1996) rule out a distribution of neutron stars
or white dwarfs. One possibility is that the dark mass distribution consists of
∼ 10M⊙ black holes; cluster evolution calculations (Lee, 1995) have shown that
this is at least feasible although stability arguments (Maoz, 1998) suggest the
cluster lifetime would be ∼ 108 years, considerably less than the age of the
Galaxy. In this paper, we wish to determine if, stability arguments aside, the
spectrum of Sgr A* could be due to a plasma trapped within the potential well
of a dark cluster of arbitrary objects.
In addition to large scale gaseous features, there is ample evidence for the
existence of rather strong stellar winds in and around Sgr A* itself. The key wind
sources appear to be the cluster of mass-losing, blue, luminous stars comprising
the IRS 16 assemblage, located within several arcseconds of Sgr A*. A variety
of observations over the years (for a review see Morris & Serabyn, 1996) provide
clear evidence of a hypersonic wind, with a velocity of vw ∼ 500 − 1000 km
Mw ∼
s−1, a number density nw ∼ 103−4 cm−3, and a total mass loss rate
3 − 4 × 10−3M⊙ yr−1 pervading the inner parsec of the Galaxy.
If the dark
matter is distributed, it is likely that a portion of this wind is captured by
the dark compact cluster and that it settles within the cluster's potential well.
Although the potential well of a cluster does not include a cusp such as that due
to a black hole, the trapped plasma might conceivably still account for at least
some of Sgr A*'s radiative characteristics.
In §2 we discuss the simulation of the gas flow through a distributed dark
mass cluster. In §3 we describe the resulting spectrum and present some prelim-
inary semi-analytical spectral calculations based on Sgr A* being a point mass
while in §4 we summarize our analysis.
2. The 3D Hydrodynamical Model
In the classical Bondi-Hoyle scenario (Bondi & Hoyle, 1944), the mass accretion
rate for a uniform hypersonic adiabatic gas flowing past a centralized mass is
MBH = πRA
2mHnwvw ,
(1)
2 is the accretion radius and M is the mass of the central-
where RA ≡ 2GM/vw
ized object(s). At the GC, for the conditions described in the Introduction, we
would therefore expect an accretion rate MBH ∼ 1021−22g sec−1, with a capture
radius RA ∼ 0.01 − 0.02 pc. Since this accretion rate is sub-Eddington for a
∼ one million solar mass concentration, the accreting gas is unimpeded by the
escaping radiation field and is thus essentially in hydrodynamic free-fall starting
at RA. Our initial, simplistic, numerical simulations of this process, where we
assume a point object and uniform flow (Ruffert & Melia, 1994; Coker & Melia,
1996) have verified these expectations.
2
2.1. The Stellar Wind Sources
The GC wind, however, is unlikely to be uniform since the winds from many stars
contribute to the mass ejection. We assume that the early-type stars enclosed
(in projection) within the Western Arc, the Northern Arm, and the Bar produce
the observed wind. Thus far, 25 such stars have been identified (Genzel, et. al.,
1996), though the stellar wind characteristics of only 8 have been determined
from their He I line emission (Najarro, et. al., 1997); the relavant characteristics
of these 25 stars are summarized in Table 1. Two sources, IRS 13E1 and IRS 7W,
seem to dominate the mass outflow with their high wind velocity (∼ 1000 km
sec−1) and a mass loss rate of more than 2 × 10−4 M⊙ yr−1 each. Unfortunately,
the temperature of the stellar winds is not well known, and so for simplicity we
have assumed that all the winds are Mach 30; this corresponds to a temperature
of 104−5K. In addition, for the sources that are used in these calculations, their
location in z (i.e., along the line of sight) is determined randomly with the
condition that the overall distribution in this direction approximately matches
that in x and y.
Table 1.
Parameters for Galactic Center Wind Sources
Star
IRS 16NE
IRS 16NW
IRS 16C
IRS 16SW
IRS 13E1
IRS 7W
AF
IRS 15SWb
IRS 15NEb
IRS 29Nc
IRS 33Ec
IRS 34Wc
IRS 1Wc
IRS 9NWcd
IRS 6Wc
AF NWcd
BLUMb
IRS 9Sb
Unnamed 1b
IRS 16SEb
IRS 29NEb
IRS 7SEb
Unnamed 2b
IRS 7Eb
AF NWWb
xa(arcsec)
ya(arcsec)
za(arcsec)
v (km sec−1)
M(10−5 M⊙ yr−1)
5.5
7.3
-7.1
4.9
-1.5
-5.1
8.5
3.5
1.5
-6.4
7.8
-3.8
3.6
-2.1
550
750
650
650
1000
1000
700
700
750
750
750
750
750
750
750
750
9.5
5.3
10.5
15.5
79.1
20.7
8.7
16.5
18.0
12.9
12.9
12.9
12.9
12.9
12.9
12.9
-2.6
0.2
-1.0
-0.6
3.4
4.1
7.3
1.5
-1.6
1.6
0.0
3.9
-5.3
-2.5
8.1
8.3
9.2
-5.5
1.3
-1.4
1.1
-2.7
3.8
-4.2
10.2
0.8
1.0
0.2
-1.3
-1.7
4.8
-6.7
10.1
11.4
1.4
-3.0
1.6
0.3
-6.2
1.6
-3.1
-5.0
-9.2
-0.6
-1.4
1.8
3.0
-4.2
4.9
-2.7
aRelative to Sgr A* in l-b coordinates (negative x is east and negative y is south of Sgr A*)
bStar not used in these calculations
cWind velocity and mass loss rate fixed (see text)
dStar position changed slightly due to finite physical resolution
The sources are assumed to be stationary over the duration of the simu-
lation. The stars without any observed He I line emission have been assigned
3
a wind velocity of 750 km sec−1 and an equal mass loss rate chosen such that
the total mass ejected by the 14 stars used here is equal to 3 × 10−3 M⊙ yr−1.
Note that although we have matched the overall mass outflow rate to the ob-
servations, we have only used 14 of the 25 stars in the sample. There are two
principal reasons for this: (1) stars further away than 10 arcsec (in projection)
from Sgr A* are outside of our volume of solution and therefore could not be
included, and (2) due to our computational resolution limits, we needed to avoid
excessively large local stellar densities.
In a complex flow, generated by many wind sources, the wind velocity and
density are not uniform, so the accretion radius is not independent of angle.
To set the length scale for the simulation, we shall therefore adopt the value
RA = .018pc (for which 1′′ = 2.3 RA) as a reasonable mean representation of
this quantity.
2.2. The Dark Cluster Potential
We wish to study the emission characteristics of a hot, magnetized plasma
"trapped" within the dark cluster's gravitational potential well. Following Haller
& Melia (1996), we will represent the gravitational potential of the dark clus-
ter with an "η-model" (Tremaine, et. al., 1994). This function represents an
isotropic mass distribution with a single parameter. We here restrict our exam-
ination to the case η = 2.5 since this provides the closest approximation to a
King model that is physically realizable (i.e., a nonnegative distribution func-
tion). We scale the mass so that 2 × 106 M⊙ are enclosed within 0.01 pc and
the total integrated mass of the dark cluster is 2.7 × 106 M⊙. Thus, writing r
in units of RA we get for the enclosed mass as a function of r,
Mη(r) = 2.7 × 106(cid:18) 13.84r
1 + 13.84r(cid:19)5/2
M⊙ .
(2)
A more recent assessment of the enclosed mass (Genzel, et. al., 1997) places
a yet more rigorous constraint on the possibility of a distributed dark matter
component. These newer observations may indeed invalidate the idea that any
realistic stable distribution of objects can account for the observed gravitational
potential.
2.3. The Hydrodynamics Code
We use a modified version of the numerical finite difference algorithm ZEUS,
a general purpose code for MHD fluids developed at NCSA (Stone & Norman,
1992; Norman, 1994). The code was run on the massively parallel Cray T3E at
NASA's Goddard Space Flight Center under the High Performance Computing
Challenge program. Zone sizes are geometrically scaled by a factor of 1.02 so that
the central zones are ∼ 20 times smaller than the outermost zones, mimicking the
"multiply nested grids" arrangement used by other researchers (e.g., Ruffert &
Melia, 1994). This allows for maximal resolution of the central region (within the
computer memory limits available) while sufficiently resolving the wind sources
and minimizing zone-to-zone boundary effects. The total volume is (40RA)3 or
∼ (0.7pc)3 with the center of the spherically symmetric dark cluster distribution
being located at the origin.
4
The density of the gas initially filling the volume of solution is set to a
small value and the velocity is set to zero, while the internal energy density is
chosen such that the initial gas temperature is ∼ 102 K. Free outflow conditions
are imposed on the outermost zones and each time step is determined by the
Courant condition with a Courant number of 0.5. The 14 stellar wind sources
are modeled by forcing the velocity in 14 subregions of 125 zones each to be
constant with time while the densities in these subvolumes are set so that the
total mass flow into the volume of solution from each source is given by Table 1.
Also, the magnetic field of the winds is assumed to always be at equipartition
with the thermal energy density. The angular momentum and mass accretion
rates are calculated by summing the relevant quantity in zones located within
0.1RA of the origin.
We assume the magnetic field is perfectly tangled and thus ignore the ef-
fects of the magnetic field on the large scale kinematics. We take the medium
to be an adiabatic polytropic gas, with γ = 5/3. Building on previous work
(Melia, 1994; Coker & Melia, 1997), we have included a first order approxi-
mation to magnetic dissipative heating as well as an accurate expression for the
cooling due to magnetic bremsstrahlung, thermal bremsstrahlung, line emission,
radiative recombination, and 2 photon continuum emission for a gas with cos-
mic abundance. For magnetic heating, we assume that the magnetic field never
rises above equipartition. If compression and flux conservation would otherwise
dictate a magnetic field larger than the equipartition value, the field lines are
assumed to reconnect rapidly, converting the magnetic field energy into thermal
energy, thereby re-establishing equipartition conditions. The cooling function
includes a multiple-Gaussian fit to the relevant cooling emissivities provided by
N. Gehrels (see Gehrels & Williams, 1993, and references cited therein), though
with the thermal bremsstrahlung portion supplanted with more accurate ex-
pressions that are valid over a broader range of physical conditions and with
the inclusion of magnetic bremsstrahlung. For details on the cooling expressions
used in the hydrodynamics code as well as the spectrum calculations below, see
Melia & Coker (1999). Note that cooling due to Comptonization and any pair
production have not yet been included since they are not thought to be signifi-
cant in the vicinity of Sgr A*. Also, it is assumed that the optical depth is small
throughout the volume of solution.
2.4. Results of the 3D Simulation
The integrated emissivity along the line-of-sight, at ∼ 1450 years after the start
of the simulation, is shown in Figure 1. The grey scale is logarithmic with
solid white corresponding to a frequency-integrated intensity of ∼ 1.1 × 105 erg
cm−2 s−1 steradian−1, and black corresponding to ∼ 1 × 10−2 erg cm−2 s−1
steradian−1. Sgr A* is located at the center of the image. Of particular interest
in this image is the appearance of streaks of high-velocity, high-density gas
("streamers") that are very reminiscent of features, such as the so-called "Bullet"
seen near the Galactic center (Yusef-Zadeh, et al. 1996). In our simulation, these
structures are produced predominantly within the wind-wind collision regions,
and in the future, we shall consider in greater detail the possibility that the
observed high-velocity gas components near Sgr A* are produced in this fashion.
5
Figure 1.
The line-of-sight integrated emissivity along the z-axis.
Note the dominant role played by IRS 13E1 to the lower right of the compact
radio source.
After reaching equilibrium several sound crossing times (∼ 1000 years) after
the start of the simulation, the enclosed mass and energy begin to fluctuate
aperiodically on time scales of less than a few decades and with an amplitude
of up to 50%, reflecting the turbulent cell nature of the flow in and out of the
central region. Typically 2.7 × 10−3 M⊙ of gas is trapped within the cluster at
any given time. Also, although the gas is generally supersonic, with most of the
energy in kinetic form, the thermal energy can be boosted rather suddenly when
the enclosed magnetic field energy is dissipated. This occurs when strong shocks
pass through the central region; the shocks compress the field sufficiently to the
point where it reaches, or even surpasses, equipartition and dissipation ensues.
6
3. The Spectrum of Sgr A*
3.1. The Spectrum due to a Dark Cluster
In order to calculate the observed continuum spectrum, we assume that the
observer is positioned along the negative z-axis at infinity and we sum the emis-
sion from all zones that are located at a projected distance, Rxy, of less than
0.1 RA. Since the size of Sgr A* (at λ7 mm) is ≈ 1012−13 cm (Bower & Backer,
1998) and the smallest cell size in our simulation is 7 × 1014 cm, in order to
minimize the inaccuracy due to numerical fluctuations, we have calculated the
spectrum from a central region roughly 10 times this size (0.1RA), to include at
least 100 zones. Thus clearly our predicted spectrum constitutes an upper limit
to the actual emission expected from Sgr A*. At the temperature and density
that we encounter here, the dominant components of the continuum emissivity
are electron-ion and electron-electron bremsstrahlung; we ignore line emission
in these spectral calculations.
As discussed in Melia (1998), the density in the central region reaches a peak
value of roughly 108 cm−3 and the temperature is never greater than about 108
K. Thus, since electrons begin to emit significant synchrotron radiation only
above a few times 109 K, the gas can only emit cyclotron radiation. However,
the cyclotron emissivity is insignificant compared to bremsstrahlung so that the
final spectrum, as shown in Figure 2, is a bremsstrahlung spectrum that,in the
radio, falls more than 4 orders of magnitude short of the observed luminosity
of Sgr A* (compare with Figure 6b in Melia, 1998). Figure 2 shows spectra for
3 points in time during the 3D hydrodynamical simulation. The solid, dotted,
and dashed curves are at approximately 1200, 1300, and 1400 years, respectively,
after the start of the simulation. The flattening of the density, temperature, and
magnetic field profiles is a direct consequence of the shallowness of the dark
cluster potential compared to the steep potential gradient encountered by gas
falling into a black hole. The high-energy shoulder of the emission shown in
Figure 3 is characteristic of the highest temperature Tmax (∼ 108 K) attained
by the gas and is consistent with the brightness temperature limit associated
with Sgr A*. Similarly, the X-ray and γ-ray emission is significantly below the
observed upper limits (Predehl & Truemper, 1994; Goldwurm, et. al., 1994).
3.2. The Spectrum due to a Black Hole
Although we have not yet run a 3D hydrodynamical simulation using the grav-
itational potential due to a point mass, we have undertaken semi-analytical
calculations that improve upon earlier work (Melia, 1994). Although details will
be presented elsewhere (Coker & Melia, 1999), we here present some preliminary
results. We now explicitly solve for the velocity of the spherical flow using the
relativistic Euler equation and integrate over µ, the cosine of the angle between
the line-of-sight and the flow. We also include effects due to inverse Compton
scattering and the index of refraction.
In recent work (Lo, et. al., 1998) it has been suggested that, at mm wave-
lengths at least, the intrinsic size of Sgr A* varies as να with −1.9 < α < −0.7.
It is difficult to compare the absolute observed size, which is usually the FWHM
of a best-fit Gaussian (e.g. Krichbaum, et. al., 1998), to a theoretical size,
which is usually given as the surface of last scatterring where τ (ν) ≡ 1; the
7
Figure 2.
Plot of the predicted luminosity density versus frequency.
observational size will tend to be larger than the theoretical size with these def-
initions. However, Figure 3 shows a plot of the predicited intrinsic size of Sgr
A* along with present observational values (see Lo, et. al. 1998 and references
cited therein for details). The slope of the line is ∼ −0.6, somewhat less steep
than the observed value. This is not surprising since this particular model does
not produce enough low frequency magnetic bremsstrahlung at large radii. The
slope should steepen when a more optimal fit to the spectrum is found. Note
that the lower size bound on the plot corresponds to the size of the black hole
itself (1Rs ∼ 6µas).
4. Summary
It does not appear, based on these calculations, that the gravitational poten-
tial of a distributed dark mass, be it due to a compact cluster of stellar rem-
nants or an even more exotic collection of objects (e.g. Tsiklauri & Viollier,
8
Figure 3.
& Melia, 1999).
Plot of the predicted size of Sgr A* versus frequency (Coker
1998), can compress the gas from stellar winds to the point where the tem-
perature, density, and magnetic field can produce an observationally significant
cyclotron/synchrotron emissivity at GHz frequencies. Although the spatial res-
olution of the simulation near the origin can be improved over that used here,
it is unlikely that it will alter the results; the lack of compression is due to the
inherently flat potential of any distributed dark cluster. Thus, aside from the
issue of whether any dark cluster can account for the observed GC potential and
whether any such cluster is stable over a significant fraction of the age of the
Galaxy, our 3D hydrodynamical simulation has shown that the radio emissivity
from the gas trapped within any such cluster cannot reprodude the spectrum of
Sgr A*. We are left to conclude that either Sgr A* is unrelated to the accreting
or trapped GC gas or that Sgr A* is the signature of an accreting, massive black
hole. However, models of the structure and mechanism of radio emission from
such a black hole are now faced with futher contraints due to recent observations
of the intrinsic source structure of Sgr A*. Coupled with the observed spectral
energy distribution of Sgr A*, we may be close to differentiating between the
various accretion models.
Acknowledgments. This work was partially supported by NASA grant
NGT-51637 and has made use of NASA's Astrophysics Data System Abstract
Service.
9
References
Backer, D.C. 1994, in The nuclei of normal galaxies, ed. R. Genzel & A.I. Harris
(Dordrecht: Kluwer), 403.
Bondi, H, & Hoyle, F. 1944, MNRAS, 104, 273.
Bower, G.C. & Backer, D.C. 1998, ApJ, 496, 97L.
Coker, R., & Melia, F. 1996, in ASP Conf. Vol. 102, The Galactic Center: 4th
ESO/CTIO Workshop, ed. R. Gredel (San Francisco: ASP), 403.
Coker, R. & Melia, F., 1997, ApJ, 488, L149.
Coker, R. & Melia, F., 1999, in preperation.
Eckart, A. & Genzel, R., 1998, these proceedings.
Gehrels, N. & Williams, E. D. 1993, ApJ, 418, L25.
Genzel, R., Thatte, N., Krabbe, A., Kroker, H., & Tacconi-Garman, L.E., 1996,
ApJ, 472, 153.
Genzel, R., Eckart, A., Ott, T. & Eisenhauer, F. 1997, MNRAS, 291, 219.
Goldwurm, A., et. al., 1994, Nature, 371, 589.
Haller, J.M., & Melia, F. 1996, ApJ, 464, 774.
Haller, J.M., Rieke, M.J., Rieke, G.H., Tamblyn, P., Close, L., & Melia, F. 1996,
ApJ, 456, 194.
Herbst, T.M., Beckwith, S.V.W., Forrest, W.J., Pipher, J.L. 1993, AJ, 105, 956.
Lee, H.M. 1995, MNRAS, 272, 605.
Lo, K.Y., Shen, Z.-Q., Zhao, J.-H., & Ho, P. 1998, ApJ, 508, L61.
Lutz, D., Krabbe, A. & Genzel, R. 1993, ApJ, 418, 244.
Maoz, E. 1998, ApJ, 494, L181.
Melia, F. 1994, ApJ, 426, 577.
Melia, F. 1998, these proceedings.
Melia, F. & Coker, R. 1999, ApJ, in press.
Mezger, P.G., Duschl, W.J. & Zylka, R. 1996, A&A Rev., 7, 289.
Morris, M. & Serabyn, E. 1996, ARA&A, 34, 645.
Najarro, F., Krabbe, A., Genzel, R., Lutz, D., Kudritzki, R., & Hillier, D. 1997,
A&A, 325, 700.
Narayan, R., Yi, I. & Mahadevan, R. 1995, Nature, 374, 623.
Norman, M. 1994, BAAS, 184, 50.01.
Predehl, P. & Truemper, J. 1994, a, 290, 29.
Roberts, D.A. & Goss, W.M. 1993, ApJS, 86, 133.
Ruffert, M., & Melia, F. 1994, A.A.Letters, 288, L29.
Stone, J.M., & Norman, M.L. 1992, ApJS, 80, 753.
Tremaine, S., Richstone, D. O., Byun, Y., Dressler, A., Faber, S. M., Grillmair,
C., Kormendy, J., & Lauer, T. R. 1994, AJ, 107, 634.
Yusef-Zadeh, F., Roberts, D.A., Goss, W.M., Frail, D. & Green, A. 1996, ApJ,
466, L25.
10
|
0704.2204 | 1 | 0704 | 2007-04-17T17:55:29 | The Differential Rotation of Kappa1 Ceti as Observed by MOST | [
"astro-ph"
] | We first reported evidence for differential rotation of Kappa1 Ceti in Paper I. In this paper we demonstrate that the differential rotation pattern closely matches that for the Sun. This result is based on additional MOST (Microvariability & Oscillations of STars) observations in 2004 and 2005, to complement the 2003 observations discussed in Paper I. Using StarSpotz, a program developed specifically to analyze MOST photometry, we have solved for k, the differential rotation coefficient, and P_{EQ}, the equatorial rotation period using the light curves from all three years. The spots range in latitude from 10 to 75 degrees and k = 0.090^{+0.006}_{-0.005} -- less than the solar value but consistent with the younger age of the star. k is also well constrained by the independent spectroscopic estimate of vsini. We demonstrate independently that the pattern of differential rotation with latitude in fact conforms to solar.
Details are given of the parallel tempering formalism used in finding the most robust solution which gives P_{EQ} = 8.77^{+0.03}_{-0.04} days -- smaller than that usually adopted, implying an age < 750 My. Our values of P_{EQ} and k can explain the range of rotation periods determined by others by spots or activity at a variety of latitudes. Historically, Ca II activity seems to occur consistently between latitudes 50 and 60 degrees which might indicate a permanent magnetic feature. Knowledge of k and P_{EQ} are key to understanding the dynamo mechanism and rotation structure in the convective zone as well assessing age for solar-type stars. We recently published values of k and P_{EQ} for epsilon Eri based on MOST photometry and expect to analyze MOST light curves for several more spotted, solar-type stars. | astro-ph | astro-ph |
Draft October 28, 2018
Preprint typeset using LATEX style emulateapj v. 08/22/09
THE DIFFERENTIAL ROTATION OF κ1 CETI AS OBSERVED BY MOST1 -- II
Gordon A.H. Walker2, Bryce Croll3, Rainer Kuschnig3, Andrew Walker4, Slavek M. Rucinski5, Jaymie M.
Matthews3, David B. Guenther6, Anthony F.J. Moffat7, Dimitar Sasselov8, Werner W. Weiss9
Draft October 28, 2018
ABSTRACT
We first reported evidence for differential rotation of κ1 Ceti in Paper I. In this paper we demon-
strate that the differential rotation pattern closely matches that for the Sun. This result is based on
additional MOST (Microvariability & Oscillations of STars) observations in 2004 and 2005, to comple-
ment the 2003 observations discussed in Paper I. Using StarSpotz, a program developed specifically
to analyze MOST photometry, we have solved for k, the differential rotation coefficient, and PEQ, the
equatorial rotation period using the light curves from all three years. The spots range in latitude from
10◦ to 75◦ and k = 0.090+0.006
−0.005 -- less than the solar value but consistent with the younger age of the
star. k is also well constrained by the independent spectroscopic estimate of vsini. We demonstrate
independently that the pattern of differential rotation with latitude in fact conforms to solar.
Details are given of the parallel tempering formalism used in finding the most robust solution which
gives PEQ = 8.77+0.03
−0.04 days -- smaller than that usually adopted, implying an age < 750 My. Our
values of PEQ and k can explain the range of rotation periods determined by others by spots or
activity at a variety of latitudes. Historically, Ca II activity seems to occur consistently between
latitudes 50◦ and 60◦ which might indicate a permanent magnetic feature. Knowledge of k and PEQ
are key to understanding the dynamo mechanism and rotation structure in the convective zone as
well assessing age for solar-type stars. We recently published values of k and PEQ for ǫ Eri based
on MOST photometry and expect to analyse MOST light curves for several more spotted, solar-type
stars.
Subject headings:
individual: Kappa1 Ceti, HD 20630 -- stars:
stars: activity -- stars:
stars: starspots -- stars: rotation -- stars: oscillations -- stars: exoplanets
late-type --
1. INTRODUCTION
Differential rotation and convection provide the engine
for the solar dynamo (eg. Ossendrijver (2003)). The
mechanism suggested by Lebedinskii (1941) for differ-
ential rotation seems to be the most widely accepted.
Angular momentum is continually redistributed in the
deep convective envelope which occupies some 30% of
the outer solar radius resulting in nonuniform rotation
because the convective turbulence is affected by the Cori-
olis force (Kitchatinov 2005). There is good evidence
1 Based on data from the MOST satellite, a Canadian Space
Agency mission, jointly operated by Dynacon Inc., the University
of Toronto Institute of Aerospace Studies and the University of
British Columbia with the assistance of the University of Vienna.
2 1234 Hewlett Place, Victoria, BC V8S 4P7, Canada; gor-
[email protected]
3 Dept.
Physics & Astronomy, UBC, 6224 Agricultural
Road, Vancouver, BC V6T 1Z1, Canada; [email protected],
[email protected], [email protected]
4 Sumus Technology Ltd.; [email protected]
5 Dept. Astronomy & Astrophysics, David Dunlap Obs., Univ.
Toronto P.O. Box 360, Richmond Hill, ON L4C 4Y6, Canada;
[email protected]
6 Department of Astronomy and Physics, St. Mary's University
Halifax, NS B3H 3C3, Canada; [email protected]
7 D´ept de physique, Univ de Montr´eal C.P. 6128, Succ. Centre-
Ville, Montr´eal, QC H3C 3J7, and Obs du mont M´egantic, Canada;
[email protected]
8 Harvard-Smithsonian Center for Astrophysics, 60 Garden
Street, Cambridge, MA 02138, USA; [email protected]
9 Institut fur Astronomie, Universitat Wien Turkenschanzstrasse
17, A -- 1180 Wien, Austria; [email protected]
that a younger Sun would have rotated more rapidly and
gradually lost angular momentum through magnetic cou-
pling to the solar wind and there is currently a consid-
erable theoretical effort to develope quantitative models
of differential rotation in solar-type stars (see, for exam-
ple, reviews by Kitchatinov (2005) and Thompson et al.
(2003)).
For stars other than the Sun, the simultaneous detec-
tion of two or more spots at different latitudes allows a
direct measurement of differential rotation. The MOST
photometric satellite (Walker et al. 2003a) offers contin-
uous photometry of target stars with an unprecedented
precision for weeks at a time making it it possible to
track spots accurately.
We (Rucinski et al. 2004) have already reported the
detection of a pair of spots in the 2003 MOST light curve
of κ1 Ceti (HD 20630, HIP 15457, HR 996; V = 4.83,
B − V = 0.68, G5V). The photometry was part of the
satellite commissioning and prior to a considerable im-
provement pointing in early 2004. The larger spot had a
well-defined 8.9 d rotation period while the smaller one
had a period ≥9.3 d. Only the latitude of the larger spot
could be determined with any confidence. Spectroscopic
observations by Shkolnik et al. (2003), mostly from 2002,
of the Ca II K-line emission reversals were also synchro-
nized to a period of about 9.3 d. Rucinski et al. (2004)
also determined a stellar radii of R∗ = 0.95 ± 0.10 RJ.
To better decipher the spot activity and rotation of
κ1 Ceti and to model spot distributions on other MOST
2
Walker et al.
targets, one of us, BC, developed the program Starspotz
(Croll et al. (2006b); Croll (2006a)). It was applied first
to a 2005 MOST light curve covering three rotations of ǫ
Eri where two spots were detected at different latitudes
rotating with different periods. From this, we derived a
differential rotation coefficient for ǫ Eri (see section 4),
k = 0.11+.03
−.02, in agreement with models by Brown et al.
(2004) for young Sun-like stars rotating roughly twice as
fast as the Sun.
In this paper, we apply Starspotz to the three separate
MOST light curves of κ1 Ceti -- from 2003, 2004 and 2005.
We specifically solve for the astrophysically important
elements of inclination, differential rotation, equatorial
period and rotation speed and find that the differential
rotation profile is identical to solar.We use parallel tem-
pering methods to fully explore the realistic parameter
space and identify the global χ2 minimum to the data-set
that signifies a best-fitting unique starspot configuration.
Markov Chain Monte Carlo (MCMC) methods are then
used to explore in detail the parameter space around this
global minimum, to define best-fitting values and their
uncertainties and to explore possible correlations among
the fitted and derived parameters.
2. MOST PHOTOMETRY AND LIGHT CURVES
The MOST satellite, launched on June 2003, is fully de-
scribed by Walker et al. (2003a). A 15/17.3 cm Rumak-
Maksutov telescope feeds two CCDs, one for tracking and
the other for science, through a single, custom, broad-
band filter (350 -- 700 nm). Starlight from primary sci-
ence targets (V ≤ 6) is projected onto the science CCD
as a fixed (Fabry) image of the telescope entrance pupil
covering some 1500 pixels. The experiment was designed
to detect photometric variations with periods of minutes
at micro-magnitude precision and does not use compar-
ison stars or flat-fielding for calibration. There is no di-
rect connection to any photometric system. Tracking jit-
ter was dramatically reduced early in 2004 to ∼1 arcsec
which, subsequently, led to significantly higher photo-
metric precision.
The observations reported here were reduced by RK.
Outlying data points generated by poor tracking or cos-
mic ray hits were removed. At certain orbital phases,
MOST suffers from parasitic light, mostly Earthshine,
the amount and phase depending on the stellar coor-
dinates, spacecraft roll and season of the year. Data
are recorded for seven Fabry images adjacent to the tar-
get to track the background. The background signals
were combined in a mean and subtracted from the tar-
get photometry. This procedure also corrected for bias,
dark and background signals and their variations. Re-
ductions basically followed the scheme already outlined
by Rucinski et al. (2004).
A log of the MOST observations of κ1 Ceti for each
of 2003, 2004 and 2005 is given in Table 1. For the
light curve analysis in this paper, data were binned at
the MOST orbital period of 101.413 min. Figure 1 dis-
plays full light curves for each of the three years with
time given as JD-2451545 (heliocentric). The solid line
in the lower panel displays the difference between the
two most divergent background readings from the seven
background Fabry images. There is no obvious corre-
lation between structure in the differential background
signal and the light curve. The complete light curve
can be downloaded from the MOST Public Archive at
www.astro.ubc.ca/MOST.
The formal rms point-to-point precison, σ, is given in
Table 1. The shapes of the light curves ultimately used
for the StarSpotz analysis depend to a degree on just how
the background is removed. Despite having the summed
signal values from 7 adjacent Fabry images we cannot
interpolate a 'true' background but subtraction of the
mean background is a consistent option. In each of the
three data sets used here the 7 background values are
closely identical when averaged over an orbit but the dif-
ferences between the extreme backgrounds can be signif-
icant when the ambient background is high (eg. pointing
close to the Moon) which is why that difference is shown
in the lower panel for each light curve.
We have assumed that the true errors in the individ-
ual orbital means making up each light curve are pro-
portional to the difference between the maximum and
minimum background values. These errors have been
arbitrarily scaled by one-fourth for the formal StarSpotz
analysis and these 'errors' are shown as bars in Figures 4,
5, and 6. The individual bars correspond approximately
to the standard deviation of the background.
3. STARSPOTZ
StarSpotz (Croll et al. 2006b) is a program based on
SPOTMODEL (Rib´arik et al. 2003; Rib´arik 2002), that
is designed to provide a graphical user interface for pho-
tometric spot modeling. The functionality included in
StarSpotz is designed to take advantage of the nearly
continuous photometry returned by MOST to quantify
the non-uniqueness problem often associated with pho-
tometric spot modeling. Basically, StarSpotz uses the
analytic models of Budding (1977) or Dorren (1987) to
model the drop in intensity caused by NSP OT circular,
non-overlapping spots. The model proposed in Budding
(1977) is used in this application.
Fitting can be performed by a variety of methods in-
cluding a Marquardt-Levenberg non-linear least squares
algorithm (Marquardt 1963; Levenberg 1944; Press et al.
1992), or Markov Chain Monte Carlo (MCMC) function-
ality (Gregory 2005b; Gamerman et al. 1999; Gilks et al.
1996). The parameters of interest are the stellar incli-
nation angle, i, the unspotted intensity of the star, U ,
and the period, epoch, latitude and angular size of the
ath spot: pa, Ea, βa, and γa. Croll et al. (2006b) have
demonstrated how the Marquardt-Levenberg non-linear
least squares algorithm could be used to fit the cycle to
cycle variation in ǫ Eri. Croll (2006a) investigated the
same observations of ǫ Eri demonstrating that Markov
Chain Monte Carlo functionality is very helpful in defin-
ing correlations and possible degeneracies between the
fitted parameters.
The search for additional local minima in χ2 space indi-
cating other plausible spot configurations that produce a
reasonable fit to the light-intensity curve is an important
part of the modeling because it allows one to quantify
the uniqueness of a solution. We argue that this task
Differential rotation of κ1 Ceti
3
TABLE 1
MOST Observations of κ1 Ceti
year
start
dates
HJD -- 2451545
2003 5 Nov
2004 15 Oct
2005 14 Oct
1403.461 to 1433.901
1748.578 to 1768.989
2112.721 to 2125.065
total
days
30.44
20.41
12.34
duty cycle
%
96
99
83
exp
sec
cycle
sec mmag
σa
40
30
30
50
35
35
0.5
0.30
0.19
a for data binned at the MOST orbital period of 101.413 min.
of completely searching parameter space for other phys-
ically plausible local minima is better suited to a pro-
cess called parallel tempering (Gregory 2005a; Gregory
2005b), discussed here in §4.2, rather than the Unique-
ness test methods in Croll et al. (2006b).
The present and subsequent releases of the freely avail-
able program StarSpotz, including the full source-code,
executable, and documentation are available at:
www.astro.ubc.ca/MOST/StarSpotz.html
Release 3.0 includes the functionality discussed which
allows one to fit multiple epochs of data as well as the
parallel tempering functionality.
4. THE SPOTS ON κ1 CETI
As the MOST observations of κ1 Ceti come from three
well-seperated epochs with independent starspot groups
from epoch to epoch, the fitting process was consider-
ably more complicated than those detailed in Croll et al.
(2006b) and Croll (2006a). Rather than fitting for the pe-
riod, p, of the individual spots in each epoch we derive a
common equatorial period, PEQ, and the differential ro-
tation coefficient, k, among the three epochs where these
are given by:
Pβ = PEQ/(1 − k sin2β)
(1)
where Pβ is the period at latitude β. There are five fewer
parameters to be fitted because the periods of the in-
dividual spots are no longer independent variables, but
completely defined by k, PEQ, and β. Analyses of so-
lar differential rotation usually include terms in sin4β
(Kitchatinov 2005) but we did not expect to detect spots
at such high latitudes where this term would be impor-
tant.
Common limb-darkening coefficients, u = 0.684, and
flux ratios between the spot and unspotted photosphere,
κω = 0.22 were assumed. The former was based on
the compilations of D´ıaz-Cordov´es et al. (1995), while
the latter comes from the canonical value for sunspots
(Tspot ∼ 4000 K, Tphot ∼ 5800 K). The 27 parameters to
be fitted are summarized in Table 2.
Initial
investigations of the light curves using the
StarSpotz standard spot model indicated that different
numbers of spots were required for the various epochs.
The 2003 and 2005 MOST data-sets could be well ex-
plained by two starspots rotating with different periods,
at different latitudes. The behaviour observed in the
2004 MOST light curve, on the other hand, required a
third spot. Thus the following analysis explicitly assumes
two, three and two starspots for the 2003-2005 MOST
epochs, respectively.
In this present application we use the same paradigm
that was applied to κ1 Ceti by Rucinski et al. (2004), and
ǫ Eri by Croll (2006a) and Croll et al. (2006b) where the
spot parameters (E, β, γ) were assumed to be constant
throughout the observations, and that the cycle-to-cycle
variability observed is due solely to spots rotating dif-
ferentially. Other interpretations are certainly plausible
including circular spots with time-varying spot sizes and
latitudes such as those applied to IM Peg in Rib´arik et al.
(2003). The Maximum Entropy & Tikhonov reconstruc-
tion technique of Lanza et al. (1998), which assumes
pixellated spots of various contrasts, is another example.
We feel that with the limitations of photometric spot
modeling, our assumption of unchanging spot parame-
ters is the most suitable for the timescale of the MOST
light curves.
It provides an opportunity to determine
a unique solution, as well as recovering astrophysically
important parameters.
We recognise, given the obvious changes in the κ1 Ceti
light curve over a year, that the assumption of spot
parameters constancy over a few weeks is a simplifica-
tion but we do not expect significant spot migration on
timescales of less than a month.
4.1. StarSpotz Markov Chain Monte Carlo (MCMC)
functionality
The Markov Chain Monte Carlo (MCMC) functional-
ity used in StarSpotz is fully described in Croll (2006a).
The goal is to take an n-step intelligent random walk
around the parameter space of interest while recording
the point in parameter space for each of the K fitted
parameters for each step. This samples the posterior
parameter distribution, which can be thought of in pho-
tometric spot modeling as simply the range in each of the
K fitted parameters that provide a reasonable fit to the
light curve. The advantage of a MCMC is that it does not
simply follow the path of steepest descent - resulting in
the possibility that it becomes stuck in a local minimum
- but allows the MCMC to fully explore the immediate
parameter space. The K parameters can thus be defined
by a K dimensional vector y.
T stands for the "virtual temperature" and determines
the probability that the MCMC chain will accept large
deviations to higher regions in χ2 space (Sajina et al.
2006). The virtual temperature should properly be unity
(T = 1) to sample the posterior parameter distribution
when reduced χ2 is near unity. However, in photomet-
ric spot modeling, nature is often more complicated than
one's model can take into account, resulting in anoma-
lously high reduced χ2 values. Setting T to be the mini-
mum reduced χ2 value observed produces the same effect
4
Walker et al.
as scaling reduced χ2 to one, and thus effectively sam-
ples the posterior parameter distribution. In this present
application Tmin will refer to this value of T equal to the
minimum reduced χ2 observed.
A suitable burn-in period as described in Croll (2006a)
is often used, or a χ2 cut where points with χ2
values above this user-specified minimum are excised
from the analysis. A user-specified thinning factor, f ,
(Croll 2006a) is also used. To ensure proper conver-
gence and mixing we employ the tests as proposed by
Gelman and Rubin (1992) and explicitly defined for our
purposes in Croll (2006a). We ensure the vector R
falls as close to 1.0 as possible for each of the K fitted
parameters. The marginalized likelihood (Sajina et al.
2006; Lewis and Bridle 2002) as described in §3.1 of Croll
(2006a) is used to define our 68% credible regions, while
the mean likelihood (Sajina et al. 2006) is used for com-
parison. The mean likelihood is produced using the χ2
values in a particular bin of a histogram of parameter
values, while the marginalized likelihood directly reflects
the fraction of times that the MCMC chains are in a
particular bin. Peaks in the marginalized likelihood that
are inconsistent with peaks in the mean likelihood can be
indicative of the chains having yet to converge, or larger
multidimensional parameter space rather than a better
fit (Lewis and Bridle 2002). In our analysis we quote the
minimum to maximum range of the 68% credible regions.
4.2. Parallel Tempering
Gregory (2005a) applied parallel tempering to a radial
velocity search for extrasolar planets, as the parameter
space he investigated had widely seperated regions in pa-
rameter space that provided a reasonable fit to the data.
In our application as we fit for a large number of param-
eters, K=27, we use parallel tempering to explore our
multi-dimensional parameter space and search for other
local minima in χ2 space indicative of other plausible so-
lutions, and thus starspot configurations, that produce a
reasonable fit to the light-intensity curve.
In parallel tempering, MP multiple MCMC chains are
run simultaneously each with different values of the vir-
tual temperature, Tm, where m is an index that runs
from 1 to MP . Chains with higher values of T are
easily able to jump to radically different areas in pa-
rameter space ensuring that the chain does not become
stuck in a local minimum. Chains with lower values of
T can refine themselves and descend into local minima
and possibly the global minimum in χ2 space (Gregory
2005a). The chain with T = 1.0×Tmin is the target dis-
tribution and is referred to as the "cold sampler." The
parameters of the "cold sampler" are used for analysis
to return the posterior parameter distribution. A set
mean fraction of the time, l, a proposal is made for
the parameters y of adjacent chains to be exchanged.
If this proposal is accepted than adjacent chains are cho-
sen at random and the K parameters y for each chain
are exchanged.
In this way, this method allows radi-
cally different areas in parameter space to be investi-
gated in the high temperature chains, before reaching
the "cold sampler" where a local minimum can be sought
efficiently. We run 11 parallel chains with virtual temper-
atures, Tm, of: T = {100×Tmin, 50.0×Tmin, 33.0×Tmin
,
3.00×Tmin,
2.00×Tmin, 1.50×Tmin 1.25×Tmin, 1.00×Tmin }.
20.0×Tmin,
10.0×Tmin,
5.0×Tmin,
In a typical parallel tempering scheme the number of
iterations between swap proposals is set to some con-
stant integer, B (e.g. B ≈ 10). With the large number
of parameters being fitted across the three epochs of data
(K=27) in the current application, it was found that this
typical scheme was not able to efficiently reach the low-χ2
valleys in the multi-dimensional parameter space indica-
tive of a reasonable fit to the light-intensity curve. Thus
the number of iterations between swap proposals, Bm,
was set to a variable number between the chains. Bm
was set to a larger number for the lower temperature
chains to allow these chains a greater number of itera-
tions to refine themselves and explore the low-χ2 areas.
This was found to be a suitable compromise as the higher
virtual temperature chains required fewer iterations to
reach radically different areas of parameter space, while
the lower virtual temperature chains required nearly an
order of magnitude more iterations to reach lower ar-
eas in χ2 space indicative of reasonable fits to the light-
curve. The number of chain points per chain was set
to B = {Bmin, Bmin, Bmin, Bmin, Bmin, Bmin, Bmin,
2.0×Bmin, 2.0×Bmin, 3.0×Bmin, 4.0×Bmin}. Bmin was
set to twenty (Bmin=20), within an order of magnitude
of the value proposed by Gregory (2005a); extensive tests
have proven this choice to be reasonable. l was set as 0.80
to allow for exchanges between adjacent MCMC chains
on average 80% percent of the time. That means a swap
is approved if ul < l, where ul
is a random number
(ul ǫ[0, 1]). Significant experimentation was needed to
choose and refine suitable choices of Tm, Bm, Bmin, and
l.
In the present application we start 8×11 of these par-
allel tempering chains starting from random points in
our acceptable parameter space as given in the Paral-
lel Tempering initial allowed range column of Table 2.
MCMC fitting was then implemented for n=4000 steps
to ensure that the starting point for the parallel tem-
pering chains provided a mediocre fit to the light curve.
A mediocre fit was desired to ensure that the starting
point for the parallel tempering chains provided a rea-
sonable fit to the light curve so as to reduce the required
burn-in period, while not starting from within a signifi-
cant local minimum so as to allow the parallel tempering
chains to efficiently explore the parameter space. Each of
these 8×11 MCMC parallel tempering chains were then
run for nx=13000 exchanges, resulting in nx×20×4 =
1.04×106 total steps for each of the 8 "cold samplers."
A burn-in period was not used in this application for our
analysis. Rather a χ2 cut was used where all points of
the parallel tempering "cold samplers" with reduced χ2
greater than some particular minimum, χ2
CU T , were ex-
cised from the analysis. This serves a similar function to
a burn-in period for the initial points, while allowing the
analysis to focus on the chain points that provide a rea-
sonable fit to the light curve. Given the large number of
parameters we fit for, K=27, this is useful as the paral-
lel tempering chains spend an inordinate amount of time
exploring parameter spaces that do not provide a good
Differential rotation of κ1 Ceti
5
fit to the data, and would otherwise complicate the anal-
ysis. A thinning factor, f , of 20 were used in all chains
for the parallel tempering analysis. The 8×11 parallel
tempering chains required a total computational time of
7.0 CPU days on a Pentium processor with a clock speed
of 3.2 Ghz with 1.0 GB of memory. This value of nx was
chosen to ensure that the vector R, in this case RP arallel
(Gelman and Rubin 1992; Croll 2006a), fell as close to
1.0 as possible for each of the K = 27 parameters. R
< ≈1.3 signifies that the chain points have roughly con-
verged and thus we are close to an equilibrium distribu-
tion. To determine RP arallel we do not use the above χ2
cut but instead use a burn-in period of nburn = 30000
steps. This is because we are interested in ensuring the
parallel tempering chains have converged while exploring
our parameter space as a whole, rather than the narrow
region of parameter space signifying the global χ2 mini-
mum. Analysis of these 8 "cold samplers" is given below
in §4.2.1.
4.2.1. κ1 Ceti Parallel Tempering results
The Parallel Tempering results as applied to MOST's
2003-2005 observations of κ1 Ceti indicated that there is
a single unique solution to the data-set that provides the
best fit. Following the nx=13000 exchanges as described
above the vector RP arallel was investigated for each of
the K=27 parameters. It was found that RP arallel fell
below 1.6 for all parameters, and RP arallel fell below 1.3
for most parameters. These low values of RP arallel indi-
cate adequate convergence has been obtained. Each of
the components of the vector RP arallel are given in Table
2. These low values for all the K parameters of RP arallel
indicate that our parallel tempering chains have approxi-
mately converged to an equilibrium, and adequately sam-
pled the realistic parameter space.
For our analysis we set Tmin as 2.96 as it is the global
minimum observed in the below section (§4.3) and a re-
duced χ2
CU T of 3.38 as it was judged this was reasonably
close to the global χ2 minimum to focus the analysis ex-
clusively on the solutions that provided a reasonable fit
to the light curve. Our 68% credible regions are given
by the marginalized likelihood following the χ2 cut. The
immediate parameter space indicated by these results are
given in Table 3. This parameter space was thus worthy
of more detailed investigation to explore this χ2 global
minimum, as well as to define best-fit values and appro-
priate uncertainties for the fitted and derived parameters
of interest. Also, possible correlations between the fitted
parameters were investigated for.
Detailed investigation of the parallel tempering results
indicate that all other local minima produce significantly
worse fits to the data-sets. The only minor exception
indicated by the parallel tempering results is that the
second spot in the 2003 epoch can also be found at a
latitude, β2003 2, in the southern hemisphere as well as
the northern hemisphere. Detailed investigation indicate
that the northern hemisphere solution provides a signif-
icantly better fit and thus is the solution that will be
investigated. Given the fact that all other local minima
have been ruled out, and the low values of RP arallel for
all K parameters, we feel fully justified in stating that
this parameter space is the unique realistic solution to
the observed light curve given our assumptions.
4.3. κ1 Ceti MCMC application
In §4.2 we demonstrated that the solution presented
is unique for the 2003-2005 MOST epochs under the k,
PEQ, paradigm. Thus we used the MCMC methods dis-
cussed in Croll (2006a) and summarized above in §4.1 to
investigate possible correlations and derive best-fit values
and uncertainties in the fitted and derived parameters for
κ1 Ceti. We used M =4 chains starting from reasonably
different points in parameter space, although within the
local χ2 minimum that defined the parameter space of
interest of §4.2. The prior ranges are given in Table 2.
The choices in prior are identical to those used for the
Parallel Tempering section except a slightly more con-
servative range is used for the inclination angle, 30o < i
< 80o, to limit the MCMC chain to a realistic parameter
space. We run each of the M =4 chains for n=7.0×106
steps, requiring a total computational time of 8.0 CPU
days on the aforementioned processor. This large value of
n was motivated to ensure that the R vector, in this case
RM CM C , fell below 1.20 for all K=27 parameters, and
fell below 1.05 for most parameters, indicative of suitable
convergence. The values of each of the K components of
RM CM C are given in Table 2.
The results of this analysis, henceforth referred to as
Solution 1, can be seen in Figures 2 and 3, and are sum-
marized in Table 3. As one might expect, especially given
the similar results of ǫ Eri (Croll 2006a), the latitudes
of the various spots are moderately correlated with the
inclination angle of the system, i. However, the differen-
tial rotation coefficient, k, is largely uncorrelated with i.
This is in stark contrast to the results of ǫ Eri where the
correlation of spot latitude with i produced a moderate
correlation between k and i. Our derived value of k =
0.085 - 0.096 is therefore very robust as it is does not
require an independent estimate of i.
It is important to note, though, that the 2003 and 2005
MOST data-sets do not severely limit the allowed best-
fit range of the differential rotation coefficient, k. The
2003 and 2005 MOST data-sets were best-fit with spots
that were not widely seperated in latitude, and thus these
epochs contributed only marginally to the determination
of the differential rotation coefficient. The 2004 MOST
data-set most severely limited the best-fit range of the
differential rotation coefficient, as it required three spots
ranging in latitude from the mid-southern hemisphere,
to the equator, to near the north pole.
Also, although we have used a relatively liberal flat
prior on the stellar inclination angle, i, of 30o < i <
80o, the value of i returned from our fit gives vsini =
4.64 - 4.94 km s−1, entirely consistent with the spectro-
scopic value of vsini = 4.64 ± 0.11 km s−1 determined by
Rucinski et al. (2004) from the width of the Ca II K-line
reversals. This agreement is a very useful sanity check
and gives us great confidence that our fitted results and
our assumptions of circular spots are in close agreement
with the actual behaviour of the star during these three
epochs.
The minimum χ2 point observed in our MCMC chains
6
Walker et al.
κ1 Ceti MCMC fitted parameters
TABLE 2
parameter
prior allowed
range
Parallel Tempering RP arallel RM CM C
initial allowed range
i (o)
k
PEQ (d)
U2003
E2003 1 (JD-2451545)
β2003 1 (o)
γ2003 1 (o)
E2003 2 (JD-2451545)
β2003 2 (o)
γ2003 2 (o)
U2004
E2004 1 (JD-2451545)
β2004 1 (o)
γ2004 1 (o)
E2004 2 (JD-2451545)
β2004 2 (o)
γ2004 2 (o)
E2004 3 (JD-2451545)
β2004 3 (o)
γ2004 3 (o)
U2005
E2005 1 (JD-2451545)
β2005 1 (o)
γ2005 1 (o)
E2005 2 (JD-2451545)
β2005 2 (o)
γ2005 2 (o)
25 - 85, 30 - 80 a
−0.75 - 0.75
8.0 - 10.5
0.99 - 1.01
1409.0 - 1413.0
−90 - 90
0.0 - 30
1404.0 - 1408.0
−90 - 90
0.0 - 30
0.99 - 1.015
1754.7 - 1758.7
−90 - 90
0.0 - 30
1761.5 - 1765.5
−90 - 90
0.0 - 30
1767.1 - 1771.1
−90 - 90
0.0 - 30
0.99 - 1.01
2115.0 - 2119.0
−90 - 90
0.0 - 30
2118.1 - 2122.1
−90 - 90
0.0 - 30
25 - 85
−0.6 - 0.6
8.0 - 10.5
1.000
1409.0 - 1413.0
−10.0 - 60.0
4.0 - 8.0
1404.0 - 1408.0
−10.0 - 60.0
4.0 - 8.0
1.000
1754.7 - 1758.7
−10.0 - 60.0
2.5 - 6.0
1761.5 - 1765.5
−10.0 - 60.0
2.5 - 6.0
1767.1 - 1771.1
−10.0 - 60.0
2.5 - 6.0
1.000
2115.0 - 2119.0
−10.0 - 60.0
4.0 - 8.0
2118.1 - 2122.1
−10.0 - 60.0
4.0 - 8.0
1.033
1.443
1.477
1.415
1.338
1.144
1.110
1.051
1.548
1.312
1.318
1.179
1.338
1.126
1.391
1.596
1.209
1.079
1.490
1.237
1.235
1.182
1.249
1.122
1.289
1.074
1.057
1.010
1.053
1.012
1.024
1.006
1.023
1.031
1.009
1.026
1.026
1.124
1.004
1.013
1.022
1.005
1.098
1.100
1.039
1.042
1.035
1.179
1.057
1.059
1.070
1.038
1.072
1.092
a The left hand values were used for the parallel tempering application (§4.2.1), while the
values on the right were used for the MCMC application (§4.3).
is also quoted in Table 3. The fit to the light curves in
2003, 2004, 2005 and the views of the modelled spots
shown from the line of sight (LOS) of this minimum χ2
solution are given in Figures 4, 5, and 6.
Our current results are largely consistent with the anal-
ysis of the 2003 MOST data by Rucinski et al. (2004) as
can be seen in the comparison given in Table 3. The
only discrepancies of note are that our MCMC analysis
indicates a slightly smaller value of the inclination an-
gle, i = 57.8 − 63.5o, and that we find a shorter period
of the 2nd spot, p2003 2 = 9.022 − 9.094 d. The period
found by Rucinski et al. (2004) for the 2nd spot depends
on effective removal of the variations caused by the larger
spot and it was not possible to asssign a reliable latitude.
They also asumed that the spots were black making them
smaller than those in this paper.
Differential rotation of κ1 Ceti
7
4.4. κ1 Ceti 2004 Parallel Tempering and MCMC
application
Given the good agreement of the MOST 2003, 2004
and 2005 κ1 Ceti observations with the assumed solar-
type differential rotation profile in Equation 1 and sum-
marized in Figure 7, we decided to test independently
whether solar-type differential rotation is indeed present
in κ1 Ceti. The 2004 light curve was chosen because, as
noted above, it constrained the equatorial period, PEQ,
and differential rotation coefficient, k, more significantly
than the light curves from the other two years.
For this independent test we fitted explicitly for the
periods of each of the three spots: p2004 1, p2004 2, and
p2004 3. The K=13 fitted parameters are summarized in
Table 4. For these we used the same priors, and for the
parallel tempering chains we started from the same ran-
dom ranges in parameter space, as summarized in Table
2. We assumed i= 60.6o corresponding to the minimum
χ2 value found above. This assumption for the inclina-
tion angle is justified because it results in vsini = 4.77km
s−1(assuming PEQ ≈ 8.78), a value close to the spectro-
scopic value measured by Rucinski et al. (2004).
Parallel tempering as described in §4.2 was used to
find the unique global minimum. Starting from random
points in acceptable parameter space we implemented
MCMC fitting for n=3000 steps to ensure a mediocre
fit to the light curve. Parallel tempering chains were
then implemented with nx=1800 exchanges resulting in
144000 steps for each of the 8 "cold samplers". We used
a burn-in period of the first 5000 steps to determine the
R vector, in this case R2004
P arallel. All K=13 parameters
of R2004
P arallel fell below 1.4, and below 1.2 for most, as
summarized in Table 4. Analysis is performed using a
Tmin of 1.50 and a reduced χ2
CU T of 2.0. The parallel
tempering results identify a unique global minimum as
summarized in Table 4, that is nearly identical to Solu-
tion 1 as summarized above in §4.3 and Table 3.
Thus after using parallel tempering to identify the
unique global minimum, we explored this parameter
space with our MCMC techniques as described above
in §4.1 and §4.3. We use n=3.7×106 steps, and then use
a burn-in period of 1000 steps to determine the R vec-
tor (in this case R2004
M CM C
fell below 1.02 for all K=13 parameters as summarized
in Table 4, and thus demonstrated more than adequate
convergence. Our MCMC results are also summarized in
Table 4.
M CM C ). The parameters of R2004
As summarized in the bottom-panel of Figure 7 these
results indicate that the 2004 MOST data-set provides a
strong independent argument that solar-type differential
rotation pattern defined by Equation 1 has been observed
in κ1 Ceti. Using only the 2004 data-set the values of
the differential rotation coefficient, and equatorial period
are: k = 0.092 - 0.101, and PEQ = 8.65 - 8.72d. These
values were determined by averaging for each point of
the MCMC chains the three possible k and PEQ values,
between spots 1 & 2, 1 & 3, and 2 & 3.
5. DISCUSSION
The MCMC Solution 1 in Table 3 provides our best
values for the various spot and stellar parameters. The
solutions are expressed by the 68% credible ranges rather
than best values with standard deviations since the like-
lihood histograms are non-gaussian. Figure 7 is a good
summary of our analysis. The periods and β 68%
marginalized contours are shown for each of the spots to-
gether with mean and limiting curves for k (0.090, 0.085,
0.096) and the range of PEQ (8.77, 8.74, 8.81d). Each
of the three years is distinguished by a different color.
β ranges from 10◦ and 75◦ with the 2004 data obvi-
ously providing the most rigorous constraint on k. In-
deed, it demonstrates that Rucinski et al. (2004) had a
most challenging task in demonstrating differential ro-
tation because the spots in 2003 were at very similar
latitudes with the second spot being particularly small.
The analysis defined k using Equation 1 which is de-
rived from the solar pattern. The agreement of all seven
spots with the form of the k curve as well as the good
agreement of the spot solutions in Figures 5, 6 and 7 sug-
gests that the the differential rotation curve for κ1 Cet
is closely similar to solar. The analysis of the 2004 light
curve independently of any assumption about k offers
strong confirmation that the pattern is indeed solar.
For
the Sun, k has been derived quite
inde-
pendently from either sunspot latitudes and periods
(Newton and Nunn 1951) or surface radial velocities
(Howard et al. 1983) yielding k = 0.19 and 0.12, respc-
tively (Kitchatinov 2005). The large discrepancy be-
tween these values may be related in part to the different
behavior of sunspot and photospheric motions.
In our
case, we depend on spots to determine k and our value
of 0.09 is significantly lower than either of those for the
Sun. This is in line with calculations by Brown et al.
(2004) who find that k should increase with age.
Gudel et al. (1997) estimated an age of 750 Myr for
κ1 Cet from their estimated 9.2 d rotation period. Our
value of PEQ = 8.77 d suggests a still younger age.
All of the photometric periods found to date for κ1 Cet
can be explained by spots appearing at different lati-
tudes. A period of 9.09 d is given in the Hipparcos cat-
alog (ESA 1997) while Messina & Guinan (2002) quote
a value of 9.214 d. Baliunas et al. (1995) monitored Ca
II H & K photoelectrically between 1967 and 1991 and
found a rotational period of 9.4 ± 0.1 d (Baliunas et al.
1983) and Shkolnik et al. (2003) found a closely simi-
lar period of ∼9.3 d mostly from spectra taken in 2002.
While the apparent persistance of this period which cor-
responds to a range of 50◦ to 60◦ in latitude for some 35
years maybe fortuitous, it might be related to some large
scale magnetic structure.
The Natural Sciences and Engineering Research Coun-
cil of Canada supports the research of B.C., D.B.G.,
J.M.M., A.F.J.M., S.M.R., G.A.H.W.. Additional sup-
port for A.F.J.M. comes from FCAR (Qu´ebec). R.K. is
supported by the Canadian Space Agency. W.W.W. is
supported by the Austrian Space Agency and the Aus-
trian Science Fund (P14984).
8
Walker et al.
κ1 Ceti MCMC fitted parameters
TABLE 3
parameter
fitted
Rucinski
2004 results
Parallel Tempering Minimum χ2
results
solution
MCMC
Solution 1
Reduced χ2
ν a
i (o)
k
PEQ
vsini (km s−1) c
equatorial speed (km s−1) c
u
κω
U2003
E2003 1 (JD-2451545)
p2003 1 (days)
β2003 1 (o)
γ2003 1 (o)
E2003 2 (JD-2451545)
p2003 2 (days)
β2003 2 (o)
γ2003 2 (o)
U2004
E2004 1 (JD-2451545)
p2004 1 (days)
β2004 1 (o)
γ2004 1 (o)
E2004 2 (JD-2451545)
p2004 2 (days)
β2004 2 (o)
γ2004 2 (o)
E2004 3 (JD-2451545)
p2004 3 (days)
β2004 3 (o)
γ2004 3 (o)
U2005
E2005 1 (JD-2451545)
p2005 1 (days)
β2005 1 (o)
γ2005 1 (o)
E2005 2 (JD-2451545)
p2005 2 (days)
β2005 2 (o)
γ2005 2 (o)
n/a
n/a
yes
yes
yes
derived
derived
assumed
assumed
yes
yes
derived
yes
yes
yes
-
-
70 ± 4
-
-
4.64 ± 0.11 b
-
0.80
0.00
-
-
8.9 ± 0.1 d
40 ± 7
11.1 ± 0.6
3.145 - 3.374
817
44.9 - 74.1
0.094 - 0.117
8.71 - 8.90
3.83 - 5.34
5.40 - 5.53
0.6840
0.220
0.9964 - 1.0031
1410.94 - 1410.97
8.978 - 9.007
16.7 - 36.9
11.05 - 11.93
2.936
817
60.6
0.087
8.784
4.77
5.47
0.684
0.220
1.0011
1410.93
9.008
32.4
11.75
2.968- 3.018
817
57.8 - 63.5
0.085 - 0.096
8.74 - 8.81
4.64 - 4.94
5.46 - 5.50
0.684
0.220
1.0003 - 1.0017
1410.92 - 1410.94
8.990 - 9.015
29.5 - 34.8
11.63 - 11.86
-
1405.85 - 1406.23
1406.10
1406.07 - 1406.16
derived
9.3 − 9.7 d
yes
yes
yes
yes
derived
yes
yes
yes
derived
yes
yes
yes
derived
yes
yes
yes
yes
derived
yes
yes
yes
derived
yes
yes
-
-
n/a
n/a
n/a
n/a
n/a
n/a
n/a
n/a
n/a
n/a
n/a
n/a
n/a
n/a
n/a
n/a
n/a
n/a
n/a
n/a
n/a
n/a
8.948 - 9.160
-45.7 - 58.9
5.08 - 9.41
1.0025 - 1.0092
1756.69 - 1756.78
8.721 - 8.939
-0.8 - 15.0
7.17 - 8.51
9.070
37.2
5.95
1.0149
1756.68
8.820
12.7
7.99
9.022 - 9.094
32.9 - 39.8
5.68 - 6.18
1.0129 - 1.0150
1756.66 - 1756.70
8.787 - 8.851
9.3 - 16.8
7.73 - 8.09
1763.53 - 1763.65
1763.54
1763.49 - 1763.56
9.039 - 9.316
-43.9 - -28.7
7.14 - 19.21
1769.01 - 1769.32
9.572 - 9.751
59.2 - 76.9
7.53 - 13.23
1.0000 - 1.0053
2116.96 - 2117.09
9.340 - 9.419
41.8 - 63.0
8.30 - 9.37
2120.09 - 2120.21
9.094 - 9.253
28.0 - 56.0
7.69 - 8.31
9.200
-46.3
16.76
1769.09
9.572
77.3
13.04
1.0042
2116.94
9.372
58.4
9.74
9.153 - 9.231
-47.6 - -43.2
14.44 - 17.31
1769.03 - 1769.16
9.542 - 9.640
74.9 - 78.4
11.60 - 13.53
1.0029 - 1.0051
2116.95 - 2117.05
9.365 - 9.423
55.4 - 62.1
9.24 - 10.28
2120.22
2120.21 - 2120.30
9.248
49.6
8.55
9.166 - 9.252
42.9 - 50.1
7.94 - 8.52
a ν is the number of binned data points minus the number of fitted parameters.
b spectroscopic value from Rucinski et al. (2004), not derived photometrically.
c These values determined using R∗ = 0.95 RJ.
d This parameter fitted, rather than derived (Rucinski et al. 2004).
REFERENCES
Baliunas, S.L., Donahue, R.A., Soon, W.H., Horne, J.H., Frazer,
J., Woodard-Eklund, L., Bradford, M., Rao, L.M., Wilson,
O.C., Zhang, Q., Bennett, W., Briggs, J., Carroll, S.M.,
Duncan, D.K., Figueroa, D., Lanning, H.H., Misch, T., Mueller,
J., Noyes, R.W., Poppe, D., Porter, A.C., Robinson, C.R.,
Russell, J., Shelton, J.C., Soyumer, T., Vaughan, A.H.,
Whitney, J.H., 1995, ApJ, 438, 269
Baliunas, S.L., Hartmann, L., Noyes, R.W., Vaughan, H.,
Preston, G.W., Frazer, J., Lanning, H., Middelkoop, F.,
Mihalas, S., 1983, ApJ, 275, 752
Brown, B. P., Browning, M. K., Brun, A. S., Toomre, J. 2004,
Proceedings of the SOHO 14 / GONG 2004 Workshop (ESA
SP-559). "Helio- and Asteroseismology: Towards a Golden
Future". 12-16 July, 2004. New Haven, Connecticut, USA.
Editor: D. Danesy., p.341
Brun A.S., Toomre J. 2002, ApJ, 570, 865
Budding, E. 1977, Ap&SS, 48, 207
Collier Cameron, A. 2002, Astron. Nachr., 323, 336
Croll, B. 2006, PASP, accepted 18 July
Croll, B., Walker, G. A. H., Kuschnig, R., Matthews, J.M., Rowe,
J.F., Walker, A., Rucinski, S.M., Hatzes, A.P., Cochran, W.D.,
Robb, R.M., Guenther, D.B., Moffat, A.F.J., Sasselov, D.,
Weiss, W.W., 2006, ApJ, 648, 607
D´ıaz-Cordov´es, J., Claret, A., Gimin´enez, A. 1995, Ap&SS, 110,
329
Dorren, J.D. 1987, ApJ, 320, 756
European Space Agency. 1997. The Hipparcos and Tycho
Catalogues (ESA SP-1200)(Noordwijk: ESA) (HIP)
Gamerman D., 1997, Markov Chain Monte Carlo: Stochastic
Simulation for Bayesian Inference, Chapman and Hall, London
Gelman, A., Rubin, D.B. 1992, Statistical Science, 7(4), 457
Gilks, W.R., Richardson, S., Spiegelhalter, D.J., 1996, Markov
Chain Monte Carlo in Practice, Champan and Hall, London
Gregory, P.C. 2005a, ApJ, 631, 1198
Differential rotation of κ1 Ceti
9
κ1 Ceti 2004 MCMC fitted parameters
TABLE 4
parameter
fitted
Reduced χ2
ν a
i (o)
k
vsini (km s−1) b
PEQ
u
κω
U2004
E2004 1 (JD-2451545)
p2004 1 (days)
β2004 1 (o)
γ2004 1 (o)
E2004 2 (JD-2451545)
p2004 2 (days)
β2004 2 (o)
γ2004 2 (o)
E2004 3 (JD-2451545)
p2004 3 (days)
β2004 3 (o)
γ2004 3 (o)
n/a
n/a
assumed
derived
derived
derived
assumed
assumed
yes
yes
yes
yes
yes
yes
yes
yes
yes
yes
yes
yes
yes
R2004
P arallel Parallel Tempering R2004
results
M CM C
n/a
n/a
n/a
n/a
n/a
n/a
n/a
n/a
1.122
1.110
1.127
1.183
1.092
1.315
1.279
1.162
1.074
1.034
1.102
1.203
1.127
1.50 - 1.82
269
60.6
0.095 - 0.111
4.80 - 4.86
8.624 - 8.721
0.6840
0.220
1.0051 - 1.0150
1756.66 - 1756.74
8.704 - 8.773
9.1 - 22.3
7.51 - 7.80
1763.40 - 1763.50
9.050 - 9.162
-45.8 - -38.6
11.21 - 15.80
1768.94 - 1769.07
9.518 - 9.598
71.0 - 77.9
9.04 - 13.11
n/a
n/a
n/a
n/a
n/a
n/a
n/a
n/a
1.011
1.001
1.003
1.003
1.002
1.000
1.002
1.013
1.016
1.002
1.000
1.003
1.003
MCMC
Solution
1.46 - 1.52
269
60.6
0.092 - 0.101
4.80 - 4.84
8.65 - 8.72
0.684
0.220
1.0139 - 1.0150
1756.64 - 1756.68
8.736 - 8.788
15.6 - 21.7
7.64 - 7.84
1763.43 - 1763.49
9.112 - 9.183
-47.5 - -43.9
14.65 - 16.83
1768.97 - 1769.08
9.510 - 9.579
76.6 - 78.0
12.64 - 13.17
a ν is the number of binned data points minus the number of fitted parameters.
b These values determined using R∗ = 0.95 RJ.
Gregory, P.C. 2005b, Bayesian Logical Data Analysis for the
Physical Sciences: A Comparative Approach with Mathematica
Support (Cambridge: Cambridge Univ. Press)
Gudel, M., Guinan, E.F., & Skinner, S.L., 1997, ApJ, 483, 947
Hall, D.S. & Busby, M.R. 1990, in Active Close Binaries, NATO
ASIC Proceedings 319, Nato Advance Study Institute, p. 377
Henry, G.W., Eaton, J.A., Hamer, J. & Hall, D.S. 1995, ApJS,
97, 513
Howard, R., Adkins, J.M., Boyden, J.E., Cragg, T.A., Gregory,
T.S., Labonte, B.J., Padilla, S.P., Webster, L. 1983, Solar Phys.
83, 321
Kitchatinov, L.L. 2005, Physics - Uspekhi 48 (5), 449
Lanza, A. F., Catalano, S., Cutispoto, G., Pagano, I., Rodono,
M., 1998, A&A, 332, 541
Lebedinskii, A.I. 1941, Astron. Zh., 18, 10
Levenberg, K., 1944, Quart. Appl. Math. 2, 164
Lewis, A., Bridle, S. 2002, PhRvD, 66, 103511
Marquardt, D.W., 1963, SIAM J. Appl. Math, 11, 431
Messina, S. & Guinan, E. F. 2002, A&A, 393, 225
Newton, H.W., Nunn, M.L. 1951, MNRAS, 111, 413
Ossendrijver, M. 2003, Astronomy and Astrophysics Review,
Volume 11, Issue 4, 287
Press, W.H., Flannery, B., Teukolsky, S.A., Vetterling, W.T.
1992. Numerical Recipes in Fortran, 2nd ed. (Cambridge:
Cambridge Univ. Press)
Rib´arik, G., 2002, Occasional Technical Notes from Konkoly
Observatory No. 12,
http://www.konkoly.hu/Mitteilungen/otn12.ps.Z
Rib´arik, G., Ol´ah, K., Strassmeier, K. G. 2003, Astronomische
Nachrichten, 324, 202
Rucinski, S.M., Walker, G. A. H., Matthews, J.M., Kuschnig, R.,
Shkolnik, E., Marchenko, S., Bohlender, D.A., Guenther, D. B.,
Moffat, A. F. J., Sasselov, D., Weiss, W. W. 2004, PASP, 116,
1093
Sajina, A., Scott, D., Dennefeld, M., Dole, H., Lacy, M., Lagache,
G., astro-ph 0603614
Shkolnik, E., Walker, G.A.H., Bohlender, D.A., 2003, ApJ,
597,1092
Thompson, M.J., Christensen-Dalsgaard, J., Miesch, M.S. 2003,
Annu. Rev. Astron. Astrophys., 41, 599
Walker, G.A.H., Matthews, J.M., Kuschnig, R., Johnson, R.,
Rucinski, S., Pazder, J., Burley, G., Walker, A., Skaret, K., Zee,
R., Grocott, S., Carroll, K., Sinclair, P., Sturgeon, D., Harron,
J., 2003a, PASP, 115, 1023
10
Walker et al.
1.00
0.99
0.98
0.97
0.96
0.01
0.00
1.00
0.99
0.98
0.01
0.00
1.00
0.99
0.98
0.01
0.00
2003
1405
1410
1415
1420
1425
1430
1435
2004
1745
1750
1755
1760
1765
1770
2005
2105
2110
2115
2125
HJD - 2451545
2120
2130
2135
i
a
n
g
s
d
e
z
i
l
l
a
m
r
o
n
Fig. 1. -- MOST light curves for κ1 Ceti from 2003, 2004, and 2005. The points are the mean signals from individual 101.413 min satellite
orbits. The black connected symbols in the lower part of each panel are the the differences between the highest and lowest of 7 adjacent
simultaneously recorded (background) Fabry images (see text). The 'dip' at 1762 in 2004 corresponds to a total eclipse of the Moon.
Differential rotation of κ1 Ceti
11
0.04
0.035
0.03
0.025
0.02
0.015
0.01
0.005
0
0.07
0.08
0.09
0.1
0.11
0.12
k
0.018
0.016
0.014
0.012
0.01
0.008
0.006
0.004
0.002
0
0.014
0.012
0.01
0.008
0.006
0.004
0.002
0
50 52 54 56 58 60 62 64 66
i
8.65
8.7
8.75
8.8
8.85
8.9
PEQ
Fig. 2. -- The marginalized likelihood for each of the fitted parameters is shown by the thick curve. The thin dotted curve shows the
mean likelihood. The best-fitting value (the peak of the distribution) and the 68% credible regions determined from the marginalized
likelihood are shown by the solid vertical, and dashed vertical lines. The unusual spike in the mean likelihood curve for a low value of k in
the k histogram is likely due to low numbers of statistics, as indicated by the negligible value of the marginal likelihood curve.
12
Walker et al.
68
66
64
62
60
58
56
54
52
50
48
68
66
64
62
60
58
56
54
52
50
48
68
66
64
62
60
58
56
54
52
50
48
68
66
64
62
60
58
56
54
52
50
48
i
i
i
i
68
66
64
62
60
58
56
54
52
50
48
25
30
35
40
45
50
β20032
68
66
64
62
60
58
56
54
52
50
48
-52 -50 -48 -46 -44 -42 -40 -38 -36
β20042
68
66
64
62
60
58
56
54
52
50
48
45
50
55
60
65
70
β20051
68
66
64
62
60
58
56
54
52
50
48
0.06 0.07 0.08 0.09 0.1 0.11 0.12 0.13
k
i
i
i
i
20 22 24 26 28 30 32 34 36 38 40
β20031
0
5
10
15
β20041
20
25
30
64 66 68 70 72 74 76 78 80 82
β20043
30
35
40
50
55
60
45
β20052
Fig. 3. -- The solid lines are the 68% and 95% credible regions of the marginalized likelihood for various parameters. The 68% mean
likelihood credible regions are shown by dotted contours (see text). Obviously there is little correlation between k and i in the bottom
right panel.
Differential rotation of κ1 Ceti
13
1.000
0.995
0.990
0.985
0.980
0.975
0.970
0.965
0.960
0.955
0.020
0.015
0.010
0.005
0.000
-0.005
-0.010
-0.015
-0.020
1405.0
1410.0
1415.0
1420.0
HJD - 2451545
1425.0
1430.0
Fig. 4. -- The best-fitting two spot solution for κ1 Ceti in 2003 (rotating counter-clockwise from top) and seen from the line of sight at
phases 0.00, 0.25, 0.50, and 0.75 (from left) of the first spot. Middle: the MOST light curve with errors (see text). The continuous line is
the solution from the "Minimum χ2" column of Table 3. The dotted line indicates the unspotted normalized signal of the star (U =1.0011).
Vertical dashed lines indicate phases 0.00, 0.25, 0.50 and 0.75. Bottom: residuals from the model on the same scale.
l
i
a
n
g
S
d
e
z
i
l
a
m
r
o
N
14
Walker et al.
1.015
1.010
1.005
1.000
0.995
0.990
0.985
0.015
0.010
0.005
0.000
-0.005
-0.010
-0.015
1750.0
1755.0
1760.0
HJD - 2451545
1765.0
Fig. 5. -- The best-fitting three spot solution for κ1 Ceti in 2004 (rotating counter-clockwise from top) and seen from the line of sight at
phases 0.00, 0.25, 0.50, and 0.75 (from left) of the first spot. Middle: the MOST light curve with errors (see text). The continuous line is
the solution from the "Minimum χ2" column of Table 3. The dotted line indicates the unspotted normalized signal of the star (U =1.0149).
Vertical dashed lines indicate phases 0.00, 0.25, 0.50 and 0.75. Bottom: residuals from the model on the same scale.
l
i
a
n
g
S
d
e
z
i
l
a
m
r
o
N
Differential rotation of κ1 Ceti
15
1.005
1.000
0.995
0.990
0.985
0.980
0.010
0.005
0.000
-0.005
-0.010
2112.0
2114.0
2116.0
2118.0
2120.0
2122.0
2124.0
2126.0
HJD - 2451545
Fig. 6. -- The best-fitting two spot solution for κ1 in Ceti 2005 (rotating counter-clockwise from top) and seen from the line of sight at
phases 0.00, 0.25, 0.50, and 0.75 (from left) of the first spot. Middle: the MOST light curve with errors (see text). The continuous line is
the solution from the "Minimum χ2" column of Table 3. The dotted line indicates the unspotted normalized signal of the star (U =1.0042).
Vertical dashed lines indicate phases 0.00, 0.25, 0.50 and 0.75. Bottom: residuals from the model on the same scale.
l
i
a
n
g
S
d
e
z
i
l
a
m
r
o
N
16
Walker et al.
80
70
60
50
40
30
20
10
8.6
8.8
9
9.2
p (days)
9.4
9.6
9.8
0
10
80
70
60
50
40
30
20
10
β (o)
β (o)
8.6
8.8
9
9.2
p (days)
9.4
9.6
9.8
0
10
Fig. 7. -- Top: The 68 % marginalized likelihood contours for the modulus of the latitude, β, and spot period are shown by dots for
each spot -- 2003 (green), 2004 (blue) and 2005 (pink). The central red curve is the solar-period, latitude relation (Equation 1) with the
mean values taken from Solution 1 of Table 3, k = 0.090 and PEQ = 8.77d. The other two red lines correspond to k = 0.085, PEQ = 8.74d,
and k = 0.096, PEQ = 8.81d. The vertical dashed lines indicate the range of rotational periods (p = 9.4 ± 0.1 days) found in Ca II H & K
line reversals over ∼30 years. Bottom: The 68 % marginalized likelihood contours for the 2004 data only (cyan) while fitting not for k and
PEQ but explicitly for the periods, p2004 1, p2004 2, and p2004 3. The black curve is the best-fitting curve (k = 0.096 and PEQ = 8.68d)
for the 2004 data, while the red curves are the same as in the upper panel. As can be seen, the black curve fits the data well and thus the
observed differential rotation pattern is closely similar to solar.
|
astro-ph/0403478 | 1 | 0403 | 2004-03-19T15:48:58 | Galaxy Kinematics with SALT | [
"astro-ph"
] | The combination of dynamical and photometric properties of galaxies offers a largely un-tapped source of information on how galaxies assembled and where stars formed. Bi-dimensional kinematic measurements have been the stumbling block. The light-gathering power of SALT coupled with the high-throughput performance of the Prime Focus Imaging Spectrograph (PFIS) yield a superb facility for measuring velocity-ellipsoids of stars and gas in galaxies out to gigaparsec distances. From these data dynamical asymmetries arising from lopsided or elliptical halos may be probed; disk-mass and mass-decompositions may be uniquely determined; mechanisms for disk heating constrained; and a zeropoint for the mass-to-light ratios of stellar populations set. A number of groups within the SALT consortium are interested in making these measurements using a variety of different, but complementary approaches. The scientific potential from their synthesis is very promising. We describe some unusual observational modes in which PFIS may be used to probe the shape of dark-matter halos and the content of galaxy disks. | astro-ph | astro-ph |
First Robert Stobie SALT Workshop
Science with SALT Workshop Proceedings, Vol. 2, 2004
D.A.H. Buckley
Galaxy Kinematics with SALT
M. A. Bershady1, M. A. W. Verheijen2, D. R. Andersen3, R. A.
Swaters4, and K. B. Westfall1
1Department of Astronomy, University of Wisconsin -- Madison, 475 N.
Charter St, Madison, WI 57306, USA
2Astrophysikalisches Institut Potsdam, An der Sternwarte 16, 14482,
Potsdam, Germany
3Max Planck Institut fur Astronomie, Konigstuhl 17, 69117 Heidelberg,
Germany
4Department of Astronomy, University of Maryland, College Park, MD
20742, USA
Abstract.
The combination of dynamical and photometric properties of galaxies
offers a largely un-tapped source of information on how galaxies assembled
and where stars formed. Bi-dimensional kinematic measurements have
been the stumbling block. The light-gathering power of SALT coupled
with the high-throughput performance of the Prime Focus Imaging Spec-
trograph (PFIS) yield a superb facility for measuring velocity-ellipsoids
of stars and gas in galaxies out to gigaparsec distances. From these data
dynamical asymmetries arising from lopsided or elliptical halos may be
probed; disk-mass and mass-decompositions may be uniquely determined;
mechanisms for disk heating constrained; and a zeropoint for the mass-
to-light ratios of stellar populations set. A number of groups within the
SALT consortium are interested in making these measurements using a
variety of different, but complementary approaches. The scientific po-
tential from their synthesis is very promising. We describe some unusual
observational modes in which PFIS may be used to probe the shape of
dark-matter halos and the content of galaxy disks.
1.
Introduction
Extant information about galaxies comes primarily from broad-band optical
images. These deliver a wealth of information on when and where stars have
formed. Missing is an accurate understanding of the distribution of mass and the
details of its assembly -- knowledge which requires dynamical arguments substan-
tiated with kinematic measurements. Hence kinematic information is essential
for a basic understanding of galaxy formation and evolution. The last decade
has seen the growth of bi-dimensional optical spectroscopy -- once the purview
of radio (HI) observations. SALT can continue this development. With SALT's
Prime Focus Imaging Spectrograph we will be able to measure velocity-ellipsoid
maps for a variety of dynamical tracers which include (collisionless) stellar pop-
1
2
Bershady et al.
Figure 1.
SparsePak (left, Bershady et al. 2003, 2004) and PPAK
(right, Verheijen et al. 2004) IFUs, which respectively feed spectro-
graphs on the WIYN and Calar Alto 3.5m telescopes, both capable of
∼11 km/s (σ) resolution. Units are presented at the same subtended
angular scale at the telescope focal-plane; physical dimensions scale
with respective telescope+fore-optic f -ratios. Light fibers transmit to
the spectrograph; dark fibers serve as mechanical buffers.
ulations and ('sticky') ionized gas. These measurements will be accessible via
Fabry-Perot and dispersed spectroscopy over a very large (8′) field.
As promising for SALT is the critical mass of consortium scientists with
overlapping interests in galaxy dynamics (see, for example, contributions by
Balona, Cecil, Cote, Crawford, Ferrarese, Sellwood, Sparke, Wilcots, Williams,
and Ziegler in this volume). A few specific scientific prospects of interest to
members of the consortium include (but are not limited to!) the frequency and
amplitude of lopsided disks and triaxial halos as probed by dynamical and pho-
tometric asymmetries; mergers and perturbations inferred studies of kinematic
irregularities; disk heating mechanisms judged from velocity dispersion ellipsoids
(σz/σR, Gerssen et al. 2000); feed-back processes in disks and the ISM; bar dy-
namics and their pattern-speeds; the Tully-Fisher relation and its evolution; and
direct dynamical mass decompositions determined by combining rotation curves
with stellar velocity dispersions.
Here we focus on how well we can directly separate disk from halo mass in
spirals to yield dark-halo profiles, the zeropoint for stellar M/L ratio, and plau-
sible constraints on the faint-end of the stellar IMF. These are of fundamental
interest, respectively for testing both hierarchical cold-dark-matter structure-
formation scenarios and star-formation models. This particular science case al-
lows us to illustrate a novel application of PFIS to bi-dimensional spectroscopy
with wide application. The concept combines narrow-band filters with dispersed
spectroscopy to form what we call massively multi-plexed spectroscopy, or MMS.
The approach complements other methods using fluid-dynamical modeling and
kinematic measurements of barred potentials (Weiner et al. 2001; Gerssen et al.
2003; Debattista & Williams 2004).
Galaxy Kinematics with SALT
3
Figure 2.
SparsePak spectra of a K0.5 III template-star and
azimuthally-averaged fiber rings for NGC 3982 at 6 different radii,
labeled in units of disk radial scale-length hR. Auto- and cross-
correlations are shown (right) for ∼10nm of spectra in the MgI region.
2. Mass Decomposition
The traditional technique of decomposing galaxy rotation curves into disk and
halo mass contributions is inherently degenerate (see discussion by Sellwood,
this volume). One way to break the disk-halo degeneracy is to measure directly
the disk potential via the vertical motions and scale-height of the stars. To
within factors of order unity: σz = pπG z0 µ M/L, where σz is the vertical
component of the disk velocity dispersion, z0 is the disk scale height, µ the sur-
face light-intensity, and M/L the mass-to-light ratio (hence µ M/L is the disk
surface mas-density, Σ). Recent studies of edge-on galaxies (Kregel et al. 2002)
permit vertical scale-heights to be inferred with reasonable accuracy for face-on
systems given the observed correlations between the radial scale-length, hR, ro-
tation speed, type, and z0. Face-on disks are ideal for measuring σz, typically the
smallest component of the velocity ellipsoid, but low inclination makes measure-
ment of inclination and rotation speed difficult. Inclined disks require careful
correction and decomposition of the observed, line-of-sight velocity-dispersion
to extract σz. Past attempts to realize this kinematic approach (e.g., Bottema
4
Bershady et al.
10
8
6
4
2
0
K
L
/
M
g
o
l
1
0
−1
−2
14
16
18
20
22
24
constant star−formation rate
stars + gas
1.7−2.6h
1h
3.5h
const. SFR, stars only
simple
stellar
pop.
Age
(Gyr)
1010
109
108
107
0
1
2
3
4
1
2
3
4
B−K
Figure 3. Direct kinematic surface-mass density and M/L measure-
ments from integral-field spectroscopy at 515nm and optical-NIR pho-
tometry of NGC 3982: Surface mass-density (Σ) and surface-brightness
(µ) vs radius (left); K-band M/L vs rest-frame B − K color (right).
Dashed line (left) represents a reference Σ for the MW in the solar
neighborhood (Kuijken & Gilmore 1991). Model curves (right) are
adopted from Bruzual & Charlot (1993) and Bell & De Jong (2000).
Error limits show systematic and random components.
1997) have not overcome these problems, and have been limited by long-slit
spectroscopy that does not reach well beyond a single disk scale length.
The observational approach can be rectified by application of integral-field
observations of nearly face-on disks at medium spectral resolutions (7000 <
λ/∆λ < 15000) and multi-band optical and NIR imaging. Two suitable integral-
field units (IFUs) are shown in Figure 1. Galaxy sizes can be carefully matched
to the IFUs. Observations with these units yield kinematic maps from which
accurate inclinations can be derived to i ∼ 15◦ (Andersen & Bershady 2003).
Most importantly, the area sampled with IFUs increases with radius. For nearly
face-on disks, one may average the signal in rings, which greatly enhances the
limiting depth and radius of the observations.
Using the two IFUs shown in Figure 1 we have undertaken a survey of
nearly face-on, nearby disk galaxies to determine their disk masses. The survey
has yielded Hα kinematic maps for ∼100 normal, late-type galaxies. We are now
gathering absorption-line spectroscopic observations (for stellar kinematics) for
40 of these galaxies, selected to have regular velocity fields (see Verheijen et al.
2004, Figure 2).
Results from a pilot target, NGC 3982, yield σz out to 3.5 disk scale-lengths
in this albeit high-surface-brightness disk in both the MgI (515nm; Figure 2) and
CaII triplet (860nm) regions. Cross-correlations yield similar results and show
that the trend of disk surface-density well traces near-infrared surface-brightness
-- mass traces light! -- as shown in Figure 3. Contrary to our earlier, preliminary
reports, a renewed analysis of the data indicates a K-band M/L of the disk
which is consistent with a maximum-disk situation in the sense of van Albada &
Sancisi (1986). The colors and dynamical M/L measurements in Figure 3 agree
Galaxy Kinematics with SALT
5
Examples of Massively Multiplexed Spectroscopy (MMS)
Figure 4.
compared to Long-slit and Fabry-Perot spectroscopy with PFIS. Each
present a different slice of the spatial (θ) × spatial (θ) × wavelength
(λ) data-cube. "NB" is a narrow-band filter, which in the case of PFIS
has a bandwidth of λ/∆λ = 50. Grey regions denote used portions
of the CCD. Long-slit spectroscopy makes best use of the CCD array,
while MMS and Fabry-Perot modes use only ∼50% compared to the
Long-slit case.
well with stellar population synthesis models of, e.g., Bell & De Jong (2001),
yet this source lies over 2 mag below TF. Here is a small, blue high-surface-
brightness, actively star-forming disk, which is rapidly rotating and dominated
by normal, baryonic matter out to 3-4 scale-lengths of the light distribution.
At face value, this calibration of Bell & De Jong's (2001) models implies, as
they note, a Salpeter-like IMF truncated below 0.35 M⊙. It is critical to test
the generality of this result in a variety of "normal" systems over a range in
type, color and surface-brightness. Our on-going survey will begin to address
this issue, but the low-surface-brightness regime will be difficult to probe at the
required spectral resolution using 4m-class telescopes.
3. Prospects with the Prime Focus Imaging Spectrograph
The power of PFIS lies in its unique combination of 3 Fabry-Perot etalons span-
ning a factor of 50 in resolution, 6 gratings covering from the UV-atmospheric
cutoff to 850 nm at resolutions up to λ/∆λ ∼ 13000, and a suite of narrow-
band filters filling an octave at optical wavelengths. All of these options are
6
Bershady et al.
configurable with imaging, longslit, or multi-slit mode. Most of the gratings are
high-transmissivity volume-phase holographic (VPH) elements, tunable with an
articulating camera.
For most galaxy kinematic studies, long-slit measurements are inefficient:
Two-dimensional sampling is needed to gain velocity field information and to
gather more light at large radii. Fabry-Perot imaging, in contrast, is excel-
lent for gas and stellar velocity fields based on single, strong lines. How-
ever, for velocity dispersion work in narrow-lined (low-mass) systems, ideally
many (weak) lines are probed, which requires scanning a significant band-pass
(λ/∆λ < 100). Hence an intermediate trade between the spectral and spa-
tial sampling of longslit and Fabry-Perot spectroscopy is needed. Integral- or
formatted-field spectroscopy, with limited spatial coverage is one alternative,
but at first-blush PFIS appears to have no such capability.
Other possibilities include using different tracers, such has planetary neb-
ulae (e.g., Ciardullo et al. 2003; Merrett et al. 2003). Traditionally this re-
quires narrow-band imaging and multi-object spectroscopic follow-up. PFIS is
well suited for both. While this is efficient, newer techniques, such as counter-
dispersed imaging (e.g., Douglas et al. 2002) also can be accomplished with
PFIS. Another approach for multi-object emission-line kinematics would be to
use slitless, dispersed imaging, i.e., slitless grism spectroscopy, at high spec-
tral resolution over a limited "on" and "off" band-pass using a combination
of gratings and narrow-band filters. The on and off bands replace the func-
tion of counter-dispersion. The technique should be considered by prospective
PFIS users. We describe, instead, an alternative method using gratings and
narrow-band filters with slits for IFU-like spectroscopy -- suitable for emission
and absorption line studies of extended sources.
3.1. MMS: Massively Multiplexed Spectroscopy
The wavelength-spatial multiplex trade for galaxy kinematic studies can be ac-
complished with PFIS by combining the suite of λ/∆λ ∼ 50 narrow-band filters
with the tunable VPH gratings. The filters serve to limit the range of the dis-
persed spectra on the detector, thereby allowing for an increase in the spatial
multiplex: Multiple, parallel slits or slit-lets can be placed on a mask and pro-
duce non-overlapping spectra covering roughly, e.g., 10nm centered at 515nm --
just what is shown in Figure 2 from SparsePak. The concept for mapping ex-
tended sources is illustrated in Figure 4; multi-object applications are possible
(Crawford, this volume).1
The minimum spatial separation in the dispersion direction is ∼1 arcmin at
the highest spectral resolutions (grating angle α = 50◦), and decreases linearly
with resolution (∝ tan α, i.e., the Littrow condition for the VPH gratings).
At the highest spectral resolutions MMS yields a spatial multiplex increase of
roughly a factor of 4. For lower-dispersion setting, the spatial multiplex
can be significantly increased. At any resolution, the slits can be staggered
to yield a variety of two-dimension sampling patterns (Figure 4) that resemble
1The Fabry-Perot image is not truly monochromatic (there is a radial field-dependence to the
band-pass), and the MMS slits are not uniformly spaced due to field-dependence of the grating
incidence angle.
Galaxy Kinematics with SALT
7
PFIS/FP
RFPI
GHASP
VIMOS
SAURON
INTEGRAL
OASIS
1000 10000
10000
1000
100
10
PFIS/VPH + NB
PFIS/VPH
PPAK
SparsePak
DensePak
GMOS
PMAS
GIRAFFE/
ARGUS
Figure 5. Grasp versus spectral power for two-dimensional spectro-
scopic systems on 4m and 8m-class telescopes (solid and dashed lines,
respectively, with Fabry-Perot instruments shown as filled circles). To-
tal grasp is the product of area × solid-angle (AΩ). Spectral power
is the product of the spectral resolution, λ/∆λ, times the number of
spectral resolution elements, N∆λ. Variations in the shapes of covered
parameter space depends on how a given instrument achieves a range of
spectral resolution and spatial sampling, i.e., through changes in grat-
ings, slit-widths, or both. Note the unique location of PFIS in these
diagrams is significantly extended with the use of MMS.
e.g., SparsePak. Such data are suitable for constructing velocity fields and dis-
persion maps. Hence MMS should provide a competitive approach to the stellar
dynamical studies illustrated in the previous section.
It is useful to examine where PFIS lies in comparison to other bi-dimensional
spectrographs in terms of "information gathering power," which we loosely
parametrize with the two quantities "total grasp" and "spectral power". These
parameters are defined in Figure 5, where it is seen that PFIS has very large
grasp while spanning a wide range in spectral power. The addition of MMS
extends the range in grasp and spectral power, bridging between Fabry-Perot
and traditional slit-spectroscopy modes. The light-gathering power of PFIS and
SALT is over an order of magnitude larger than the 4m-class IFU instruments
we have discussed. As such, PFIS can extend the study of galaxy kinematics
significantly into the low surface-brightness regime.
4. Summary and Discussion
Galaxy kinematic measurements with SALT's PFIS can be used to answer many
outstanding, and fundamental questions about the structure and formation of
galaxies. Highlighted here is one example of unique, dynamical mass decompo-
8
Bershady et al.
sitions of spiral galaxies. A number of observational modes are available with
PFIS. We have described one powerful, new mode, referred to as "massively
multiplexed spectroscopy," (MMS), which is enabled by the unique combination
of VPH gratings and narrow-band filters in PFIS. With MMS and Fabry-Perot
imaging, PFIS is a highly competitive survey instrument for two-dimensional
studies of gas and stars in nearby galaxies. There are several groups within the
SALT consortium interested in applying these to a similar range of dynamical
problems. This holds the promise of fruitful collaboration, cross-checks, and
innovation.
What about higher redshift studies (see Ziegler and Crawford, this vol-
ume)? PFIS can sample a 0.6 arcsec slitwidth, but the sub-arcsec image quality
enjoyed regularly by VLT and Gemini is needed unless we remain content with
spatially-unresolved line-widths. Should it prove feasible to phase the SALT pri-
mary segements, we may consider pushing aggressively for significantly improved
image quality. This will have a major impact on both resolved, extragalactic as
well as high-resolution stellar spectroscopic observations. But SALT is merely
a 9m telescope; for extended sources observed at significant spectral resolution,
the area -- solid-angle product (AΩ) is what counts. As other facilities rush for the
highest possible angular resolution that, with today's detectors, may be better
suited for tomorrow's 30m-class telescopes, we should keep good aim on the type
of ground-breaking science enabled by our large field of view and the exquisite,
multi-faceted spectroscopic capabilities of PFIS.
Acknowledgments. This work is funded by NSF AST-0307417, NASA
LTSA NAG5-6032 and STScI/GO-9126.
References
Andersen, D. R. & Bershady, M. A. 2003, ApJ, 599, L79
Bell, E. & de Jong 2001, ApJ, 550, 212
Bershady, M. A., et al. 2003, ApJS, submitted
Bershady, M. A., et al. 2004, to appear in PASP
Bottema, R. 1997, A&A, 328, 517
Ciardullo, R., et al. 2003, submitted to ApJ
Debattista, V. P. & Williams, T. B. 2004, to appear in ApJ
Douglas, N. G. et al. 2002, PASP, 114, 1234
Gerssen, J., Kuijken, K., & Merrifield, M. R. 2000, MNRAS, 317, 545
Gerssen, J., Kuijken, K., & Merrifield, M. R. 2003, MNRAS, 345, 261
Kregel, M., van der Kruit, P. C., & de Grijs, R. 2002, MNRAS, 334, 646
Kuijken, K. & Gilmore, G. 1991, ApJ, 367, L9
Merrett, H. R., et al. 2003, MNRAS, 346, 62
van Albada, T. S. & Sancisis, R. 1986, Phil. Trans. R. Soc. Lon., 320, 447
Verheijen, M. A. W., et al. 2004, AN, 325, 151
Weiner, B. J., Sellwood, J. A. & Williams, T. B. 2001, ApJ, 546, 931
|
0812.0293 | 1 | 0812 | 2008-12-01T13:37:35 | The Gaia Era: synergy between space missions and ground based surveys | [
"astro-ph"
] | The Gaia mission is expected to provide highly accurate astrometric, photometric, and spectroscopic measurements for about $10^9$ objects. Automated classification of detected sources is a key part of the data processing. Here a few aspects of the Gaia classification process are presented. Information from other surveys at longer wavelengths, and from follow-up ground based observations will be complementary to Gaia data especially at faint magnitudes, and will offer a great opportunity to understand our Galaxy. | astro-ph | astro-ph | Mem. S.A.It. Vol. 75, 282
c(cid:13) SAIt 2008
Memorie della
The Gaia Ea: yegy bewee a
e ii
ad g d baed vey
A. Vallenari1, R. Sordo,1
INAF -- Osservatorio Astronomico, Vic. dell'Osservatorio 5 I-35122 Padova, Italy e-mail:
[email protected]
Abstract. The Gaia mission is expected to provide highly accurate astrometric, photomet-
ric, and spectroscopic measurements for about 109 objects. Automated classification of de-
tected sources is a key part of the data processing. Here a few aspects of the Gaia classifica-
tion process are presented. Information from other surveys at longer wavelengths, and from
follow-up ground based observations will be complementary to Gaia data especially at faint
magnitudes, and will offer a great opportunity to understand our Galaxy.
Key words. Stars: astrometry -- Galaxy: structure
1. Introduction
ESA's Gaia mission, to be launched in 2011,
is meant to obtain accurate position, parallax
and proper motion for 109 object all over the
sky, up to magnitude G=20 with an astromet-
ric accuracy at the µarcsec level. The low-
dispersion spectroscopy obtained in the BP and
RP passbands (330 -- 1000 nm, resolution of 3 --
30 nm) will be used not only to correct the as-
trometry for color effects, but also to obtain a
characterization of the sources themselves. The
Radial Velocity Spectrograph (RVS, 840 -- 890
nm, RP=11 500) will measure radial velocities,
with a precision of few km s−1, up to G=16.
Gaia will observe the whole sky for five years
(plus a possible year extension) achieving a
mean of ∼ 80 observations for each source.
The final catalog will be available in 2020, pre-
ceded by an early data release. The data reduc-
tion is a great challenge: the size of Gaia re-
lated data will be about 1015 bytes, while the
Send offprint requests to: A. Vallenari
final data delivered to the community would
be of about 20 TB. The estimated computa-
tion size will be of the order of 1021 Flops (see
Mignard et al 2008).
2. The classification of Gaia objects
Gaia does not use a full sky input Catalog.
However, in the early stages of the mission, the
Guide Star Catalog (Lasker et al 2008) (GSC-
II)will be used as inputs for the initial source
list to support the identification of the objects.
The GSC-II is constructed from the scanned
images of the Palomar and UK Schmidt photo-
graphic survey digitized at the Space Telescope
Science Institute. It makes use of Tycho-2
data as reference for the astrometric calibra-
tion. This is a good example of synergy be-
tween ground based surveys and spatial mis-
sions. Outside the Galactic plane the Catalog
is complete down to RF ∼ 20, but stellar clas-
sification is reliable at 90% level at RF ∼ 19.5.
Coordinates are provided with a mean preci-
8
0
0
2
c
e
D
1
]
h
p
-
o
r
t
s
a
[
1
v
3
9
2
0
.
2
1
8
0
:
v
i
X
r
a
Vallenari & Sordo: The Gaia Era
283
′′2 -- 0.
′′28 for about 9 × 108 objects.
sion of 0.
The final Gaia catalog is expected to provide
positions, parallaxes, proper motions, radial
velocities, photometry in the two broad bands
BP/RP, discrete classification of sources, astro-
physical parameters (APs) for single stars (i.e.
Teff, log g,...), and possibly the parametrization
of special sources (galaxies, QSOs). To avoid
biases, and to built a reference frame for as-
trometry, Gaia will re-classify the observed ob-
jects. Such a large amount of data can be clas-
sified only in an automated way. In the Gaia
project, the classification algorithms are based
on both supervised and unsupervised methods
(see Smith et al 2008), first producing a dis-
crete classification of the objects, i.e. divid-
ing objects having higher probability of being
stars, galaxies, and QSOs, then estimating its
APs by comparison with a set of templates.
Finally, the treatment of the outliers will relay
on unsupervised methods. The scientific com-
munity involved in Gaia is working to calculate
extensive libraries of synthetic and observa-
tional templates with improved physics for all
the classes of objects to be used as training data
for the classification algorithms. In the classifi-
cation task, Gaia data should be complemented
by external (i.e. non-Gaia) information. The
simplest way to do this is via astrometric cross
matching to existing catalogs. The most obvi-
ous candidates are the spectral and/or morpho-
logical classifications from SDSS and 2MASS,
later also UKIDSS and PS1. FIRST could also
be useful for the QSOs. This information could
easily be introduced in a multi-component dis-
crete classifier by means of the introduction of
priors as pre-data estimate that a object belongs
to a given class (see Bailer-Jones et al 2008,
for a detailed description of the method).
3. Training data for object
classification
The Gaia object classification includes as
well the determination of the APs of stars and
possibly galaxies. As we state in the previous
Section, supervised methods require the com-
parison with a set of templates, either observed
or synthetic, as training data sets. While
observational programs have already started
to built a homogeneous sample of stellar
templates, it is clear that training data cannot
be purely observational, since a large variety
and uniform coverage is requested for the
parameter distributions. It turns out that high
resolution and high quality synthetic libraries
are of fundamental importance. New extensive
calculations of sets of spectral stellar libraries
with improved physics are on the way. They
cover the two Gaia spectral ranges: 300 -- 1100
nm at 0.1 nm resolution, and 840 -- 890 nm at
0.001 nm resolution. These new libraries span
a large range in atmospheric parameters, from
super-metal-rich to very metal-poor stars, from
cool stars to hot, from dwarfs to giant stars,
with small steps in all parameters, typically
∆Teff =250 K (for cool stars), ∆ log g=0.5
dex, ∆[Fe/H]=0.5 dex. Depending on Teff,
these libraries rely on MARCS (F,G,K stars:
Gustafsson et al 2008), PHOENIX (cool and
C stars: Brott & Hauschildt 2005), KURUCZ,
TLUSTY models including magnetic field, pe-
culiar abundances, mass loss (A,B,Be,O stars:
Sordo & Munari 2006;
Bouret et al 2008;
Alvarez & Plez 1998; Kochukhov et al 2005).
included
WDA and WDB objects
(Castanheira et al 2006).
models
are based on different assumptions: KURUCZ
are LTE, plane-parallel, MARCS implement
also spherical symmetry while PHOENIX and
TLUSTY (hot stars) can calculate NLTE mod-
els both in plane-parallel mode and spherical
symmetry (see for a more detailed discussion
Gustafsson et al 2008; Sordo et al 2008). A
comparable effort is carried on in the galaxy
domain. We remind that Gaia will extend the
existing surveys of galaxies (see or instance
the SDSS covering only a fifth of the sky)
since it will be able to detect about 107 unre-
solved galaxies down to G=20 covering the
whole sky for the first time since photographic
surveys (UK, ESO, Palomar Schmidt) of 30
years ago in a larger spectral range. Large
synthetic libraries of galaxy spectra covering
the main Hubble types in the Gaia spectral
range at a sampling of 1 nm or less are under
construction (see Tsalmantza et al 2007). At
present, a library of about 3800 galaxy at zero
redshift, and a second one of about 140,000
spectra at changing redshift are available.
are
Those
284
Vallenari & Sordo: The Gaia Era
Finally, QSOs synthetic and semi-empirical
libraries (Claeskens et al 2006) are calculated.
4. QSO classification: Gaia reference
frame
A high precision reference frame in Gaia is ob-
viously mandatory to reach a high accuracy of
10 µas in the astrometry. With such a request,
the astrometry must be self-calibrating and for
this reason Gaia must observe a large num-
ber of quasars to define a high precision refer-
ence frame. This quasar sample must be very
clean, showing only a low contamination by
other objects. A probabilistic classifier is built
to select objects with higher probability. Care
is paid to construct a pure sample of objects,
discarding for instance QSOs with low equiv-
alent width emission lines which can be easily
confused with F-G-K stars (4000-8000 K) hav-
ing high extinction (AV ∼ 8 − 10). Preliminary
results show that a pure sample of QSOs at
65% level complete down to G=20 can be se-
lected. This is adeguate to establish a refer-
ence frame for Gaia: Gaia will observe 500 000
QSOs brighter than G=20, but only the ob-
jects with the most accurate positions (G <
18) will be used to built a reference frame.
Following our preliminary estimates, a sample
of 250 000 can be recovered with no more than
13 contaminats (see Bailer-Jones et al 2008). It
is clear that the Gaia reference frame needs
to be aligned with the International Celestial
Reference Frame (ICRF) with the highest ac-
curacy. The ICRF is based on VLBI positions
of about 700 extragalactic radio sources. For
this reason an observational program is ini-
tiated at the VLBI to identify suitable radio
sources to align the two reference systems. At
the moment, only a few objects can be use-
ful to this purpose, either because they are not
bright enough at optical wavelengths, or be-
cause they have extended emissions in the ra-
dio which precludes to reach the requested as-
trometric accuracy (Bourda et al 2008).
5. Gaia and complementary surveys
Gaia will be of fundamental importance to
study the Galactic structure at low latitudes: the
position and the velocity of the OB stars con
be measured without assuming rotation curve
or extinction. The distances of OB stars at 4
Kpc with Av= 4 mag extinction will be de-
rived with an accuracy of 13%, space veloci-
ties with an accuracy of a few Km/s. Fainter
stars can be measured as well, giving impor-
tant information about the mass distribution.
However it should be noticed that at faint mag-
nitudes (G ∼ 20), the expected accuracy is de-
graded and stellar APs cannot be reliably de-
termined. Once that a three-dimensional map
of the Galactic plane is derived, and that dis-
tances and kinematics are known for all the
star forming regions, we will be able to trace
the disk and spiral structure of the Galaxy.
The challenge will be to built dynamical the-
ories to reproduce the density fields and the
velocity fields. To distinguish between m=2
and m=4 spiral arm structure recent simula-
tions find out that the potential needs to be
known to 10%, the line-of-sight velocity ac-
curacy needs to be better than 20 Km/s, dis-
tances shoud be known with uncertainties bet-
ter than 30% (Minchev & Quillen 2008). This
is well within the possibility limits of the Gaia
survey. However, two main problems should
be reminded: the first is related to the dust ob-
scuration, which might hamper the determina-
tion of the star mass density, while the sec-
ond is due to the confusion of the stars in
the field of view (FOV). Concerning extinc-
tion, its knowledge will become a limiting fac-
tor in the determination of the stellar luminosi-
ties and APs. The estimate made on the basis
of G, BP, RP photometry only present some
degree of degeneration: it is difficult to de-
rive both the extinction and the extinction law
for late type giants. Using RVS information,
both parameters can be derived. In addition,
the use diffuse interstellar bands (DIBs) are ex-
tinction tracers will be explored. The strongest
expected DIB in the surveyed range is at 862
nm. Its equivalent width well correlates with
the interstellar reddening (Munari et al 2008).
A large effort in going on in the Gaia commu-
nity to ensure a proper determination of the ex-
tinction, testing and comparing different meth-
ods, including the use of infrared passbands in
combination with Gaia passbands (Knude &
Vallenari & Sordo: The Gaia Era
285
Lindstroem 2007). In addition, Gaia is confu-
sion limited (BP/RP images of different stars
are superposed on the FOV) when the total
star density per transit (sum of the star number
of both FOVs) is higher than 750,000 stars/sq
deg. This means that the central inner degrees
will not be well measured. Bright stars can
probably still be recovered, but the precision
will decrease. Simulations are still ongoing
(Marrese 2008). Infrared surveys are comple-
mentary to Gaia to understand the structure
of the inner disk and deal with dust obscura-
tion. Current astrometric data in the infrared
are of poor quality, even if great improvements
were made in the recent past (see 2MASS and
DENIS). The UKIDSS will cover only a part
of the sky (7000 sq deg), but it will be much
deeper (K∼ 18 − 19 on the Galactic plane,
and K∼ 21 at higher latitudes over 35 sq deg).
Interferometric and adaptive optics astrometry
can be very promising, but they can only pro-
vide high precision relative astrometry in small
fields. However, absolute parallaxes in small
fields can be derived with high precision from
these observations when enough suitable extra-
galactic reference sources are detected in the
field of view. Pan STARRS and LSST optical-
near infrared surveys covering Northern and
Southern sky respectively, are expected to ob-
serve 1010 stars down to magnitudes brighter
than 24 mag. reaching an accuracy of about
3-25, 1-10 mas, respectively on the parallax
determination. One of the space infrared sur-
vey which is foreseen in the near future is
JASMINE, which is expected to cover the
Bulge and the inner disk regions. JASMINE
however, will not go very deep (z< 14) ob-
serving about 107 stars with a precision of 0.01
mas on the parallaxes. All those surveys will
be of great importance to map highly obscured
regions where Gaia is not efficient, although
they cannot reach the same accuracy on as-
trometry. Finally, since Gaia radial velocities
will be measured only for G < 16, spectro-
scopic ground based follow-up with 8m class
telescopes need to be planned.
In conclusion, the Gaia mission will pro-
vide highly accurate astrometric, photometric,
and spectroscopic measurements for a large
sample of objects. The high quality of Gaia
data, especially on astrometry will not be
reached by any of the planned surveys in the
near future. Gaia data complemented by infor-
mation coming from surveys at longer wave-
lengths, and from follow-up ground based ob-
servations will offer a great opportunity to un-
veil the formation and the structure Galaxy.
Acknowledgements. This short review relies on the
work of many members of the Gaia community, who
are gratefully acknowledged for their valuable con-
tribution to the project.
References
Alvarez R. & Plez B., 1998, A&A 330, 1109
Bailer-Jones, C. A. L. et al 2008, MNRAS,
in press
Bourda G., Charlot, P., Le Champion, J-F,
2008, A&A submitted
Bouret, J.-C. et al. 2008, in preparation
Brott, I., & Hauschildt, P. H. 2005, The
Three-Dimensional Universe with Gaia, 576,
565
Castanheira, B. G., et al 2006, A&A, 450,
331
Claeskens, J.-F., et al 2006, MNRAS, 367,
879
Gustafsson, B., et al. 2008, ArXiv e-prints,
805, arXiv:0805.0554
Knude J. & Lindstroem 2007, GAIA-C8-TN-
NBI-JK-002
Kochukhov, O., Khan, S., & Shulyak, D.
2005, A&A, 433, 671
Lasker, B.M., Lattanzi, M., et al 2008, AJ
136,735
Marrese, P, 2008, GAIA-C5-TN-LEI-PM-
003
Minchev & Quillen 2008, MNRAS 386,1579
Mignard, F., et al. 2008, in IAUS 248, p.224
Munari U., Tommasella, L.,Fiorucci, L., et al.
2008, A&A 488, 969
Smith K. et al 2008, ASP Conference Series,
Vol. 394, p.539
Sordo, R., & Munari, U. 2006, A&A, 452,
735
Sordo R., Vallenari A., et al in Mem.Sait 2008
in press
Tsalmantza, P., et al. 2007, A&A, 470, 761
|
astro-ph/0107153 | 2 | 0107 | 2002-04-16T15:36:41 | Implications of O and Mg abundances in metal-poor halo stars for stellar iron yields | [
"astro-ph"
] | Inhomogeneous chemical evolution models of galaxies which try to reproduce the scatter seen in element-to-iron ratios of metal-poor halo stars are heavily dependent on theoretical nucleosynthesis yields of core-collapse supernovae. Hence inhomogeneous models present themselves as a test for stellar nucleosyn- thesis calculations. Applying an inhomogeneous chemical evolution model to our Galaxy reveals a number of shortcomings of existing theoretical nucleosynthesis yields. One problem is the predicted scatter in [O/Fe] and [Mg/Fe] which is too large compared to the one observed in metal-poor halo stars. This can be either due to the O or Mg yields or due to the Fe yields (or both). However, O and Mg are alpha-elements that are produced mainly during hydrostatic burning and thus are not affected by the theoretical uncertainties afflicting the collapse and explosion of a massive star. Stellar iron yields, on the other hand, depend heavily on the choice of the mass-cut between ejecta and proto neutron star and are therefore very uncertain. We present Fe yield distributions as function of progenitor mass that are consistent with the abundance distribution of metal- poor halo stars and are in agreement with observed Ni yields of SNe II with known progenitor masses. The iron yields of lower-mass SNe II (in the range 10-20 Msol) are well constrained by those observations. Present observations, however, do not allow to determine a unique solution for higher-mass SNe. Nevertheless, the main dependence of the stellar iron yield as function of progenitor mass may be derived and can be used as constraint for future supernova/hypernova models. The results are of importance for the earliest stages of galaxy formation when the ISM is dominated by chemical inhomogeneities and the instantaneous mixing approximation is not valid. | astro-ph | astro-ph | Astronomy & Astrophysics manuscript no.
(will be inserted by hand later)
Implications of O and Mg abundances in metal-poor halo stars
for stellar iron yields
2
0
0
2
r
p
A
6
1
2
v
3
5
1
7
0
1
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
D. Argast1
,
2, M. Samland1, F.-K. Thielemann2, and O. E. Gerhard1
1 Astronomisches Institut der Universitat Basel, Venusstrasse 7, CH-4102 Binningen, Switzerland
2 Institut fur Physik der Universitat Basel, Klingelbergstrasse 82, CH-4056 Basel, Switzerland
Received . . . / Accepted . . .
Abstract. Inhomogeneous chemical evolution models of galaxies which try to reproduce the scatter seen in element-
to-iron ratios of metal-poor halo stars are heavily dependent on theoretical nucleosynthesis yields of core-collapse
supernovae (SNe II). Hence inhomogeneous chemical evolution models present themselves as a test for stellar
nucleosynthesis calculations. Applying such a model to our Galaxy reveals a number of shortcomings of existing
nucleosynthesis yields. One problem is the predicted scatter in [O/Fe] and [Mg/Fe] which is too large compared
to the one observed in metal-poor halo stars. This can be either due to the oxygen or magnesium yields or due
to the iron yields (or both). However, oxygen and magnesium are α-elements that are produced mainly during
hydrostatic burning and thus are not affected by the theoretical uncertainties afflicting the collapse and explosion
of a massive star. Stellar iron yields, on the other hand, depend heavily on the choice of the mass-cut between
ejecta and proto-neutron star and are therefore very uncertain. We present iron yield distributions as function of
progenitor mass that are consistent with the abundance distribution of metal-poor halo stars and are in agreement
with observed 56Ni yields of core-collapse supernovae with known progenitor masses. The iron yields of lower-mass
SNe II (in the range 10 − 20 M⊙) are well constrained by these observations. Present observations, however, do
not allow us to determine a unique solution for higher-mass SNe. Nevertheless, the main dependence of the stellar
iron yields as function of progenitor mass can be derived and may be used as a constraint for future core-collapse
supernova/hypernova models. A prediction of hypernova models is the existence of ultra α-element enhanced
stars at metallicities [Fe/H] ≤ −2.5, which can be tested by future observations. The results are of importance
for the earliest stages of galaxy formation when the ISM is dominated by local chemical inhomogeneities and the
instantaneous mixing approximation is not valid.
Key words. Physical processes: nucleosynthesis -- Stars: abundances -- ISM: abundances -- Galaxy: abundances --
Galaxy: evolution -- Galaxy: halo
1. Introduction
The key to the formation and evolution of the Galaxy
lies buried in the kinematic properties and the chemical
composition of its stars. Especially old, metal-poor halo
stars and globular clusters are ideal tracers of the forma-
tion process. Although many of the properties of the halo
component and its substructures have been unveiled, it is
still not possible to decide whether the Galaxy formed by a
fast monolithic collapse (Eggen, Lynden-Bell & Sandage,
1962), by the slower merging and accretion of subgalac-
tic fragments (Searle & Zinn 1978) or within the context
of a hybrid picture, combining aspects of both scenarios.
Recently, Chiba & Beers (2000) made an extensive inves-
tigation to address this question, concluding that a hybrid
scenario, where the inner part of the halo formed by a fast,
Send offprint requests to: D. Argast,
e-mail: [email protected]
dissipative collapse and the outer halo is made up of the
remnants of accreted subgalactic fragments, best explains
the observational data. It also seems to be consistent with
the theory of galaxy formation based on cold dark matter
scenarios (see e.g. Steinmetz & Muller 1995; Gnedin 1996;
Klypin et al. 1999; Moore et al. 1999; Pearce et al. 1999;
Bekki & Chiba 2000; Navarro & Steinmetz 2000).
However, the kinematic structure of the halo alone is
not sufficient to draw a conclusive picture of the formation
of the Galaxy. Old, unevolved metal-poor halo stars allow
us to probe the chemical composition and (in)homogeneity
of the early interstellar medium (ISM) and its evolution
with time, since element abundances in the stellar atmo-
spheres of those stars directly reflect the chemical com-
position of the material out of which they formed. It is
almost impossible to determine the age of single stars (ex-
cept in a few cases where radioactive thorium or uranium
was detected, see e.g. Cayrel et al. 2001). Therefore, the
2
D. Argast et al.: Implications of O and Mg abundances for stellar iron yields
metallicity Z or iron abundance [Fe/H] of a star is taken
as an age estimate, knowing that an age -- metallicity rela-
tion can only be used in a statistical sense for the bulk of
stars (see e.g. Argast et al. 2000, hereafter Paper I).
Common chemical evolution models mostly assume
that the metal-rich ejecta of supernovae (SNe) are
mixed instantaneously and homogeneously into the ISM.
Models using this approximation, together with theoreti-
cal nucleosynthesis yields of type Ia and type II SNe, can
explain the behaviour of element-to-iron ratios ([el/Fe])
of stars as function of metallicity [Fe/H] for many ele-
ments and for [Fe/H] ≥ −2. This shows that the instan-
taneous mixing approximation is valid at this stage and --
since at these metallicities even some of the lowest mass
core-collapse supernovae (SNe II) have exploded -- that
the stellar yields averaged over the initial mass function
(IMF) are for most elements accurate within a factor of
two (see e.g. Samland 1997).
However, observations of very metal-poor stars show
significant scatter in [el/Fe] ratios at [Fe/H] < −2, im-
plying that the ISM was not well mixed at this stage
(Paper I). These local chemical
inhomogeneities were
probably mainly caused by SNe II, since progenitors of
SN Ia have much longer lifetimes and are unimportant for
the chemical enrichment of the ISM until approximately
[Fe/H] ≥ −1. At these early stages of galaxy formation,
the instantaneous mixing approximation is not valid and
yields depending on the mass of individual SNe II become
important. Therefore, accurate nucleosynthesis yields as
a function of progenitor mass are crucial for the under-
standing of the earliest stages of galaxy formation.
In Paper I, a stochastic chemical evolution model
was presented which accounts for local chemical inho-
mogeneities caused by SNe II with different progenitor
masses. The model successfully reproduces the scatter in
[el/Fe] ratios as function of [Fe/H] for some elements like Si
or Ca, but fails quantitatively in the case of the two most
abundant α-elements, O and Mg. The scatter in [O/Fe]
and [Mg/Fe] is much larger than observed and predicts
stars with [O/Fe] and [Mg/Fe] ≤ −1.0. This result de-
pends mainly on the employed stellar yields, demonstrat-
ing that either the oxygen/magnesium or the iron yields
(or both) as a function of progenitor mass are not well
determined by existing nucleosynthesis models.
The solution to this problem is important for the un-
derstanding of the chemical evolution of our Galaxy. In
this work, we try to reconcile element abundance observa-
tions of metal-poor halo stars with the predictions of our
inhomogeneous chemical evolution model by changing the
progenitor mass dependence of stellar yields. The forma-
tion of oxygen and magnesium in hydrostatic burning and
ejection during a SN event is much better understood than
the formation and ejection of 56Ni (which decays to 56Fe
and forms the bulk of the ejected iron), since the amount
of ejected 56Ni is directly linked to the still not fully under-
stood explosion mechanism (c.f. Liebendorfer et al. 2001;
Mezzacappa et al. 2001; Rampp & Janka 2000). Any at-
tempt to alter stellar yields should therefore start with
iron and iron-group elements. We present a method to de-
rive stellar iron yields as function of progenitor mass from
the observations of metal-poor halo stars, assuming given
yields of oxygen and magnesium.
In Sect. 2 we give a short description of the stochas-
tic chemical evolution model, followed by a summary of
observations and basic model results in Sect. 3. The dis-
cussion of uncertainties in stellar yields and how global
constraints on stellar iron yields can be gained from ob-
servations is given in Sect. 4. Implications for stellar iron
yields and conclusions are given in Sect. 5 and Sect. 6,
respectively.
2. The chemical evolution model
Observations of very metal-poor halo stars show a scatter
in [el/Fe] ratios of order 1 dex. This scatter gradually de-
creases at higher metallicities until a mean element abun-
dance is reached which corresponds to the [el/Fe] ratio of
the stellar yields integrated over the initial mass function.
Our stochastic chemical evolution model of Paper I follows
the enrichment history of the halo ISM in a cube with a
volume of (2.5 kpc)3, down to a resolution of (50 pc)3.
Every cell of the grid contains detailed information about
the enclosed ISM and the mass distribution of stars. For
the purpose of this paper, the enrichment of the ISM with
O, Mg, Si, Ca and Fe is computed.
At every time-step, randomly chosen cells may create
stars. The likelihood for a cell to form a star is propor-
tional to the square of the local ISM density. The mass of
a newly formed star is chosen randomly, with the con-
dition that the mass distribution of all stars follows a
Salpeter IMF. The lower and upper mass limits for stars
are taken to be 0.1 M⊙ and 50 M⊙, respectively. Newly
born stars inherit the abundance pattern of the ISM out of
which they formed, carrying therefore information about
the chemical composition of the ISM at the place and time
of their birth.
Stars in a range of 10 − 50 M⊙ are assumed to explode
as SNe II (or hypernovae, we will use the term SNe II to in-
clude hypernovae unless otherwise noted) resulting in an
enrichment of the neighbouring ISM. Intermediate mass
stars form planetary nebulae, which return only slightly
enriched material. Low mass stars do not evolve signifi-
cantly during the considered time but serve to lock up part
of the mass, affecting therefore the abundances of elements
with respect to hydrogen. Stellar yields are taken from
Thielemann et al. (1996, hereafter TH96) and Nomoto et
al. (1997). Additionally, since there are no nucleosynthesis
calculations for 10 M⊙ progenitors, their yields were set
to 1/10 of the yields of the 13 M⊙ model. We then linearly
interpolated the stellar yields given in these papers, since
we use a finer mass-grid in our simulation. The interpo-
lation gives IMF averaged values of [el/Fe] ratios which
are in good agreement (within 0.1 dex) with the observed
mean values of metal-poor stars.
The SN remnant sweeps up the enriched material in
a spherical, chemically well mixed shell. Since the explo-
D. Argast et al.: Implications of O and Mg abundances for stellar iron yields
3
Fig. 1. [O/Fe] and [Mg/Fe] ratios vs. metallicity [Fe/H] of metal-poor halo stars (squares and triangles) and model
stars (dots). Circles depict [O/Fe] and [Mg/Fe] ratios of SN II models of the given progenitor mass. (See text for
details.) In contrast to observations, model stars with subsolar [O/Fe] and [Mg/Fe] ratios are predicted by the applied
stellar yields, as visible by the location of the 13 and 15 M⊙ SNe.
sion energy of SNe II is believed to depend only slightly
on the mass of its progenitor (Woosley & Weaver 1995,
hereafter WW95; Thielemann et al. 1996), we assume that
each SN II sweeps up about 5×104 M⊙ of gas (Ryan et al.
1996; Shigeyama & Tsujimoto 1998). Stars which form out
of material enriched by a single SN inherit its abundance
ratios and therefore show an element abundance pattern
which is characteristic for this particular progenitor mass.
This will lead to a large scatter in the [el/Fe] ratios, as
long as local inhomogeneities caused by SN events dom-
inate the halo ISM. As time progresses, supernova rem-
nants overlap and the abundance pattern in each cell ap-
proaches the IMF average, leading to a decrease in the
[el/Fe] scatter at later times. Since the SN remnant for-
mation is the only dynamical process taken into account,
this model shows the least possible mixing efficiency for
the halo ISM. This is just the opposite to chemical evolu-
tion models which use the instantaneous mixing approxi-
mation. We continue our calculation up to an average iron
abundance of [Fe/H] = −1.0. At this metallicity, SN Ia
events which are not included in our model start to influ-
ence the ISM significantly. A more detailed description of
the model can be found in Paper I.
We emphasize one important result: Starting with a
primordial ISM and taking into account local inhomo-
geneities caused by SNe II, the initial scatter in [el/Fe]
ratios is determined solely by the adopted nucleosynthesis
yields. The details of the chemical evolution model only
determine how fast a chemically homogeneous ISM is
reached, i.e. how the scatter evolves with time or (equiv-
alently) iron abundance [Fe/H]. Therefore, the range of
[el/Fe] ratios of the most metal-poor stars does not de-
pend on specific model parameters but is already fixed by
the stellar yields.
3. Observations and basic model results
As mentioned in the introduction, existing nucleosynthesis
models, combined with a chemical evolution model tak-
ing local inhomogeneities into account, predict [O/Fe] and
[Mg/Fe] ratios less than solar for some metal-poor stars.
This is in contrast to observations of metal-poor halo
stars, as can be seen in Fig. 1. The left hand panel shows
the [O/Fe] ratio of observed and model stars as func-
tion of iron abundance [Fe/H] and the right hand panel
the same for [Mg/Fe], where the model stars are plot-
ted as small dots. The observational data were collected
from Magain (1989), Molaro & Bonifacio (1990), Molaro
& Castelli (1990), Peterson et al. (1990), Bessell et al.
(1991), Ryan et al. (1991), Spiesman & Wallerstein (1991),
Spite & Spite (1991), Norris et al. (1993), Beveridge &
Sneden (1994), King (1994), Nissen et al. (1994), Primas
et al. (1994), Sneden et al. (1994), Fuhrmann et al. (1995),
McWilliam et al. (1995), Balachandran & Carney (1996),
Ryan et al. (1996), Israelian et al. (1998), Jehin et al.
(1999), Boesgaard et al. (1999), Idiart & Th´evenin (2000),
Carretta et al. (2000) and Israelian et al. (2001).
Combining data from various sources is dangerous at
best, since different investigators use different methods to
derive element abundances with possibly different and un-
4
D. Argast et al.: Implications of O and Mg abundances for stellar iron yields
known systematic errors. This influences the scatter in
[el/Fe] ratios, which plays a crucial role in determining
the chemical (in)homogeneity of the ISM as function of
[Fe/H]. Unfortunately, there is no investigation with a
sample of oxygen/magnesium abundances of metal-poor
halo stars that is large enough for our purpose. Therefore,
we are forced to combine different data sets, keeping in
mind that unknown systematic errors can enlarge the in-
trinsic scatter in element abundances of metal-poor stars.
Recently, Idiart & Th´evenin (2000) and Carretta et al.
(2000) reanalyzed data previously gathered by other au-
thors and applied NLTE corrections to O, Mg and Ca
abundances, which is a first step in reducing the scatter
introduced by systematic errors. Therefore we divided the
collected data into two groups, namely the data of Idiart &
Th´evenin (2000) and Carretta et al. (2000), which is rep-
resented in Fig. 1 by triangles, and the data of all other
investigators, represented by squares. If multiple observa-
tions of a single star exist, abundances are averaged and
pentagons and diamonds are used for the first and sec-
ond group, respectively. (Averaging of data points was
only necessary in a few cases for Mg, Si and Ca abun-
dances.) Note, that the average random error in element
abundances is of the order 0.1 dex.
Also plotted in Fig. 1 as circles are [el/Fe] ratios pre-
dicted by nucleosynthesis calculations of TH96. The num-
bers in the circles give the mass of the progenitor star in
solar masses. In the picture of inhomogeneous chemical
evolution, a single SN event enriches the primordial ISM
locally (in our model by mixing with 5 × 104 M⊙ of ISM)
with its nucleosynthesis products. Depending on the mass
of the progenitor star, the resulting [O/Fe] and [Mg/Fe]
ratios in these isolated patches of ISM cover a range of over
two dex and as long as the ISM is dominated by these local
inhomogeneities, newly formed stars will show the same
range in their [el/Fe] ratios. In particular, this means that
stars with [O/Fe] and [Mg/Fe] as small as −1.0 are in-
evitably produced by our model. This is in contrast to the
bulk of observed metal-poor halo stars, which show [O/Fe]
and [Mg/Fe] ratios in the range between 0.0 and 1.2, and
is a strong indication that existing nucleosynthesis models
may correctly account for IMF averaged abundances but
fail to reproduce stellar yields as function of progenitor
mass.
4. Global constraints on stellar Fe yields
4.1. Uncertainties in O, Mg and Fe yields
Apart from the shortcomings of nucleosynthesis yields dis-
cussed in Sect. 3, there seems to be an additional uncer-
tainty concerning either the stellar yields of O and Mg
or the derivation of their abundances in metal-poor halo
stars, as shown in Fig. 2:
The theoretical nucleosynthesis yields of oxygen
(YO (m)) and magnesium (YMg (m)) show a very similar
dependence on progenitor mass m, i.e. in first order we
can write YMg (m) ≈ 6.7 · 10−2 · YO (m). Thus, for model
Fig. 2. [O/Mg] vs. [Mg/H] ratios of metal-poor halo stars.
Nucleosynthesis models predict a narrow region of possible
[O/Mg] ratios (hatched) which is not consistent with the
scatter of observations. Symbols are as in Fig. 1.
stars [O/Mg] ≈ 0.0 on average, and due to chemical inho-
mogeneities in the early ISM, model stars scatter in the
range −0.3 ≤ [O/Mg] ≤ 0.1 (hatched region in Fig. 2). In
contrast to theoretical predictions, observations of metal-
poor halo stars scatter in the range −0.3 ≤ [O/Mg] ≤ 0.5,
with a mean of [O/Mg] ≈ 0.15. This result is very impor-
tant, since it means that either even our understanding of
nucleosynthesis processes during hydrostatic burning is in-
complete or that oxygen abundances at very low metallic-
ities tend to be overestimated (or magnesium abundances
underestimated).
The problem hinted at in Fig. 2 is also connected to
the recent finding that the mean [O/Fe] ratio of metal-
poor halo stars seems to increase with decreasing metallic-
ity [Fe/H], whereas the mean [Mg/Fe] ratio seems to stay
constant (see e.g. Israelian et al. 1998, 2001; Boesgaard
et al. 1999; King 2000; but see also Rebolo et al. 2002).
This result can not be explained by changes in the sur-
face abundances due to rotation, since rotation tends to
decrease the oxygen abundance in the stellar atmosphere,
whereas magnesium abundances remain unaffected (Heger
& Langer 2000; Meynet & Maeder 2000). However, the
problem described with Fig. 2 would disappear, if the in-
crease in [O/Fe] with decreasing metallicity is not real but
due to some hidden systematic error. Then oxygen abun-
dances would have to be reduced, resulting in a smaller
scatter and lower mean in [O/Mg].
Regarding nucleosynthesis products, a crude argument
shows that (at least in the non-rotating case) we should
not expect a drastic change in the progenitor mass depen-
dence of O and Mg yields: Oxygen and magnesium are pro-
D. Argast et al.: Implications of O and Mg abundances for stellar iron yields
5
duced mainly during hydrostatic burning in the SN pro-
genitor and only a small fraction of the ejecta stems from
explosive neon- and carbon-burning (see e.g. Thielemann
et al. 1990, 1996). Magnesium is to first order a prod-
uct of hydrostatic carbon- and ensuing neon-burning in
massive stars. The amount of freshly synthesized magne-
sium depends on the available fuel, i.e. the size of the C-O
core after hydrostatic He burning, which also determines
the amount of oxygen that gets expelled in the SN event.
Thus, O and Mg yields as function of progenitor mass
should be roughly proportional to each other. A very large
mass loss during hydrostatic carbon burning could reduce
the size of the C-O core and thus decrease the amount of
synthesized magnesium for a given progenitor mass. This
would result in a larger scatter of [O/Mg] ratios than in-
dicated by the hatched region in Fig. 2. But the evolu-
tionary timescale of carbon burning is very short indeed
(≈ 5.8 · 103 yr for a 25 M⊙ star, e.g. Imbriani et al. 2001),
making a significant change in the structure of the C-O
core unlikely.
Although the hydrostatic burning phases are thought
to be well understood, one has to keep in mind that the
important (effective) 12C(α, γ)16O reaction rate is still un-
certain and that the treatment of rotation and convection
may also influence the amount of oxygen and magnesium
produced during hydrostatic burning. Recently, Heger et
al. (2000) showed that even in the case of slow rotation
important changes in the internal structure of a massive
star occur:
Rotationally induced mixing is important prior to cen-
tral He ignition. After central He ignition, the timescales
for rotationally induced mixing become too large com-
pared to the evolutionary timescales, and the further evo-
lution of the star is similar to the non-rotating case.
Nevertheless, He cores of rotating stars are more massive,
corresponding to He cores of non-rotating stars with about
25% larger initial mass. Furthermore, for a given mass of
the He core, the C-O cores of rotating stars are larger than
in the non-rotating case. At the end of central He burn-
ing, fresh He is mixed into the convective core, converting
carbon into oxygen. Therefore, the carbon abundance in
the core is decreased, whereas the oxygen abundance is in-
creased. Unfortunately, no detailed nucleosynthesis yields
including rotation have been published yet, but since the
size of the He core is increased in rotating stars, at least
changes in the yields of α-elements have to be expected.
(For a review of the changes in the stellar parameters in-
duced by rotation see Maeder & Meynet 2000.)
Contrary to oxygen and magnesium which stem from
hydrostatic burning, iron-peak nuclei are a product of ex-
plosive silicon-burning. Unfortunately, no self-consistent
models following the main-sequence evolution, collapse
and explosion of a massive star exist to date which would
allow to determine reliably the explosion energy and the
location of the mass-cut between the forming neutron star
and the ejecta (Liebendorfer et al. 2001; Mezzacappa et al.
2001; Rampp & Janka 2000). Therefore, nucleosynthesis
models treat the mass cut usually as one of several free pa-
rameters and the choice of its value can heavily influence
the abundance of ejected iron-group nuclei. For this rea-
son, we feel that oxygen and magnesium yields of nucleo-
synthesis models are more reliable than iron yields, in spite
of the uncertainties discussed above.
To illustrate this point we show a comparison of O
and Mg yields (YO (m), YMg (m), Fig. 3) and of Fe yields
(YFe (m), Fig. 4) of nucleosynthesis calculations (neglect-
ing rotation) from different authors. The models of WW95
(solar composition "C" models) are marked by filled
squares, TH96 by filled circles, Nakamura et al. (2001, 1051
erg models) by open squares and Rauscher et al. (2002,
"S" models) by open triangles. Upper points in Fig. 3 cor-
respond to O yields, lower points to Mg yields. Apart from
the dip visible in YMg (m) of the WW95 models, the O and
Mg yields of the different authors agree remarkably well.
Chemical evolution calculations by Thomas et al. (1998)
show that WW95 underestimate the average Mg yield due
to this dip. The minor differences between the models
can mostly be attributed to different progenitor models
prior to core-collapse, the employed 12C(α, γ)16O reaction
rate, the applied convection criterion (e.g. Schwarzschild
vs. Ledoux) and artificial explosion methods after core-
collapse (e.g. piston vs. artificially induced shock wave).
On the other hand, as visible in Fig. 4, YFe (m) of the
different authors differs by more than an order of magni-
tude for certain progenitor masses, which is mostly due to
the arbitrary placement of the "mass-cut" between proto-
neutron star and ejecta. In order to reconcile the results
of the inhomogeneous chemical evolution model with ob-
served [O/Fe] and [Mg/Fe] abundance ratios, it is there-
fore clearly preferable to artificially adjust YFe (m) rather
than YO (m) and YMg (m).
For the following discussion, reliable O and Mg yields
as function of progenitor mass with an estimate of their
error range are needed. To this end, we calculated best fit
curves to the YO (m) and YMg (m) yields of the different
authors, visible as dashed lines in Fig. 3. The low Mg yields
of the 20 and 22 M⊙ WW95 models were neglected for
this purpose. The deviations ∆ǫ (m) of the original O and
Mg yields Yel (m) from our best fit yields Y el (m) is defined
as
∆ǫ (m) =
Yel (m) − Y el (m)
Y el (m)
.
(1)
The error ∆ǫ (m) depends on progenitor mass, but
is generally small. To account for the uncertainty in
Yel (m) introduced by the different nucleosynthesis mod-
els, we replace in the following the original Yel (m) by
(1 ± ∆ǫ) · Y el (m), where we dropped the dependence of
∆ǫ (m) on m in the notation. For most progenitor masses,
∆ǫ ≤ 0.2 for both O and Mg and the maximal deviation is
in both cases smaller than 0.5. The dotted lines in Fig. 3
show the curves (1 ± ∆ǫ) · Y el (m) with ∆ǫ = 0.2. Since
∆ǫ is small, the impact of the uncertainties in the stellar
O and Mg yields is almost negligible for the derivation
of constraints on YFe (m). On the other hand, these small
6
D. Argast et al.: Implications of O and Mg abundances for stellar iron yields
Table 1. Best fit O and Mg yields Y el (m) proposed in
Sect. 4.1. The first column gives the progenitor mass m
and the following columns the oxygen and magnesium
yields (all values in solar masses).
m Y O (m)
2.2E-02
10
8.6E-02
11
12
1.5E-01
3.1E-01
13
4.9E-01
14
6.7E-01
15
16
8.7E-01
1.1E+00
17
1.3E+00
18
1.5E+00
19
1.7E+00
20
21
2.0E+00
2.2E+00
22
2.5E+00
23
2.8E+00
24
25
3.0E+00
3.4E+00
26
3.7E+00
27
4.0E+00
28
4.3E+00
29
30
4.6E+00
Y Mg (m) m Y O (m)
5.0E+00
1.2E-03
4.9E-03
5.4E+00
5.7E+00
8.5E-03
2.0E-02
6.1E+00
3.6E-02
6.5E+00
5.3E-02
6.9E+00
7.3E+00
6.9E-02
8.6E-02
7.7E+00
1.0E-01
8.1E+00
1.2E-01
8.6E+00
9.0E+00
1.4E-01
1.5E-01
9.5E+00
1.7E-01
1.0E+01
1.9E-01
1.0E+01
1.1E+01
2.0E-01
2.2E-01
1.1E+01
2.4E-01
1.2E+01
2.6E-01
1.2E+01
1.3E+01
2.7E-01
2.9E-01
1.3E+01
3.1E-01
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
Y Mg (m)
3.3E-01
3.5E-01
3.6E-01
3.8E-01
4.0E-01
4.2E-01
4.4E-01
4.5E-01
4.7E-01
4.9E-01
5.2E-01
5.4E-01
5.6E-01
5.9E-01
6.1E-01
6.4E-01
6.6E-01
6.8E-01
7.1E-01
7.3E-01
uncertainties may (almost) be able to explain the discrep-
ancy in the scatter of [O/Mg] ratios between observations
and model stars visible in Fig. 2. Allowing for a mean devi-
ation of ∆ǫ = 0.2, the maximal scatter in [O/Mg] over all
progenitor masses may extend to −0.25 ≤ [O/Mg] ≤ 0.3,
which is very close to the one observed. Future nucleo-
synthesis calculations have to show whether this interpre-
tation is correct or not.
For the remaining discussion, we adopt the term
(1 ± ∆ǫ) · Y el (m) with ∆ǫ = 0.2 for the stellar oxygen
and magnesium yields as a function of progenitor mass,
assuming that they reproduce the true production in mas-
sive stars well enough. The values adopted for the best fit
yields Y el (m) are given in Table 1.
4.2. The influence of Z and SNe from Population III
stars
Apart from the uncertainties in the O and Mg yields dis-
cussed in Sect. 4.1, nucleosynthesis yields may also depend
on the metallicity of the progenitor. Unfortunately, the
question how important metallicity effects are is far from
solved. Nucleosynthesis calculations in general neglect ef-
fects of mass loss due to stellar winds. WW95 present
nucleosynthesis results for a grid of metallicities from
metal-free to solar and predict a decrease in the ejected
O and Mg mass with decreasing metallicity. However, the
O yields presented lie all in the range covered by the best
fit yields (1 ± ∆ǫ) · Y el (m) with ∆ǫ = 0.2 adopted for
this paper. This is not true in the case of Mg. But since
the "dip" in YMg (m) visible in Fig. 3 gets more and more
Fig. 3. O and Mg yields from different authors as func-
tion of progenitor mass. Models are from: WW95, filled
squares; TH96, filled circles; Nakamura et al. (2001), open
squares; Rauscher et al. (2002), open triangles. Upper
points correspond to O yields, lower points to Mg yields.
Dashed and dotted lines represent best fit curves to the
different nucleosynthesis models (see text for details).
Fig. 4. Fe yields from different authors as function of pro-
genitor mass. Symbols are the same as in Fig. 3. Contrary
to O and Mg yields, different authors obtain very different
yields for a given progenitor mass. This is mostly due to
the arbitrary placement of the mass cut.
D. Argast et al.: Implications of O and Mg abundances for stellar iron yields
7
pronounced with lower Z and since it is known (Thomas
et al., 1998), that WW95 underestimate the average pro-
duction of Mg, we feel that the metallicity dependence of
Mg yields is not established well enough to include this
feature into our analysis.
Contrary to the results obtained by WW95, Maeder
(1992) showed that stellar O yields decrease with increas-
ing metallicity due to strong mass loss in stellar winds.
(No detailed results were given in the case of magne-
sium.) Stars more massive than 25M⊙ with solar metal-
licity (Z=0.02), eject large amounts of He and C in stel-
lar winds (prior to the conversion into oxygen) which re-
sults in dramatically reduced O yields. Metal-poor stars
(Z≤0.001) do not undergo an extended mass-loss phase
and their O yields are comparable to the ones given by
WW95 and TH96. Since Z≤0.001 roughly corresponds to
[Fe/H] ≤ −1.5 and we are mainly concerned with metal-
poor halo stars in this metallicity range, we can neglect
changes in O yields due to metallicity.
Recently, Heger & Woosley (2000) published nucleo-
synthesis calculations of pair-instability SNe from very
massive, metal-free (Population III) stars in the mass
range from 140M⊙ to 260M⊙. For Population III stars in
the mass range 25 − 140M⊙ and stars more massive than
260M⊙, black hole formation without ejection of nucleo-
synthesis yields seems likely (Heger & Woosley 2000). In
order to investigate the influence of those massive metal-
free stars on the enrichment of the ISM and especially
their impact on the distribution of model stars in [O/Mg]
vs. [Mg/H] plots (c.f. Sect. 4.1), we carried out several in-
homogeneous chemical evolution calculations with vary-
ing SF efficiencies and IMF shapes for the Population III
stars. The detailed results will not be shown here, but
some basic conclusions are discussed in the following:
The theoretical scatter in [O/Mg] predicted by the
massive Population III stars lies in the range −0.3 ≤
[O/Mg] ≤ 0.3. This could help to explain the scatter
in [O/Mg] observed in metal-poor stars, if we take ob-
servational errors of the order of 0.1 dex into account.
Models with a high SF efficiency for Population III stars
indeed show some stars with high [O/Mg] ratios. However,
61% of the metal-poor stars ([Fe/H] ≤ −1.0) with ob-
served O and Mg abundances show [O/Mg] ≥ 0.1 (see
Fig. 2), whereas less than 1% of the model stars have
[O/Mg] ratios in this range (the exact number depends
on the shape of the IMF). Clearly, the observations of
metal-poor stars can not be explained as a consequence
of such massive, metal-free SNe. Furthermore, the distri-
bution of model-stars in [O/Fe] and [Mg/Fe] can not be
reconciled with observations of metal-poor stars. If the SF
efficiency of Population III stars is small, these discrepan-
cies in [O/Fe] and [Mg/Fe] disappear, but the number of
model stars with [O/Mg] ≥ 0.1 decreases even further. We
therefore conclude that -- at least for the purpose of this
paper -- the (possible) influence of SNe from very massive
Population III stars can safely be neglected.
4.3. Putting constraints on Fe yields with the help of
observations
In order to reproduce the scatter of observed [O/Fe] and
[Mg/Fe] ratios of metal-poor halo stars while keeping the
oxygen and magnesium yields fixed, we have to adjust the
stellar iron yields YFe (m) as function of progenitor mass
m. Since it is not known from theory what functional form
YFe (m) follows (increasing with m, declining or a more
complex behaviour), we have the freedom to make some
ad hoc assumptions. Nevertheless, some important con-
straints on YFe (m) can be drawn from the scatter, range
and mean of observed [O/Fe] and [Mg/Fe] abundances, as
visible from Fig. 1:
1. IMF averaged stellar yields (integrated over a complete
generation of stars) should reproduce the mean oxygen
and magnesium abundances of metal-poor halo stars,
i.e. [O/Fe] ≈ 0.4 and [Mg/Fe] ≈ 0.4.
2. Stellar yields have to reproduce the range and scatter
of [el/Fe] ratios observed. Using oxygen and magne-
sium as reference this requires:
0.0 ≤ [O/Fe] ≤ 1.2,
− 0.1 ≤ [Mg/Fe] ≤ 1.2.
(2)
(3)
(Note, that the error in abundance determinations is
of order 0.1 dex.)
3. There exist a few Type II and Type Ib/c SN ob-
servations (1987A, 1993J, 1994I, 1997D, 1997ef and
1998bw) where the ejected 56Ni mass and the mass
of the progenitor was derived by analyzing and mod-
elling the light-curve (e.g. Suntzeff & Bouchet 1990;
Shigeyama & Nomoto 1990; Shigeyama et al. 1994;
Iwamoto et al. 1994, 1998, 2000; Kozma & Fransson
1998; Turatto et al. 1998; Chugai & Utrobin 2000;
Sollerman et al. 2000). These observations give im-
portant constraints on YFe (m) since they constrain
the stellar yields for some progenitor masses, although
they are not unambiguous (see Sect. 4.3.3).
4. Since observations of metal-poor halo stars show no
clear trends in [Mg/Fe] with decreasing [Fe/H] we re-
quire that modified iron yields likewise do not intro-
duce any skewness in the distribution of model stars.
In the case of oxygen, it is not clear yet whether the
apparent slope in [O/Fe] in recent abundance studies
is real or due to some systematic errors (see Fig. 1 and
Sect. 4.1).
It is clear that it is not possible to predict YFe (m)
unambiguously, since the information drawn from obser-
vations is afflicted by errors. We therefore do not attempt
to find a solution which reproduces the observations per-
fectly, but try to extract the global properties of YFe (m)
needed to explain the behaviour of observed [el/Fe] ratios
in metal-poor halo stars.
8
D. Argast et al.: Implications of O and Mg abundances for stellar iron yields
4.3.1. IMF averaged iron yields
Since the yields of TH96 were calibrated so that the first
constraint is fulfilled, we require that the average iron
yield of SNe II stays constant when we change the pro-
genitor mass dependence of YFe (m). Assuming a Salpeter
IMF ranging from 0.1 M⊙ to 50 M⊙ and assuming that
all stars more massive than 10 M⊙ turn into core-collapse
SNe (or hypernovae, see e.g. Nakamura et al. 2001), a
SN II produces on average (cid:10)Y O(cid:11) ≈ 1.9 M⊙ of oxygen,
(cid:10)Y Mg(cid:11) ≈ 0.12 M⊙ of magnesium and hYFei ≈ 0.095 M⊙
of iron. Leaving the average oxygen and magnesium yields
unchanged and modifying the stellar iron yields, we there-
fore always have to require that on average ≈ 0.095 M⊙ of
iron are ejected per SN. Thus, YFe (m) has to satisfy the
following condition:
hYFei =
R 50
10 YFe (m) m−2.35dm
R 50
10 m−2.35dm
≈ 0.095 M⊙.
(4)
Note that the average [el/Fe] ratio of the model stars
depends on the lower and upper mass limits of stars that
turn into SNe II and their yields. If we raise the lower
mass limit, the average oxygen yield of a SN will increase
since many stars with a low oxygen yield no longer con-
tribute to the enrichment of the ISM. The same is true for
magnesium and iron and the combination of the new av-
eraged yields may lead to slightly changed average [el/Fe]
ratios. Since there are only a few SNe with large progeni-
tor masses, changing the upper mass limit will have a very
small influence on the average [el/Fe] ratios. For the re-
maining discussion, we will keep the lower and upper mass
limits of SNe II fixed at 10 M⊙ and 50 M⊙, respectively.
4.3.2. Range and scatter of observations
The second constraint can be used to calculate the range
that YFe (m) has to cover. In our picture of inhomoge-
neous chemical evolution, we assume that the first SNe
locally enrich the primordial ISM. Stars forming out of
this enriched material therefore inherit the [el/Fe] ratios
produced by these SNe which is determined in turn by
the stellar yields Y el (m). (For the time being, we neglect
the additional uncertainty hidden in the factor (1 ± ∆ǫ).)
Thus, for the first few generations of stars formed at the
time the ISM is dominated by local chemical inhomo-
geneities, the following identity holds (with the exception
of H and He where also the abundances in the primordial
ISM have to be taken into account):
[el/Fe] = log
= log
Nel,⋆/NFe,⋆
Nel,⊙/NFe,⊙
Y el (m) /YFe (m)
= log
Mel,⊙/MFe,⊙
,
Mel,⋆/MFe,⋆
Mel,⊙/MFe,⊙
where Nel,⊙ (Nel,⋆) is the number density of a given ele-
ment el in the solar (stellar) atmosphere and Mel,⊙ (Mel,⋆)
the corresponding mass fraction. (Solar abundances were
taken from Anders & Grevesse 1989). Now, let Y el (m) be
either the stellar yields of oxygen or of magnesium and
α, β the minimal and maximal [el/Fe] ratios derived from
observations. Then the constraint gives:
α ≤ [el/Fe] ≤ β
⇐⇒
α ≤ log
Y el (m) /YFe (m)
Mel,⊙/MFe,⊙
≤ β
⇐⇒
Y el (m) · 10−β ·
MFe,⊙
Mel,⊙
≤ YFe (m)
≤ Y el (m) · 10−α ·
MFe,⊙
Mel,⊙
.
Since the yields of oxygen and magnesium as function of
progenitor mass are assumed to be known, we now have
two sets of inequalities for the stellar iron yields. The first
is derived from the minimal and maximal [O/Fe] ratios
(Eq. 2):
8.37 · 10−3 · Y O (m) ≤ YFe (m) ≤ 1.33 · 10−1 · Y O (m) , (5)
and the second from the minimal and maximal [Mg/Fe]
ratios (Eq. 3):
1.23 · 10−1 · Y Mg (m) ≤ YFe (m) ≤ 2.46 · Y Mg (m) ,
(6)
where the uncertainty in the O and Mg yields given by the
factor (1 ± ∆ǫ) was neglected. Thus, for any given progen-
itor mass, YFe (m) is only determined within a factor of 20
-- 25 and further constraints are needed to derive a reliable
iron yield.
Fig. 5 shows the stellar iron yields YFe (m) (solid line)
from Thielemann et al. (1996) and Nomoto et al. (1997),
binned with a bin size of 1 M⊙. To reproduce the range
and scatter of [O/Fe] and [Mg/Fe] ratios observed in
metal-poor halo stars, YFe (m) should remain in the re-
gion enclosed by the boundaries given by Eqs. (5) and (6),
which are shown as dashed (representing [O/Fe] = 0.0),
long dashed (representing [O/Fe] = 1.2), dotted (repre-
senting [Mg/Fe] = −0.1) and dash-dotted (representing
[Mg/Fe] = 1.2) lines. Note that the lower lines in Fig. 5
represent the upper boundaries derived from metal-poor
halo stars and vice versa. Evidently, the iron yields of SNe
with progenitor masses in the range 10 − 15 M⊙ are out-
side the boundaries given by metal-poor halo stars, leading
to the stars in our model with much too low [O/Fe] and
[Mg/Fe] abundances (Paper I, c.f. also Fig. 1). Therefore,
we can already conclude that the iron yields of these SNe
have to be reduced to be consistent with observations.
Consequently, the iron yields of some higher-mass SNe
have to be increased to keep the IMF averaged [el/Fe]
ratios constant (Eq. (4)). This can easily be achieved by
assuming a higher explosion energy than the "canonical"
1051 erg of standard SN models for the more massive stars
(M ≥ 30 M⊙), as was shown recently by Nakamura et al.
D. Argast et al.: Implications of O and Mg abundances for stellar iron yields
9
sults agree remarkably well. The progenitor mass was es-
timated to be 20 ± 2 M⊙, while 0.070+0.020
−0.015 M⊙ of 56Ni
were ejected during the SN event.
SN 1993J had a progenitor in the mass range between
12 to 15 M⊙ and ejected approximately 0.08 M⊙ of 56Ni
(Shigeyama et al. 1994; Houck & Fransson 1996), which
is very similar to SN 1994I with a 13 to 15 M⊙ progeni-
tor and 0.075 M⊙ of ejected 56Ni (Iwamoto et al. 1994).
Although the amount of 56Ni ejected by those SNe lies
at the upper limit allowed under our assumptions, Fig. 5
shows that these values are still consistent with the con-
straints given by Eqs. 5 and 6.
Also consistent with our constraints is SN 1997ef with
a progenitor mass of 30 − 35 M⊙ and a 56Ni mass of
0.15 ± 0.03 M⊙ (Iwamoto et al. 2000). The correspond-
ing iron yield of SN 1997ef is higher than predicted by
the nucleosynthesis models of TH96. This is exactly the
behaviour needed to adjust YFe (m) according to our con-
straints. SN 1997ef does not seem to be an ordinary core-
collapse supernova. The model with the best fit to the
lightcurve has an explosion energy which is about eight
times higher than the typical 1051 erg of standard SN mod-
els. Such hyperenergetic Type Ib/c SNe are also termed
hypernovae and probably indicate a change in the explo-
sion mechanism around 25 − 30 M⊙ which could result in
a discontinuity in the iron yields in this mass range.
In the case of SN 1997D, the situation is not clear.
Turatto et al. (1998) propose two possible mass ranges
for its progenitor: They favour a 26 M⊙ star (although
the progenitor mass can vary from 25 − 40 M⊙) over a
possible 8 − 10 M⊙ progenitor. A recent investigation by
Chugai & Utrobin (2000) implies a progenitor mass in the
range 8 − 12 M⊙. Both groups deduce an extremely small
amount of newly synthesized 56Ni of only ≈ 0.002 M⊙ and
an unusual low explosion energy of only a few times 1050
erg. Since the situation about the progenitor mass remains
unclear, both possible mass ranges are shown in Fig. 5.
On the basis of the small amount of synthesized oxygen of
only 0.02−0.07 M⊙ (Chugai & Utrobin 2000), we strongly
favour the low-mass progenitor hypothesis, since accord-
ing to nucleosynthesis calculations by TH96 and WW95
a high-mass progenitor would produce a large amount of
oxygen (≈ 3 M⊙ for a 25 M⊙ progenitor). Moreover, in the
latter case the observed 56Ni abundance lies completely
outside the boundaries derived in Sect. 4.3.2, as can be
seen in Fig. 5.
SN 1998bw seems to be another hypernova with a ki-
netic energy of (2 − 5) × 1052 erg and may be physically
connected to the underluminous γ-ray burst GRB980425
(Galama et al. 1998; Iwamoto et al. 1998; Iwamoto 1999a,
1999b). The hypernova model assumes a progenitor mass
of about 40 M⊙, ejecting ≈ 0.7 M⊙ of 56Ni. Recently,
Sollerman et al. (2000) observed SN 1998bw at late phases
and made detailed models of its light curve and spectra.
They propose two possible scenarios for this hypernova:
one with a progenitor mass of 40 M⊙ and ejected nickel
mass of 0.5 M⊙ and the other with a 25 M⊙ progenitor
and 0.9 M⊙ of nickel. Note that the amount of nickel pre-
Fig. 5. YFe (m) as function of progenitor mass and bound-
aries constraining the range stellar iron yields have to sat-
isfy to reproduce the scatter in [O/Fe] and [Mg/Fe] of
metal-poor halo stars. According to nucleosynthesis mod-
els of Thielemann et al. (1996) and Nomoto et al. (1997),
SNe in the range 10 − 15 M⊙ clearly eject too much iron
to be consistent with observations. Also shown are obser-
vations of core-collapse SNe with known progenitor mass
and ejected 56Ni mass.
(2001). The reader should note that Thielemann et al.
(1996) and Nomoto et al. (1997) calculated models only
for the 13, 15, 18, 20, 25, 40 and 70 M⊙ progenitors. For
the 10 M⊙ progenitor we assumed an ad hoc iron yield of
one tenth of the yield of a 13 M⊙ star and interpolated
the intermediate data points. The details of the interpo-
lation and especially the extrapolation down to the 10
M⊙ star influence the mean [el/Fe] ratio of the ISM at
late stages, when it can be considered chemically homoge-
neous. However, this does not change the conclusion that
the 13 and 15 M⊙ models of TH96 produce too much
iron (if we assume the oxygen and magnesium yields to
be correct), as is evident from Figs. 1 and 5.
4.3.3. 56Ni yields from observed core-collapse SNe
There are six core-collapse SNe with known progenitor
mass and ejected 56Ni mass (which is the main source
of 56Fe in SNe II explosions, by the decay 56Ni → 56Co
→ 56Fe), namely 1987A, 1993J, 1994I, 1997D, 1997ef and
1998bw, that are shown in Fig. 5. Of these, SN 1987A is
the most extensively studied (see e.g. Suntzeff & Bouchet
1990; Shigeyama & Nomoto 1990; Bouchet et al. 1991a,
1991b; Suntzeff et al. 1992; Bouchet & Danziger 1993;
Kozma & Fransson 1998; Fryer et al. 1999) and the re-
10
D. Argast et al.: Implications of O and Mg abundances for stellar iron yields
Fig. 6. Iron yields YFe (m) respecting the constraints given
by observations of metal-poor halo stars. YFe (m) starts
at a very low value and increases continuously. A linear
increase is not possible since the mean [el/Fe] ratios have
to be conserved.
Fig. 7. Distribution of [O/Fe] ratios vs. metallicity [Fe/H]
of model stars resulting from the iron yields shown in
Fig. 6. The slope in the distribution of the model stars is
consistent with observations of oxygen abundances. A sim-
ilar slope is introduced in the [Mg/Fe] distribution which
can not be reconciled with observed magnesium abun-
dances (See text for details, symbols are as in Fig. 1).
sumably synthesized by this 25 M⊙ SN is about 10 times
larger than predicted by SN models that use the "canoni-
cal" kinetic explosion energy of 1051 erg. Nevertheless, it is
still consistent with the constraints derived in Sect. 4.3.2
and with recent calculations of explosive nucleosynthesis
in hypernovae by Nakamura et al. (2001).
4.3.4. Slopes in [el/Fe] vs. [Fe/H] distributions
Using only the constraints discussed in Sects. 4.3.1 and
4.3.2 still allows for a wide variety of possible iron yields
YFe (m). This is demonstrated by Figs. 6 and 8, where two
simple ad hoc iron yield functions are shown. The distri-
butions of model stars resulting from these iron yields are
plotted in Figs. 7 and 9.
In Fig. 6 the iron yield YFe (m) starts at the [O/Fe]
= 1.2 boundary, increases continuously with increasing
progenitor mass and ends at the [O/Fe] = 0.0 boundary.
Consequently, low-mass SNe create a high [O/Fe] ratio in
their surrounding primordial ISM, whereas it is close to so-
lar in the neighbourhood of high-mass SNe. The resulting
[O/Fe] distribution of model stars can be seen in Fig. 7.
The distribution shows a clear trend from high to low
[O/Fe] ratios with increasing [Fe/H]. A simple least-square
fit to our data yields [O/Fe] = −0.21 × [Fe/H] + 0.01. This
is in surprisingly good agreement with the result of King
(2000), who finds the relation [O/Fe] = −0.18 × [Fe/H] +
0.02 after considering the effects of NLTE corrections to
oxygen abundance determinations from UV OH-lines.
Fig. 8. Iron yields YFe (m) respecting the constraints given
by observations of metal-poor halo stars. YFe (m) starts at
a somewhat higher value than in Fig. 6, reaches a maxi-
mum at about 30 M⊙ and decreases again.
D. Argast et al.: Implications of O and Mg abundances for stellar iron yields
11
ary, increases with progenitor mass, reaches its maximum
around 30 M⊙, decreases again and ends at the [O/Fe]
= 1.2 boundary. The resulting distribution in [O/Fe]
shows a rising slope, which is clearly in contradiction with
observations. The slope is a consequence of the fact that
the iron yields YFe (m) of SNe in the range 10 − 15 M⊙
stay very close to the boundary that represents the ratio
[O/Fe] = 0.0. SNe in this mass range form the bulk of
SN II events and it is therefore not surprising that their
large number introduces such a slope in the distribution of
model stars. Thus, this YFe (m) also has to be discarded.
These simple examples show that YFe (m) should not
run parallel to the boundaries over a large progenitor mass
interval, otherwise an unrealistic slope is introduced in
the [el/Fe] distribution of model stars. They demonstrate
further, that the information drawn from metal-poor halo
stars alone is not sufficient to derive reliable iron yields,
and that information from SN II events and the shape of
the [el/Fe] distribution as function of [Fe/H] (i.e. how fast
the scatter decreases and whether slopes are present or
not) has to be included in our analysis.
5. Implications for stellar Fe yields
In Sect. (4.3.4) we have shown that the iron yields of
lower-mass SNe (in the range 10 − 20 M⊙) are crucial to
the distribution of model stars in [el/Fe] vs. [Fe/H] plots,
since progenitors in this mass range compose the bulk (ap-
proximately 69%) of SN II events. Thus, the iron yields of
lower-mass SNe should not introduce a slope in the [el/Fe]
distribution but should cover the entire range of observed
[O/Fe] and [Mg/Fe] ratios in order to reproduce the ob-
servations. To accomplish this, YFe (m) should not lie too
close to the boundaries given in Eqs. (5) and (6) in this
mass range but should start at the lower boundary ([O/Fe]
= 1.2) and reach the upper boundary ([O/Fe] = 0.0) for
some progenitor in the mass range 10 − 20 M⊙. If the ob-
served 56Ni production of SN 1993J, 1994I and 1987A are
also taken into account, the observational constraints are
stringent enough to fix the iron yields of the low mass SNe
apart from small variations: YFe (m) starts at the lower
boundary, increases steeply in the range 10 − 15 M⊙ to
the values given by SN 1993J and 1994I and remains al-
most constant in the range 15 − 20 M⊙ (to account for SN
1987A). For the remaining discussion we therefore assume
the Fe yields in this range to be ≈ 1.5 · 10−4 M⊙ for a
10 M⊙ progenitor, ≈ 5.5 · 10−2 M⊙ for a 15 M⊙ progen-
itor and ≈ 7.0 · 10−2 M⊙ for a 20 M⊙ progenitor. (The
detailed yields resulting from our analysis are listed in
Table 2). We now take a look at possible iron yields of
higher mass SNe corresponding to the different models of
the progenitor masses of SN 1997D and 1998bw. There
are four possible combinations of the progenitor masses of
those two SNe.
S1: The first case (model S1, shown in Fig. 10) gives
the best fit to abundance observations of metal-poor halo
stars. Here, we preferred the lower mass progenitor mod-
els of SN1997D and SN 1998bw over the higher mass
Fig. 9. Distribution of [O/Fe] ratios vs. metallicity [Fe/H]
of model stars resulting from the iron yields shown in
Fig. 8. Clearly, the rising slope in the distribution of model
stars is not consistent with observations of metal-poor halo
stars. Symbols are as in Fig. 1.
The reason for the slope in our model is given by the
distribution of [O/Fe] and [Fe/H] ratios induced in the
metal-poor ISM by core-collapse SNe with different pro-
genitor masses, as indicated by the position of the cir-
cles in Fig. 7. All the low mass SNe with progenitors up
to 15 M⊙ induce high [O/Fe] and low [Fe/H] ratios in
the neighbouring primordial ISM, whereas high-mass SNe
produce low [O/Fe] and high [Fe/H] ratios (recall that
a constant mixing mass of 5 × 104 M⊙ per SN event is
assumed, c.f. Sect. 2). SNe with intermediate masses in-
duce [O/Fe] ratios that lie approximately on a straight line
connecting the two extrema. The distribution of model
stars of the first few stellar generations follows this line
closely. As the mixing and chemical enrichment of the
halo ISM proceeds, the distribution then converges to the
IMF averaged [O/Fe] ratio. Although the inhomogeneous
enrichment is responsible for the slope which is in good
agreement with observations by e.g. Israelian et al. (1998),
Boesgaard et al. (1999), and King (2000), it fails to re-
produce the scatter seen in observed oxygen abundances.
Model stars with [O/Fe] ≈ 1.2 exist only at [Fe/H] ≤ −3.5
and not at [Fe/H] ≈ −2.5, where several are observed.
Furthermore, a similar slope is introduced in the [Mg/Fe]
distribution ([Mg/Fe] = −0.26 × [Fe/H] − 0.07), where
none is seen in observations and several model stars show
[Mg/Fe] ratios as large as ≈ 1.5. Therefore, this YFe (m)
has to be rejected (see however Rebolo et al. 2002 con-
cerning [Mg/Fe]).
The situation displayed in Figs. 8 and 9 is even worse.
Here, the iron yield starts at the [O/Fe] = 0.0 bound-
12
D. Argast et al.: Implications of O and Mg abundances for stellar iron yields
Fig. 10. Model S1: Iron yields YFe (m) respecting the con-
straints deduced from metal-poor halo stars and SN obser-
vations. The 10 M⊙ model of SN 1997D and 25 M⊙ model
of SN 1998bw are assumed to be correct.
Fig. 12. Model H2: Iron yields YFe (m) respecting the con-
straints deduced from metal-poor halo stars and SN obser-
vations. The 26 M⊙ model of SN 1997D and 25 M⊙ model
of SN 1998bw are assumed to be correct.
models. The curve is characterized by a peak of 0.59 M⊙
of iron at 25 M⊙ and a slow decline of the yields down
to 9.5 · 10−2 M⊙ for the 50 M⊙ progenitor. The yields
have to decline again to meet the mean abundances ob-
served in metal-poor halo stars. Obviously, YFe (m) ful-
fils the constraints discussed in Sects. (4.3.1), (4.3.2) and
(4.3.3). Since no slope is visible in the resulting distribu-
tion of [O/Fe], [Mg/Fe], [Si/Fe] and [Ca/Fe] ratios (shown
Fig. 11. Model H1: Iron yields YFe (m) respecting the con-
straints deduced from metal-poor halo stars and SN obser-
vations. The 26 M⊙ model of SN 1997D and 40 M⊙ model
of SN 1998bw are assumed to be correct.
Fig. 13. Model S2: Iron yields YFe (m) respecting the con-
straints deduced from metal-poor halo stars and SN obser-
vations. The 10 M⊙ model of SN 1997D and 40 M⊙ model
of SN 1998bw are assumed to be correct.
D. Argast et al.: Implications of O and Mg abundances for stellar iron yields
13
Table 2. Iron yields YFe (m) proposed in Sect. 5. The first
column gives the progenitor mass m in solar masses. The
following columns give the iron mass (in solar masses)
synthesized in the corresponding SN event according to
nucleosynthesis calculations of Thielemann et al. (1996)
and Nomoto et al. (1997) -- denoted by TN -- and the
models S1, S2, H1 and H2.
m
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
TN
1.6E-02
8.0E-02
1.3E-01
1.6E-01
1.6E-01
1.4E-01
1.2E-01
1.1E-01
1.0E-01
9.0E-02
8.0E-02
7.4E-02
7.0E-02
6.5E-02
6.1E-02
5.8E-02
5.9E-02
6.1E-02
6.3E-02
6.5E-02
6.7E-02
6.9E-02
7.1E-02
7.3E-02
7.5E-02
7.7E-02
7.9E-02
8.1E-02
8.3E-02
8.5E-02
8.6E-02
8.7E-02
8.7E-02
8.7E-02
8.7E-02
8.7E-02
8.7E-02
8.7E-02
8.7E-02
8.6E-02
8.6E-02
S1
1.5E-04
1.9E-03
6.8E-03
2.0E-02
5.2E-02
5.5E-02
6.0E-02
6.3E-02
6.7E-02
7.0E-02
7.1E-02
1.1E-01
1.8E-01
2.7E-01
4.2E-01
5.9E-01
5.1E-01
4.3E-01
3.7E-01
3.2E-01
2.7E-01
2.2E-01
2.0E-01
1.8E-01
1.6E-01
1.5E-01
1.4E-01
1.3E-01
1.3E-01
1.2E-01
1.2E-01
1.1E-01
1.1E-01
1.0E-01
1.0E-01
9.8E-02
9.6E-02
9.6E-02
9.5E-02
9.5E-02
9.5E-02
S2
1.6E-04
1.9E-03
6.8E-03
2.0E-02
5.2E-02
5.5E-02
6.0E-02
6.3E-02
6.7E-02
6.9E-02
7.0E-02
7.0E-02
7.0E-02
7.0E-02
7.0E-02
7.0E-02
7.0E-02
7.0E-02
7.0E-02
7.0E-02
8.2E-02
9.8E-02
1.2E-01
1.4E-01
1.7E-01
2.0E-01
2.3E-01
2.8E-01
3.3E-01
3.9E-01
4.6E-01
5.3E-01
6.3E-01
7.6E-01
8.8E-01
1.0E-00
1.2E-00
1.4E-00
1.6E-00
1.8E-00
2.1E-00
H1
1.5E-04
1.8E-03
6.7E-03
1.9E-02
5.1E-02
5.6E-02
6.0E-02
6.3E-02
6.7E-02
6.9E-02
6.9E-02
7.1E-02
7.3E-02
7.6E-02
7.8E-02
8.1E-02
1.9E-03
9.6E-03
2.3E-02
4.9E-02
8.2E-02
1.3E-01
1.8E-01
2.4E-01
3.0E-01
3.5E-01
3.9E-01
4.4E-01
4.8E-01
5.3E-01
5.8E-01
6.4E-01
6.9E-01
7.7E-01
8.6E-01
9.3E-01
1.0E-00
1.1E-00
1.2E-00
1.2E-00
1.3E-00
H2
1.5E-04
1.8E-03
6.8E-03
1.9E-02
5.1E-02
5.6E-02
6.0E-02
6.3E-02
6.7E-02
7.1E-02
7.1E-02
1.1E-01
1.8E-01
2.7E-01
4.2E-01
5.9E-01
1.9E-03
6.7E-03
1.6E-02
3.4E-02
5.8E-02
8.9E-02
1.3E-01
1.7E-01
2.1E-01
2.4E-01
2.7E-01
3.1E-01
3.3E-01
3.6E-01
3.9E-01
4.3E-01
4.5E-01
4.9E-01
5.3E-01
5.8E-01
6.1E-01
6.5E-01
7.0E-01
7.4E-01
7.9E-01
in Fig. 14), the constraint described in Sect. (4.3.4) is
also respected. The distribution of model stars in [O/Fe],
[Mg/Fe] and [Si/Fe] is in good agreement with the distri-
bution of observed stars, whereas a few stars with too low
[Ca/Fe] ratios are predicted. However, this may be due
to the fact that Ca stems from explosive O and Si burn-
ing and therefore depend on the structure of the progeni-
tor model and the (assumed) explosion energy (Paper I).
Note, that the mean [Mg/Fe] and [Ca/Fe] ratios in the
[el/Fe] plots are slightly shifted compared to the mean of
observations. This problem also occurs when the original
yields of TH96 are used (Paper I) and will persist for ev-
ery YFe (m) we present, since we did not change the mean
iron yield of 0.095 M⊙ (Eq. (4)). Especially noteworthy is
the good agreement in [Si/Fe], since we did not include
Si in the derivation of the constraints discussed above.
Moreover, the hypothetical iron yields in the other mod-
els below all result in [Si/Fe] distributions that do not fit
the observations as well as model S1. Therefore, we feel
that model S1 gives the best fit to element abundances in
metal-poor halo stars.
H1: Fig. 11 shows the iron yields under the assumption
that the higher mass models of SN 1997D and SN 1998bw
are correct (model H1). Here YFe (m) stays almost con-
stant up to 25 M⊙, followed by a sudden plunge of the
yields down to 1.9 · 10−3 M⊙ to account for SN 1997D
and then a continuous rise to 0.79 M⊙ of synthesized iron
for the 50 M⊙ progenitor that is necessary to account for
the IMF averaged yield. This sudden decrease of the iron
yields could indicate a change in the explosion mecha-
nism from supernovae with "canonical" kinetic explosion
energies of 1051 erg to hypernovae with 10 − 100 times
higher explosion energies. However, as visible in Fig. 11,
the very low iron yield of the 26 M⊙ progenitor violates
the [O/Fe] = 1.2 and [Mg/Fe] = 1.2 boundaries derived
from observations, so we would expect model stars with
much too high [O/Fe] and [Mg/Fe] ratios. This is indeed
the case, as can be seen in the corresponding [el/Fe] dis-
tribution (Fig. 15). A closer examination of the [el/Fe]
distributions, on the other hand, reveals that these model
stars are mainly present at very low metallicities ([Fe/H]
≤ −2.5). This makes it difficult to decide, whether model
H1 with its dip in YFe (m) has to be discarded or not.
The situation for oxygen remains unclear since no oxygen
abundances were measured at metallicities where the ef-
fect is most pronounced ([Fe/H] ≤ −3.0). However, in the
range −3.0 ≤ [Fe/H] ≤ −1.5 there are many observations
with [O/Fe] ≥ 0.6 whereas the bulk of model stars in this
range shows [O/Fe] ≈ 0.4. Furthermore, many observa-
tions of Mg abundances in halo stars with [Fe/H] ≤ −3.0
exist, but only one shows a ratio of [Mg/Fe] ≥ 1.0. The re-
maining stars all have [Mg/Fe] ≤ 0.8, which is in contrast
to the predictions of the model. Contrary to O and Mg,
there are indeed several observations of metal-poor halo
stars with very high [Si/Fe] and [Ca/Fe] ratios at [Fe/H]
≤ −2.5 and the fit in [Si/Fe] and [Ca/Fe] is not too bad.
However, there are some observed stars with [Si/Fe] ≤ 0.0
that are not reproduced by the inhomogeneous chemical
evolution model. All told, model H1 clearly does not fit
the observations as well as model S1.
H2: The iron yields shown in Fig. 12 (model H2) are
a result of the assumption that the 26 M⊙ model of SN
1997D together with the 25 M⊙ model of SN 1998bw are
correct. Coincidentally, YFe (m) is also compatible with the
10 M⊙ and 40 M⊙ models of SN 1997D and SN 1998bw
due to the requirement to keep the average iron yield con-
stant. Fig. 16 shows the resulting [el/Fe] distributions. Due
14
D. Argast et al.: Implications of O and Mg abundances for stellar iron yields
to the low amount of 56Ni ejected by the 26 M⊙ progenitor
that is the same for models H1 and H2 (c.f. Figs. 11, 12 and
Table 2), the problems in [O/Fe] and [Mg/Fe] described
in the discussion of model H1 still persist. Compared to
model H1, the fit in [Si/Fe] is now significantly improved,
whereas model stars with [Ca/Fe] abundances that are
too low are again generated by the inhomogeneous chem-
ical evolution code (compare with model S1). However,
although models H1 and H2 predict metal-poor halo stars
with [O/Fe] and [Mg/Fe] ratios as high as ≈ 1.5, they
can not be clearly discarded and the discovery of stars
with metallicities [Fe/H] ≤ −2.5 and 0.8 ≤ [Mg/Fe] ≤ 1.5
would be a strong argument for the validity of the sudden
decrease in the iron yields proposed by the models H1 and
H2.
S2: Finally, in Fig. 13 we show possible iron yields as-
suming that the 10 M⊙ model of SN 1997D and 40 M⊙
model of SN 1998bw are correct (model S2). Here, YFe (m)
shows a plateau in the progenitor mass range from 15 M⊙
to 30 M⊙, followed by an increasing yield with increas-
ing progenitor mass. The resulting [el/Fe] distributions
are shown in Fig. 17. A shallow slope in the [el/Fe] ratios
is visible in every element, violating the constraint dis-
cussed in Sect. (4.3.4). Furthermore, the scatter in [O/Fe]
and [Si/Fe] clearly is not fitted by the model stars. This is
especially conspicuous in the case of Si: According to the
model we would expect the bulk of [Si/Fe] abundances
to lie in the range 0.4 ≤ [Si/Fe] ≤ 0.8. To the contrary,
most of the observed stars have [Si/Fe] ≤ 0.4. Model S2
therefore gives the worst fit to element abundance deter-
minations in metal-poor halo stars.
A physical explanation for a sudden drop of the
iron yields of SNe with progenitor masses around 25 M⊙
was suggested by Iwamoto et al. (2000): Observational
and theoretical evidence indicate that stars with main-
sequence masses Mms ≤ 25 M⊙ form neutron stars with
a typical iron yield of ≈ 0.07 M⊙, while progenitors more
massive than this limit might form black holes and, due to
the deep gravitational potential, have a very low (or no)
iron yield. This might have been the case for SN 1997D.
One of the two possible models reconstructing its light-
curve assumes a 26 M⊙ progenitor and a very low kinetic
energy of only a few times 1050 erg and an equally low
56Ni yield of ≈ 0.002 M⊙. (But note that the lower-mass
progenitor model seems more likely, c.f. Sect. 4.3.3). On
the other hand, hypernovae such as 1997ef or 1998bw with
progenitor masses around 30 M⊙ and 40 M⊙ and explo-
sion energies as high as 10 − 100 × 1051 erg might be
energetic enough to allow for high iron yields even when a
black hole forms during the SN event (see e.g. MacFadyen
et al. 2001). However, one of the models of SN 1998bw
proposes a 25 M⊙ progenitor with a kinetic energy typ-
ical for hypernovae (i.e. much larger than the explosion
energy of SN 1997D). This is in some sense a contradic-
tion to the case of SN 1997D if we assume that a black
hole formed in both cases: If the explosion mechanism is
the same for SN 1998bw and SN 1997D it is natural to
assume that the explosion energy scales with the mass of
the progenitor and it is hard to imagine a mechanism that
would account for a hypernova from a 25 M⊙ progenitor
with an explosion energy that is ≈ 100 times larger than
the explosion energy from a 26 M⊙ progenitor.
Therefore, model H1 would fit nicely into the (qual-
itative) hypernova scenario proposed by Iwamoto et al.
(2000), whereas models S1, S2 and H2 (yet) lack a physi-
cal explanation. On the other hand, model S1 gives a much
better fit to the observations than H1, H2 and S2. Models
H1 and H2 could be tested, since they predict a number of
stars with very high [O/Fe], [Mg/Fe], [Si/Fe] and [Ca/Fe]
ratios (up to 1.5 dex) at metallicities [Fe/H] < −2.5. The
discovery of such ultra α-element enhanced stars would
be a strong argument in favour of the hypernova scenario
proposed by Iwamoto et al. (2000) and the existence of a
sudden drop in the iron yields of supernovae/hypernovae
with progenitors around 25 M⊙.
Table 2 lists the numerical values of YFe (m) as func-
tion of progenitor mass m for the models discussed above.
Model S1 gives the best fit to the distribution of α-element
abundances in metal-poor halo stars while S2 gives the
worst. Although the models H1 and H2 violate the con-
straints discussed in Sect. (4.3.2) and thus do not give a
fit as good as the one of S1, they cannot be ruled out on
the basis of the observational data available to date.
6. Conclusions
Inhomogeneous chemical evolution models in conjunction
with a current set of theoretical nucleosynthesis yields pre-
dict the existence of very metal poor stars with subsolar
[O/Fe] and [Mg/Fe] ratios (Argast et al. 2000). This re-
sult is a direct consequence of the progenitor mass depen-
dence of stellar yields, since core-collapse SNe of different
masses imprint their unique element abundance patterns
on the surrounding ISM. No observational evidence of the
existence of such stars is found, and recent investigations
on the contrary indicate an increasing [O/Fe] ratio with
decreasing metallicity [Fe/H]. This result of the inhomo-
geneous chemical evolution calculations is primarily due
to the input stellar yields and not due to the details of the
model itself. This is a strong indication that the progen-
itor mass dependence of existing nucleosynthesis models
is not fully understood. This in itself is not surprising,
since no self-consistent models of the core-collapse and
the ensuing explosion exist to date (c.f. Liebendorfer et
al. 2001; Mezzacappa et al. 2001; Rampp & Janka 2000).
A crucial parameter of explosive nucleosynthesis models is
the mass-cut, i.e. the dividing line between proto-neutron
star and ejecta. This gives rise to a large uncertainty in
the amount of iron that is expelled in the explosion of a
massive star. On the other hand, oxygen and magnesium
are mainly produced during hydrostatic burning and are
therefore not strongly affected by the details of the ex-
plosion mechanism. However, the distribution of [O/Mg]
ratios of metal-poor halo stars suggests that either un-
certainties exist even for O and Mg yields, or that obser-
D. Argast et al.: Implications of O and Mg abundances for stellar iron yields
15
Fig. 14. [el/Fe] distribution of model stars for O, Mg, Si and Ca resulting from the iron yields shown in Fig. 10 (model
S1).
vations overestimate oxygen or underestimate magnesium
abundances in such metal-poor stars.
The predictions of our inhomogeneous chemical evolu-
tion model can be rectified under the assumption that the
stellar yields of oxygen and magnesium reflect the true
production in massive stars well enough, and by replac-
ing the stellar iron yields of Thielemann et al. (1996) and
Nomoto et al. (1997) by ad hoc iron yields YFe (m) as
function of progenitor mass m. These are derived in this
paper from observations of metal-poor halo stars and core-
collapse SNe with known progenitor and ejected 56Ni mass
(the main source of 56Fe by the decay 56Ni → 56Co →
56Fe). Such ad hoc iron yields have to satisfy the following
constraints: First, the IMF averaged stellar yields should
reproduce the mean [O/Fe] and [Mg/Fe] abundances of
metal-poor halo stars. Second, the range and scatter ob-
served in [O/Fe] and [Mg/Fe] ratios of metal-poor halo
stars must be reproduced. This, in conjunction with stellar
oxygen and magnesium yields, leads in turn to upper and
lower boundaries for YFe (m). Third, no slope should be
introduced by YFe (m) in the [el/Fe] distribution of model
stars that is not compatible with observations. And finally,
16
D. Argast et al.: Implications of O and Mg abundances for stellar iron yields
Fig. 15. [el/Fe] distribution of model stars for O, Mg, Si and Ca resulting from the iron yields shown in Fig. 11 (model
H1).
the progenitor mass dependence of the iron yields should
be consistent with the ejected 56Ni mass of observed core-
collapse SNe with known main-sequence mass. Here, the
situation is complicated by SN 1997D and SN 1998bw.
The models recovering their light-curves give in each case
two significantly different progenitor masses. These con-
straints severely curtail the possible iron yield distribu-
tions but are not stringent enough to determine YFe (m)
unambiguously.
The main results of this paper are summarized in the
following points:
1. Observations of O and Mg abundances in metal-poor
halo stars and of the ejected 56Ni mass in core-collapse
SNe, in conjunction with oxygen and magnesium yields
from nucleosynthesis calculations and inhomogeneous
chemical evolution models, provide a valuable tool to
constrain the amount of iron ejected in a SN event as
function of the main-sequence mass of its progenitor.
2. The [el/Fe] distribution of model stars as function of
metallicity [Fe/H] is sensitive to the iron yields of SNe
with progenitors in the mass range 10−20 M⊙. A steep
increase of YFe (m) from ≈ 1.5 · 10−4 M⊙ for a 10 M⊙
D. Argast et al.: Implications of O and Mg abundances for stellar iron yields
17
Fig. 16. [el/Fe] distribution of model stars for O, Mg, Si and Ca resulting from the iron yields shown in Fig. 12 (model
H2).
progenitor to ≈ 5.5 · 10−2 M⊙ for a 15 M⊙ progenitor
followed by a slow increase to ≈ 7.0 · 10−2 M⊙ for a
20 M⊙ progenitor is required to give an acceptable fit
to the observations of metal-poor halo stars.
3. The further trend of YFe (m) in the mass range 20 −
50 M⊙ cannot be unambiguously determined by the
available data. For this mass range we have deduced
four possible iron yield distributions (models S1, S2,
H1 and H2) that explore the available freedom. These
correspond to the four different combinations of prob-
able progenitor masses of SN 1997D and SN 1998bw.
Iron yield distributions that differ significantly form
the presented models can be excluded.
4. Model S1 gives the best fit to observations while mod-
els H1 and H2 can not be ruled out. Model S2 gives
the worst fit to observed [el/Fe] ratios in metal-poor
halo stars. A change in the explosion mechanism of
SNe II around 25 M⊙ is expected in the case of the
"H" models. A test to distinguish between models S1
and H1/H2 would be the discovery of very metal-poor
stars ([Fe/H] ≤ −2.5) that are highly enriched in α-
elements.
18
D. Argast et al.: Implications of O and Mg abundances for stellar iron yields
Fig. 17. [el/Fe] distribution of model stars for O, Mg, Si and Ca resulting from the iron yields shown in Fig. 13 (model
S2).
5. Iron yield distributions derived from observations
through inhomogeneous chemical evolution models
yield constraints on the mass-cut in a SN II event
if the detailed structure of the progenitor model is
known (i.e. the size of the iron core and the zone
that undergoes explosive Si burning). Thus, they can
be used as benchmarks for future core-collapse super-
nova/hypernova models.
In future, a large and above all homogeneously ana-
lyzed sample of O, Mg and Fe abundances in very metal-
poor stars is needed to derive more stringent constraints
on YFe (m). Not only would this allow us to determine the
exact extent of the scatter in [O/Fe] and [Mg/Fe] ratios,
but would also answer the important question whether the
scatter in [O/Mg] is real or due to some (yet) unknown
systematic errors in O and Mg abundance determinations.
If the scatter in [O/Mg] turns out to be real, updated
nucleosynthesis calculations including rotation and mass
loss due to stellar winds are needed to understand O and
Mg abundances in metal-poor halo stars.
D. Argast et al.: Implications of O and Mg abundances for stellar iron yields
19
Also very valuable would be the observation and anal-
ysis of further core-collapse supernovae/hypernovae. Only
six core-collapse SNe with known progenitor and ejected
56Ni mass are known to date, and for two of them their
progenitor masses are not clearly determined. Especially
the discovery of a SN II with a progenitor in the critical
mass range from 20 − 30 M⊙ could provide us with the
information needed to discern between the four models
presented above, or at least whether the "S" or "H" mod-
els have to be preferred. This would also be a step towards
answering the question whether a change in the explosion
mechanism of core-collapse SNe occurs, i.e. the formation
of a black hole and significant increase of the explosion
energy for m ≥ 25 M⊙.
Acknowledgements. We thank the referee R. Henry for his
valuable suggestions that helped to improve this paper sig-
nificantly. D. Argast also thanks A. Immeli for frequent and
interesting discussions. This work was supported by the Swiss
Nationalfonds.
References
Anders, E., & Grevesse, N. 1989, Geochimica et Cosmochimica
Acta, 53, 197
Argast, D., Samland, M., Gerhard, O. E., & Thielemann, F.-K.
2000, A&A, 356, 873 (Paper I)
Balachandran, S. C., & Carney, B. W. 1996, AJ, 111, 946
Bekki, K., & Chiba, M. 2000, ApJ, 534, L89
Bessel, M. S., Sutherland, R. S., & Ruan, K. 1991, ApJ, 383,
L71
Beveridge, R. C., & Sneden, C. 1994, AJ, 108, 285
Boesgaard, A. M., King, J. R., Deliyannis, C. P., & Vogt, S. S.
1999, AJ, 117, 492
Bouchet, P., Danziger, I. J., & Lucy, L. B. 1991, AJ, 102, 1135
Bouchet, P., Phillips, M. M., Suntzeff, N. B., et al. 1991, A&A,
245, 490
Bouchet, P., & Danziger, I. J. 1993, A&A, 273, 451
Carretta, E., Gratton, R. G., & Sneden, C. 2000, A&A, 356,
238
Cayrel, R., Hill, V., Beers, T. C., et al. 2001, Nature, 409, 691
Chiba, M., & Beers, T. C. 2000, AJ, 119, 2843
Chugai, N. N., & Utrobin, V. P. 2000, A&A, 354, 557
Eggen, O. J., Lynden-Bell, D., & Sandage, A. R. 1962, ApJ,
136, 748
Fryer, L. C., Colgate, S. A., & Pinto, P. A. 1999, ApJ, 511, 885
Fuhrmann, K., Axer, M., & Gehren, T. 1995, A&A, 301, 492
Galama, T. J., Vreeswijk, P. M., van Paradijs, J., et al. 1998,
Nature, 395, 670
Gnedin, N. Y. 1996, ApJ, 456, 1
Heger, A., Langer, N., & Woosley, S. E. 2000, ApJ, 528, 368
Heger, A. & Langer, N. 2000, ApJ, 544, 1016
Heger, A. & Woosley, S. E. 2001, astro-ph/0107037
Houck, J. C., & Fransson, C. 1996, ApJ, 456, 811
Imbriani, G., Limongi, M., Gialanella, L., et al. 2001, ApJ, 558,
903
Iwamoto, K., Nomoto, K., Hoflich, P., et al. 1994, ApJ 437,
L115
Iwamoto, K., Mazzali, P.A., Nomoto, K., et al. 1998, Nature
395, 672
Iwamoto, K. 1999a, ApJ, 512, L47
Iwamoto, K. 1999b, ApJ, 517, L67
Iwamoto, K., Takayoshi, N., Nomoto, K., et al. 2000, ApJ, 534,
660
Jehin, E., Magain, P., Neuforge, C., et al. 1999, A&A, 341, 241
King, J. R. 1994, ApJ, 436, 331
King, J. R. 2000, AJ, 120, 1056
Klypin, A., Kravtsov, A. V., Valenzuela, O., & Prada F. 1999,
ApJ, 522, 82
Kozma, C., & Fransson, C. 1998, ApJ, 497, 431
Liebendorfer, M., Mezzacappa, A., Thielemann, F.-K., et al.
2001, Phys. Rev. D, 6310, 3004
MacFadyen, A. I., Woosley, S. E., & Heger, A. 2001, ApJ, 550,
410
Maeder, A. 1992, A&A, 264, 105
Maeder, A. & Meynet, G. 2000, ARA&A, 38, 143
Magain, P. 1989, A&A, 209, 211
McWilliam, A., Preston, G. W., Sneden, C., & Searle, L. 1995,
AJ, 109, 2757
Meynet, G. & Maeder, A. 2000, A&A, 361, 101
Mezzacappa, A., Liebendorfer, M., Messer, O. E. B., et al.
2001, Phys. Rev. Letters, 86, 1935
Molaro, P., & Bonifacio, P. 1990, A&A, 236, L5
Molaro, P., & Castelli, F. 1990, A&A, 228, 426
Moore, B., Ghigna, S., Governato, F., et al. 1999, ApJ, 524,
L19
Nakamura, T., Umeda, H., Iwamoto, K., et al. 2001, ApJ, 555,
880
Navarro, J. F., & Steinmetz, M. 2000, ApJ, 538, 477
Nissen, P. E., Gustafsson, B., Edvardsson, B., & Gilmore, G.
1994, A&A, 285, 440
Nomoto, K., Hashimoto, M., Tsujimoto, T., et al. 1997, Nucl.
Phys., A161, 79c13
Norris, J. E., Peterson, R. C., & Beers, T. C. 1993, ApJ, 415,
797
Pearce, F. R., Jenkins, A., Frenk, C. S., et. al. 1999, ApJ, 521,
L99
Peterson, R. C., Kurucz, R. L., & Carney, B. W. 1990, ApJ,
350, 173
Primas, F., Molaro, P., & Castelli, F., 1994 A&A, 290, 885
Rampp, M., & Janka, H.-T. 2000, ApJ, 539, L33
Rauscher, T., Heger, A., Hoffman, R. D. & Woosley S. E. 2002,
astro-ph/0112478
Rebolo, R., et al. 2002, in preparation
Ryan, S. G., Norris, J. E., & Bessell, M. S. 1991, AJ, 102, 303
Ryan, S. G., Norris, J. E., & Beers, T. C. 1996, ApJ, 471, 254
Samland, M. 1997, ApJ, 496, 155
Searle, L., & Zinn, R. 1978, ApJ, 225, 357
Shigeyama, T., & Nomoto, K. 1990, ApJ, 360, 242
Shigeyama, T., Suzuki, T., Kumagai, S., & Nomoto, K. 1994,
ApJ, 420, 341
Shigeyama, T., & Tsujimoto, T. 1998, ApJ, 507, L135
Sneden, C., Preston, G. W., McWilliam, A., & Searle, L. 1994,
ApJ, 431, L27
Israelian, G., Garc´ıa L´opez, R. J., & Rebolo, R. 1998, ApJ,
Sollerman, J., Kozma, C., Fransson, C., et al. 2000, ApJ, 537,
507, 805
L127
Israelian, G., Rebolo, R., Garc´ıa L´opez, R. J., et al. 2001, ApJ,
551, 833
Idiart, T., & Th´evenin, F. 2000, ApJ, 541, 207
Spiesman, W. J., & Wallerstein, G. 1991, AJ, 102, 1790
Spite, M., & Spite, F. 1991, A&A, 252, 689
Steinmetz, M., & Muller, E. 1995, MNRAS, 276, 549
20
D. Argast et al.: Implications of O and Mg abundances for stellar iron yields
Suntzeff, N. B., & Bouchet, P. 1990, AJ, 99, 650
Suntzeff, N. B., Phillips, M. M., Elias, J. H., et al. 1992, ApJ,
384, L33
Thielemann, F.-K., Hashimoto, M., & Nomoto, K. 1990, ApJ,
349, 222
Thielemann, F.-K., Nomoto, K., & Hashimoto, M. 1996, ApJ,
460, 408 (TH96)
Thomas, D., Greggio, L. & Bender, R. 1998, MNRAS, 296, 119
Turatto, M., Mazzali, P. A., Young, T. R. et. al., 1998, ApJ
498, L129
Woosley, S. E., & Weaver, T. A. 1995, ApJS, 101, 181 (WW95)
|
0709.2915 | 1 | 0709 | 2007-09-18T20:34:04 | A Connection between Star Formation in Nuclear Rings and their Host Galaxies | [
"astro-ph"
] | We present results from a photometric H-alpha survey of 22 nuclear rings, aiming to provide insight into their star formation properties, including age distribution, dynamical timescales, star formation rates, and galactic bar influence. We find a clear relationship between the position angles and ellipticities of the rings and those of their host galaxies, which indicates the rings are in the same plane as the disk and circular. We use population synthesis models to estimate ages of each H-alpha emitting HII region, which range from 1 Myr to 10 Myrs throughout the rings. We find that approximately half of the rings contain azimuthal age gradients that encompass at least 25% of the ring, although there is no apparent relationship between the presence or absence of age gradients and the morphology of the rings or their host galaxies. NGC1343, NGC1530, and NGC4321 show clear bipolar age gradients, where the youngest HII regions are located near the two contact points of the bar and ring. We speculate in these cases that the gradients are related to an increased mass inflow rate and/or an overall higher gas density in the ring, which would allow for massive star formation to occur on short timescales, after which the galactic rotation would transport the HII regions around the ring as they age. Two-thirds of the barred galaxies show correlation between the locations of the youngest HII region(s) in the ring and the location of the contact points, which is consistent with predictions from numerical modeling. | astro-ph | astro-ph |
A Connection between Star Formation in Nuclear Rings and their
Host Galaxies
Lisa M. Mazzuca
NASA Goddard Space Flight Center, Code 441, Greenbelt, MD 20771
Johan H. Knapen
Instituto de Astrof´ısica de Canarias, E-38200 La Laguna, Spain
Centre for Astrophysics Research, University of Hertfordshire, Hatfield, Herts AL10 9AB,
U.K.
Sylvain Veilleux
Department of Astronomy, University of Maryland, College Park, MD 20742
Space Telescope Science Institute, 3700 San Martin Drive, Baltimore, MD 21218
Michael W. Regan
ABSTRACT
We present results from a photometric Hα survey of 22 nuclear rings, aiming
to provide insight into their star formation properties, including age distribution,
dynamical timescales, star formation rates, and galactic bar influence. We find
a clear relationship between the position angles and ellipticities of the rings and
those of their host galaxies, which indicates the rings are in the same plane as
the disk and circular. We use population synthesis models to estimate ages of
each Hα-emitting (Hii) region, which range from 1 Myr to 10 Myrs throughout
the rings. We find that approximately half of the rings contain azimuthal age
gradients that encompass at least 25% of the ring, although there is no apparent
relationship between the presence or absence of age gradients and the morphology
of the rings or their host galaxies. NGC 1343, NGC 1530, and NGC 4321 show
clear bipolar age gradients, where the youngest H ii regions are located near the
two contact points of the bar and ring. We speculate in these cases that the
gradients are related to an increased mass inflow rate and/or an overall higher
gas density in the ring, which would allow for massive star formation to occur on
short timescales, after which the galactic rotation would transport the H ii regions
around the ring as they age. Two-thirds of the barred galaxies show correlation
between the locations of the youngest H ii region(s) in the ring and the location of
the contact points, which is consistent with predictions from numerical modeling.
-- 2 --
Subject headings: Nuclear Rings, SFR, Barred Galaxies
1.
Introduction
Nuclear rings are viewed as a type of starburst, consisting of distinct compact areas of
young massive stars near the centers of spiral galaxies. The intense bursts, attributed by
a Star Formation Rate (SFR) that significantly exceeds the average found among normal
galaxies, can dominate the overall star formation in early Hubble type barred spiral galaxies,
and dramatically alter the structure of the host galaxy with respect to morphology, gas and
dust content, metallicity, and stellar population (e.g., Martinet 1995; Buta & Combes 1996;
Kormendy & Kennicutt 2004). Typically located within the central kiloparsec, nuclear rings
contain a mixture of neutral and ionized gas and dust with masses of 108 to 1010 M⊙ (Heller
& Shlosman 1996; Rubin, Kenney, & Young 1997).
The great majority of nuclear rings are found in barred spiral galaxies, where the bar
provides essential fuel for the formation and evolution of the ring via shock-induced channel-
ing of disk gas to the nuclear regions (e.g., Schwarz 1981; Combes & Gerin 1985; Shlosman,
Begelman, & Frank 1990; Athanassoula 1992; Knapen et al. 1995b; Heller & Shlosman 1994;
Heller & Shlosman 1996). The shocks can act as instigators for angular momentum loss,
which is crucial for directing the gas towards the central regions and forming nuclear rings.
In such a picture massive SF would result and be most intense near where the bar dust lanes
intersect the ring, called the contact points. The contact points can lead the bar by as much
as 90◦ with respect to the galactic rotation (Knapen et al. 1995b; Heller & Shlosman, 1996;
Regan, Vogel, & Teuben 1997; Regan et al. 1999; Ryder, Knapen, & Takamiya 2001; Allard,
Peletier, & Knapen 2005).
While many galaxies with nuclear star-forming rings have been identified (Knapen 2005
states that 20% of local spirals host them), it is hard to generalize on the physical mechanisms
behind the triggering and propagation of massive SF due to the inconsistent and incomplete
nature of the existing data. To allow for a more statistical approach to answering some of
the mysteries of these rings, we performed an imaging survey of 73 galaxies in Hα and the
B and I bands, many of which host nuclear rings. An overview of this survey is presented
in Knapen et al. (2006; hereafter Paper I). In Paper I we classified the morphology of the
nuclear and circumnuclear Hii regions and confirmed that (1) most late-type galaxies have a
patchy circumnuclear appearance in Hα, (2) nuclear rings occur primarily in spiral types Sa-
Sbc, and (3) found that the presence or absence of a close companion does not significantly
affect the Hα morphology around the nucleus nor the strength of the nuclear Hα peak.
-- 3 --
Of these 73 galaxies, we identified 22 that host nuclear rings as judged from our Hα
images, of which 19 were previously known as such. This paper isolates the individual H ii
regions forming the nuclear rings. We present information on the observations and data
reduction in Section 2. We derive morphological parameters of the rings using ellipse fitting,
and then compare those to the disk characteristics, in Section 3. In Section 4, we identify
and analyze the H ii regions forming the rings, compute their equivalent widths (EWs), and
estimate error budgets. Using evolutionary models, we convert the EWs of the H ii regions
to ages and discuss the assumptions we make with respect to the type of starburst activity
occurring in and around the rings in Section 5. We present our age distribution analysis,
star formation rate estimations, and a discussion of the stellar bar dynamics with respect to
the rings in Section 6, before summarizing our conclusions in Section 7.
2. Observations and Data Reduction
The ring images presented here are part of a larger William Herschel Telescope (WHT)
imaging survey conducted to characterize the nuclear and circumnuclear morphology in
nearby galaxies (Paper I). We selected galaxies with some evidence for Hα structure in
their central regions, either from the literature (e.g., Buta & Combes 1996; Buta & Crocker
1993) or from our own past work. We adopted the definition of a nuclear ring, nuclear spiral,
and psuedo-ring from Buta & Crocker (1993) in Paper I and follow it here for consistency.
Since one of the aims of our overall study is the detailed analysis of nuclear rings, our
sample was biased towards SBC galaxies known to host such features. We chose rings with
enough structure to analyze, and which are considered to be the brightest feature in the
circumnuclear region. Given the nature of our original survey from which the targets were
selected, a distance bias is to some extent unavoidable. We recognize the wide range of
distances and ring sizes in the sample, and confirm that rings in farther galaxies are larger
than those in nearby galaxies. However, since our selection was with respect to the angular
size of the ring, we find that there is no correlation of distance to the number of identifiable
H ii regions.
The images presented in this paper have been obtained using the Auxiliary Port (Aux
Port) camera on the 4.2m WHT, operated by the Isaac Newton Group in La Palma. The
camera's circular field of view (FOV) is ∼1.′8 in diameter, imaged with a 1024 × 1024 pixel
TEK CCD. The angular sizes of the nuclear rings in our sample vary between 5′′ and 30′′,
which fit well into the FOV. We used Harris B and I filters, and one of four narrow-band
Hα filters, depending on the systemic velocity of the galaxy. We refer the reader to Table 1
of Paper I for a listing of the complete sample, along with their global parameters. Table 1
-- 4 --
of the present paper lists morphological type, ring ellipticity, position angle (PA) of the disk
and bar, and derived ring size for our subsample of 22 galactic nuclear rings as identified in
Paper I.
We observed spectrophotometric standards to calibrate the Hα images, and photometric
standards to calibrate the B and I band images. We observed the standards through the same
filters as the galaxies, with exposure times ranging from 5 to 20 seconds. The instrumental
magnitude of each star in each filter was obtained using the IRAF package phot. This pho-
tometry package calculates the magnitude, among other parameters, based on user-defined
values for the radius and annulus of a sky aperture. To convert from instrumental magni-
tude to AB magnitude for the spectrophotometric standard stars, an interpolation method
was used based on results from Oke (1990). Oke defines values of AB at points where no
ambiguity exists for instrument calibration problems due to spectral resolution. Therefore,
to obtain magnitudes for the filters used in our survey, magnitudes with respect to the wave-
lengths just above and below the interested filter wavelength were chosen.
Interpolating
between these values yields the absolute magnitude for the five Hα filters. The resulting
integrated fluxes of the rings (on the order of 10−12 − 10−14 erg s−1cm−2) and luminosities
of the H ii regions (on the order of 1038 − 1040 erg s−1; see Appendix A) are consistent with
published data (e.g., Pogge 1989; Storchi-Bergmann, Wilson, & Baldwin 1996; Regan et al.
1996; Sheth et al. 2000).
3. Ring Morphology
We fitted the nuclear rings of the 22 galaxies with the IRAF task ellipse. We obtained
the central position of each ring from the peak in the I-band image, and fixed the center
while iteratively fitting elliptical isophotes to the Hα image. We masked out the nucleus and
foreground stars that could affect the fitting process, and provided initial estimates of the
major axis length, ellipticity, and PA. The task produced measurements of the azimuthally
integrated intensity, observed ellipticity, and PA of the ring major axis as a function of radius.
We determined the best-fitting ellipse, which corresponded to the peak isophotal intensity
across the radial range of the ring (see Table 1). Figure 1 shows the resulting fits of the 22
rings, overlaid on the Hα images. The pixel-based ring diameters produced from ellipse
were converted to arcseconds and kiloparsecs (see Table 1, Columns 5 & 6, respectively).
The rings vary in major axis from 0.2 kpc to 1.7 kpc, which is typical for nuclear rings (Buta
& Crocker 1993). In most cases, the galaxy distance was obtained from the Nearby Galaxies
Catalog (Tully 1988); in the cases where there were no catalog values recorded, we computed
them by dividing the recessional velocity by the Hubble constant (taken at 75 km s−1 Mpc−1)
-- 5 --
-- see Paper I for details. Table 1 lists the parameters of the isophote best representing each
ring, chosen on the basis of the peak isophotal intensity across the radial range of the ring.
We compare the ellipticity and PA of each ring to that of its host disk. Disk ellipticities
and PAs are from the RC3, with the exception of the values for NGC 1530, which are based
on the kinematical data of Regan et al. (1996). Figure 2 compares the PAs and ellipticities of
the rings to their host disks. The plots show a clear linear relationship between the ring and
its host disk with respect to both PA and ellipticity. This indicates that the rings are in the
same plane as the disk and are nearly circular when deprojected. In some cases (NGC 278,
NGC 4303, NGC 4314, IC 1438, and NGC 7742) the disk is nearly circular; those galaxies
are seen face-on and do not appear in Figure 2. The disk morphology in these cases is
consistent with that of the rings, which all have very low ellipticities (less than 0.1). The
only exception to this general statement is NGC 7570 whose ring PA and ellipticity differ
significantly from those of the disk. The ring appears to be nearly face-on so the PA of the
chosen isophotal fit may not be reliable. However, the disk high inclination (i=54◦; RC3) is
not consistent with the observed nearly circular ring. This strongly suggests that the ring is
not in the same plane as the disk.
4. Equivalent Widths of Hii Regions in Nuclear Rings
4.1.
Identification of Hii Regions
To identify the individual H ii regions forming a given ring, we used the program SEx-
tractor (Bertin & Arnouts 1996), which builds a catalog of H ii regions and their associated
fluxes from an astronomical image. The various parameters used allow for the detection of
H ii regions even at very low Hα intensities. We masked out the nucleus and extraneous stars
that were clearly not part of the ring. The catalog file produced by the program records the
position, area, and flux of each H ii region detected. Figure 3 shows the outlines of the fitted
H ii regions overlaid on the continuum subtracted Hα images.
In the case of NGC 7469,
SExtractor was unable to identify any H ii regions due to poor spatial resolution com-
pared to the small angular size of the ring. This galaxy is well known to contain a nuclear
ring of 0.′′5 in radius (Genzel et al. 1995; Davies, Tacconi, & Genzel 2004; D´ıaz-Santos et
al.2007), which is consistent with our ellipse results (Section 3). The lack of sufficient
spatial resolution in our images, however, prevents us from continuing with the analysis of
the individual H ii regions in the ring, but we refer to above citations for a detailed study of
its SF properties, based on high-resolution imaging and near-IR spectroscopy.
To obtain EWs of the H ii regions, we measure fluxes in both the I-band and Hα. We
-- 6 --
do this by first imposing the position and shape of the apertures from the SExtractor
analysis of the Hα image onto the I-band image, making use of the fact that these images
have already been registered (see Paper I). We then run SExtractor in its two-image
mode to obtain the corresponding I-band fluxes. Among the many parameters to be defined
in SExtractor, two are of particular importance, namely the initial box size for the H ii
region detections, and the threshold value used to determine the extent of an individual H ii
region. For the first parameter we chose a value corresponding to the average seeing value
in our images, namely 8 pixels or 0.′′9. For the threshold value, we chose twice the standard
deviation, σ, of the combined galaxy and sky background across the whole image.
We confirmed our parameter settings by independently varying the threshold value and
box size and comparing the resulting ring morphology and EWs to our "reference" values,
which were best representative of the morphology seen in the Hα images. Several test cases
were run by altering each parameter (fixing one parameter while modifying the other) by
±50% of the reference values. Figure 4 shows the resulting images for NGC 1343, as a
representative example. Figures 4a and 4b are the result of fixing the threshold and altering
the box size. Modifying the box size by ±50% of the reference value (8×8 pixel box) to a
12×12 box and 4×4 box, respectively, causes a significant change in the visual appearance of
the H ii regions. Increasing the box size results in an obvious over-blending of the H ii regions,
with the opposite occurring when decreasing the box size. Figures 4c and 4d of Figure 4
are the products of modifying the threshold by ±50% of the reference value (2σ), to 1.33σ
and 3σ, respectively. Changing the threshold in either direction causes a significant change
in the number of H ii region detections forming the ring. Overall, the isophotal detection
area is more sensitive to the box size, while the number of detections is more sensitive to
the threshold. For small variations of the threshold, the majority of the detected regions
remained in tact. Added detections were either as a result of further segmenting an already-
defined region or adding a new region that was now above the noise level. These parameters
are important since the background to be subtracted from the I-band and Hα data is locally
defined by SExtractor for each H ii region, and includes both the sky background and
the emission from the galactic disk. In general, this empirical analysis gives us confidence
that our parameter choices do indeed best represent the H ii region morphologies.
4.2. Equivalent Widths
To obtain EWs for the individual H ii regions, the Hα flux was divided by the I-band
flux within the same aperture as defined by SExtractor. We describe below some of the
intricacies of the EW determination. We specifically discuss the correction of those cases
-- 7 --
where the I-band flux of an H ii region is particularly low, as well as the error budget of the
EW values.
4.2.1. Negative I-band Fluxes
Many of the galaxies in our sample (16 out of 20) contained H ii regions whose I-band
emission was lower than the background due to noise. This resulted in negative I-band fluxes,
which are unphysical and cannot be used to determine the EW. Rather than changing either
the H ii region aperture or specific SExtractor parameters, we chose to use a physically
meaningful upper limit to the I-band flux. We did this by taking, for all galaxies, the lowest
I-band flux per pixel (fmin) of all of the H ii regions. Analyzing the values per pixel versus
per region avoids the possible bias introduced by the large range of region sizes, which vary
by as much as a factor of two in a given ring. We conservatively assume the lowest value to
be representative of a 6σ deviation from the mean. Where fmin is negative, we correct any
I-band flux per pixel lower than the value fcorr = 1
2fmin (equivalent to a 3σ adjustment) to
have a corrected H ii region flux Fcorr = fcorr × area. This method, thus, applies a correction
to those H ii regions whose emission is positive but very low. See Appendix A for details on
individual H ii region corrections.
To estimate the impact of the 3σ corrections on the EWs, we also corrected these fluxes
to a lower level, namely 1
3fmin (equivalent to a 2σ adjustment). The maximum impact on
EW that we found in such cases was an increase in EW of order 0.2 in log(EW), although in
many cases the difference was less than 0.1. In 10 out of the 16 galaxies affected by negative
I-band emission, 5%, on average, of the H ii regions subject to the 3σ correction would not be
subject to a 2σ correction. That is, those H ii regions with very low positive I-band emission
that were changed using the 3σ discriminator were no longer affected when compared to
the 2σ value. The other six galaxies were entirely unaffected by the correction change and
can therefore be considered more reliable. We present these findings in Figure 5 using a
3σ correction. Regions whose I-band flux was changed to the 3σ value but was unaffected
by the 2σ comparison are labeled as a diamond with a cross; these represent lower-limits
in the log(EW) since we find that these values would have increased if we had chosen a 2σ
correction. All other H ii regions with corrected I-band fluxes, including the ones that were
negative, are labeled as open diamonds. Those H ii regions where no corrections were needed
are labeled as filled diamonds, and we consider those the most reliable.
-- 8 --
4.2.2. Characterizing the Uncertainties
We estimate typical uncertainties in the EW values by considering uncertainties due to
background subtraction and H ii region detection, which both result from the application of
SExtractor. Additional contributions to the overall error budget will come from measure-
ment errors in the standard stars, the calibration uncertainties associated with our method
of continuum subtraction (see Paper I), and in the filter response curves, but we will assume
that they are negligible compared to the two main contributors.
To make a conservative estimate of the uncertainty in the background subtraction we
consider the I-band backgrounds. These will be more affected than the Hα measurements
because (1) the background is higher in I than in Hα, (2) the contrast between the H ii region
and background is much smaller than in Hα, and (3) the basic morphology of an H ii region
has been set by its Hα appearance in the I-band. From our analysis of the negative residual
I-band fluxes, we know that the maximum difference in the log of the EW between the two
cases (i.e., applying a 2σ versus 3σ correction to the affected measurements) is 0.2. Such
a difference of log(EW)=0.2 corresponds to a factor of 1.5 difference in the I-background
subtracted before measuring the flux of an H ii region, where log(EW)=0.1 would correspond
to a factor of 1.3. Even considering the small field of view of our I-band images, these factors
cover the uncertainty in the determination of the background.
The other main contributor of the EW uncertainties is in the flux measurement of the
H ii regions. An RMS value, based on Gaussian noise statistics, is associated with each
detection. On average, we estimate the RMS to be 5% of the detection, which equates to
0.02 in the log of the EW. This constant RMS value is due to the fact that both flux and
source size play a role in the noise (e.g., detections with weaker fluxes cover smaller areas).
We thus arrive at a conservative error budget consisting of two components, 0.2 and 0.02
in the log(EW). These combine in a worst case scenario to a factor of 1.6, or an error budget
in log(EW) of 0.2. We take 3σ of the total error budget as a guideline for discriminating
between EWs (and thus H ii regions). As a consequence, two EW values must be 0.6 apart in
the log to make them significantly different. Whereas this is by all means a large uncertainty,
Figure 5 shows that the values can range by as much as 3 in the log for a given ring during the
evolution of massive stars so that, in fact, most of our measurements lead to straightforward
and believable conclusions on at least relative ages, as discussed in Section 6.1. We stress
that the error budget determined here is a very conservative one, and that for most H ii
regions the actual uncertainty will typically be significantly smaller.
-- 9 --
5. Age Dating
We use Starburst99 (Leitherer et al. 1999) stellar population modeling to estimate the
age of the individual H ii regions. The models present the time evolution of the EW(Hα) for
varying IMF properties and metallicities, which will affect the evolution of the population.
Figure 6 shows the time evolution of EW(Hα) for both an instantaneous starburst (i.e., all
stars are formed simultaneously in an instantaneous starburst episode) and a continuous
starburst (i.e., the star formation rate is constant in time). While the starburst choice can
greatly affect the determination of the ages of the population greater than ∼5 Myr, it is not
clear which model best represents the stellar activity in nuclear rings. Since active massive
star formation is prevalent in H ii regions, adopting an instantaneous burst would seem to fit
the model predictions of younger ages better over the range of 1Myr - 10Myr. If, on the other
hand, there is an abundant supply of gas entering the ring, starbursts may be a composite
of episodic bursts (Elmegreen 1997; Allard et al. 2006; Sarzi et al. 2007). Star formation in
the ring would cyclically enter an active starburst period until it exhausts the surrounding
gas supply, and then cease star formation until enough gas sparks a new starburst period.
Figure 7 shows Starburst99 model curves for an instantaneous and a continuous star-
burst with different IMF slopes, mass cut-offs, and metallicities. We vary the IMFs in both
scenarios and find that the evolution model is not very sensitive to the upper limit masses
(i.e., 30 M⊙ and 100 M⊙) or the IMF slopes (α = 2.35, 3.3), especially in the instanta-
neous starburst case. The models also confirm that high metallicity H ii regions show lower
EW(Hα) while low metallicity regions have invariably high EW(Hα). Regardless of the
metallicity or type of burst, however, we find that all of the predicted evolution curves show
a significant monotonic drop of EW with increasing age. Moreover, the variation shown in
the model curves (Fig. 7) is most likely an overestimate of the real variations in the nuclear
rings under consideration here. For instance, Allard et al. (2006) and Sarzi et al. (2007)
found Solar metallicity values for all nuclear rings in their sample, and there is no convincing
evidence for changes in either the IMF slope or upper mass cutoff, except possibly in rather
extreme environments such as the Galactic Center (e.g., reviews from Elmegreen 2002 and
Kroupa 2002). Since we are more interested in the relative age difference from one H ii region
to the next, the sharply declining curve (in either starburst case) allows for an unambiguous
discussions of age gradients. We adopted an instantaneous stellar starburst with a Salpeter
(1955) IMF between Mlow = 1 M⊙ and Mup = 100 M⊙, and a solar metallicity, which is a
good predictor for circumnuclear regions of barred galaxies (Allard et al. 2006).
-- 10 --
6. Discussion
6.1. Age Distribution
The results from SExtractor were used to map the ages of each H ii region according
to their azimuthal location in the ring. The right panels of Figure 5 show the resolved H ii
regions, which are color-coded according to age, in terms of EW. We assigned the physical
center of the ring (as computed with ellipse) to be (0,0), and viewed an H ii region's
azimuthal position in a counter-clockwise rotation from North to East. The color ranges are
unique to each ring and, thus, do not denote the same EWs for all rings. Although we did
not deproject the rings, all H ii regions have a unique position, which allows us to search for
any unambiguous distribution patterns. The PA of the major axes of the ring and the bar
are indicated in the image. To complement the images, plots are shown in the left panels,
which graph the log(EW) versus PA, and again indicate the PAs of the ring and bar major
axes.
We use B − I color index maps, shown in Figure 8, as morphological indicators of the
ring kinematics. These maps provide direct tracers of the bar dust lanes, which show the flow
of gas to the ring. In the cases where the dust lanes were present, the azimuthal direction of
the ring rotation was consistent with that of the galaxy in all cases. The contact points of
the dust lanes with the ring are clearly visible in most of our images and indicate an offset
of ∼90◦ between the contact points and the bar (see Section 6.3 for details). The dust lanes
also show directional flow around the ring, assuming that rotation continues in the same
direction as seen with the bar dust lanes and spiral arms. The adopted rotational direction
of the rings is listed in Table 2 and indicated for each ring containing a bar in Figure 5.
We find that several galaxies, both barred and non-barred, exhibit relative azimuthal
age gradients within the ring, as also observed by Ryder, Knapen, & Takamiya (2001) and
Allard et al. (2006) for NGC 4321. We only consider age gradients over a significant portion
of the ring (greater than 25% of the ring) with amplitudes that exceed the uncertainties for
differentiating between points over the range of interest. We highlight our findings below, and
summarize the characterization of the age distributions for all of the rings (e.g., azimuthal
gradient, radial gradient, flat distribution, or no recognizable pattern) in Table 3 (Column 2).
• NGC 278 (Fig. 5a): The entire ring as seen in Hα is very small in size (0.2 kpc in radius)
with 20 individual H ii regions detected that comprise the ring. This ring is the smallest
one recognized in the galaxy; a second "outer" nuclear ring appears at ∼ 1 kpc and is
discussed in detail in Knapen et al. (2004). This outer ring is considered 'non-standard'
as it is very wide, and more a region of star formation with rather abrupt delimitations
-- 11 --
on the inner and outer radial range. In addition to this, we were in practice unable to
derive a recognizable ring pattern of H ii regions using SExtractor, and have thus
opted for concentrating on the inner nuclear ring. The host galaxy does not contain a
bar in the optical nor near-IR wavelengths (Knapen et al. 2004). There is a decrease
in the log(EW) (i.e., an increase in age) over approximately one third of the inner
ring ranging from the positions of the H ii regions centered at azimuth 124◦ to 236◦.
This indicates a counterclockwise rotation within the ring. A radial gradient spans the
same range, with the younger H ii regions lying on the inner side of the ring. A similar
pattern, although not as clearly defined, appears over the range of 12◦ to 100◦.
• NGC 473 (Fig. 5b): Two-thirds of the ring can be seen in the Hα image, defined by
27 H ii regions. A galactic bar exists at a major axis PA of 164◦ along the major axis.
The bar dust lanes indicate a counterclockwise flow of material around the ring. There
is a sharply sloped age gradient over the range of 38◦ to 203◦ from younger to older,
respectively.
• NGC 613 (Fig. 5c): There exists a small partial nuclear ring containing 11 individual
star-forming H ii regions with a strong nucleus. A fairly flat distribution in the log(EW)
appears around the entire ring with ages averaging no less than 10 Myrs. The galactic
bar is located at PA = 110◦ along its major axis. Although a gradient is not defined, one
of the contact points (at 200◦) coincides within 5◦ to the youngest hotspot (PA = 205◦)
in its respective hemisphere.
• NGC 1300 (Fig. 5d): Twelve H ii regions form a small complete ring in Hα although
there is no clear pattern in EWs.
• NGC 1343 (Fig. 5e): The ring in this barred spiral galaxy is seen as complete in Hα
with 20 resolved H ii regions. The mapping of EW to H ii region location indicates a
clear bi-polar age pattern around the ring with EW maxima (i.e., relative youngest age
in the ring) at PA = 140 and 309◦. There is a distinct age gradient approximately from
one contact point to the other contact point (located at 172◦ and 352◦, respectively),
where the age increases counterclockwise from the bar contact points.
• NGC 1530 (Fig. 5f): Three-quarters of the ring can be seen in Hα containing 9 distinct
H ii regions. The galaxy has a large bar at PA = 122◦ and two open spiral arms
originating from its ends, rotating clockwise. We find bipolar age gradients, one in
each hemisphere, as bisected by the position of the bar. The youngest H ii regions are
near the bar PA, (i.e., not the contact points) over the range of 283◦ to 207◦ and 115◦
to 7◦, with the ring rotation also clockwise. Half of the ring is distinctly younger than
the other half, where the turnover points coincide with the ring major axis PA (= 25◦).
-- 12 --
• NGC 4303 (Fig. 5g): Only seven H ii regions define the small ring. The ring is observed
to be three-quarters complete in Hα with bipolar age maxima at PA = 114◦ and 304◦.
There are no obvious age gradients in between the maxima, but the latter are offset
by less than 25◦ from the bar contact points (100◦ and 280◦).
• NGC 4314 (Fig. 5h): The small ring consists of 10 H ii regions showing a smooth and
well-defined age gradient. The age gradient flows from older to younger in the same
direction as the presumed ring rotation (clockwise). The youngest H ii region is within
20◦ of one of the contact points (PA = 45◦).
• NGC 4321 (Fig. 5i): The well resolved full ring contains 57 detected H ii regions.
Because missing data prevent the calculation of Hα equivalent widths and ages, we
refer to Allard et al. (2006) for a similar analysis in Hβ. They find a bipolar age
gradient corresponding to the ring PA (at 170◦ and 350◦) and the location of the
youngest H ii region in each hemisphere. The age increases in a counter-clockwise
direction along the direction of flow in the ring.
• NGC 5248 (Fig. 5j): Nineteen H ii regions define the full, elliptically shaped ring. The
youngest H ii region is located 4◦ away from of one of the bar contact points (PA = 47◦).
• NGC 5728 (Fig. 5k): Eighteen H ii regions define the partial ring, which is three-
quarters complete. A very steep but small gradient exists from PA = 147◦ to PA = 98◦
from younger to older, according to the clockwise directional flow of the ring. One
contact point (PA = 303◦) is within 2◦ of the youngest H ii region in the ring. The
second youngest H ii region (PA = 147◦) is within 24◦ of the second contact point
(PA = 123◦), although it does precede that contact point.
• NGC 5905 (Fig. 5l): The ring is small but complete in Hα. It consists of nine H ii
regions and is relatively flat with respect to age (except for one H ii region). There
appears to be no correlation in age with respect to the bar contact points.
• NGC 5945 (Fig. 5m): Only three H ii regions define this patchy ring, which contains
low EW values (all less than 1.5 in the log). Buta & Crocker (1993) indicate a nuclear
ring, but our images do not have enough resolution to confirm this.
• NGC 5953 (Fig. 5n): The non-barred galaxy has a nuclear ring that appears complete
with 23 H ii regions. Although we know of no bar to directly fuel the ring, it is well
formed and resolved both azimuthally and radially.
• NGC 6503 (Fig. 5o): The elliptically shaped ring contains 84 resolved H ii regions in
this unbarred spiral. No gradients are evident.
-- 13 --
• NGC 6951 (Fig. 5p): The ring is small but complete with eight detected H ii regions.
The youngest H ii region (PA = 339◦) is within 16◦ of its respective contact point. The
ages steeply increase (by a factor of 10) from 339◦ to 218◦ in a clockwise direction,
consistent with the directional flow of the galaxy.
• NGC 7217 (Fig. 5q): The ring is three-quarters complete in Hα. Twenty-two H ii
regions were detected in the ring of this unbarred galaxy. Ages increase in a clockwise
direction from PA = 109◦ to PA = 6◦ .
• IC 1438 (Fig. 5r): This barred galaxy hosts a complete ring with 11 H ii regions
detected. There exists an increasing age gradient from PA = 32◦ to PA = 211◦ in a
counterclockwise direction. This direction is consistent with the ring rotational flow.
The youngest H ii region is within 2◦ of its respective bar contact point (PA = 304◦)
• NGC 7570 (Fig. 5s): Eighteen H ii regions comprise a mostly complete ring in this
barred galaxy. An increasing age gradient ranges counterclockwise from PA = 112◦ to
PA = 295◦. There is also a radial gradient spanning the same range as the azimuthal
gradient, with two distinct arcs at nearly identical slopes, and where the inner arc
appears consistently older than the outer one. No correlation to the bar contact points
is evident.
• NGC 7716 (Fig. 5t): Thirty H ii regions form a complete ring in this barred galaxy.
No gradients or correlations to the bar are observed.
• NGC 7742 (Fig. 5u): The circular and well resolved ring contains 38 H ii regions. No
gradients are observed.
While many of the rings contain age gradients, a few show no gradient, namely NGC 613,
NGC 5905, and NGC 5945. This type of distribution is qualitatively more consistent with
Elmegreen's (1994) predictions, which suggest that the massive star formation in the nuclear
ring is shocked into existence due to gravitational instabilities in the nuclear region around
the ILRs. Therefore, age gradients would not exist. Comparisons of EW to age for these
three rings also indicate that these starbursts are relatively older than those seen in the other
rings, on the order of 10 Myrs and higher. Elmegreen (1997) suggests that these rings (i.e.,
with no gradient) could last for as long as 100 Myrs before SF ceases if the inflowing gas rate
is high enough. The SFRs of these three rings (2M⊙/yr, 2M⊙/yr, and 4M⊙/yr, respectively;
see Section 6.2 and Table 1 for SFR and inflow discussion) are among the higher values in
our sample but are consistent with other rates documented (Regan et al. 1997; Jogee et al.
2002). If we take into account that NGC 613 and NGC 5945 are partial rings in the optical,
then the SFRs calculated here are lower limits.
-- 14 --
6.2.
Inflow and Star Formation Rates
In this Section we explore possible reasons for the result that only three galaxies show
bi-polar age gradients around their rings, whereas many of the rings (9 of 22) show partial
gradients along 25% - 30% of their rings, with the rest (10 of 22) showing no gradient.
We find no apparent relationships between the presence or absence of age gradients and
the morphology of the rings or their host galaxies. For example, whereas NGC 4321, with its
bi-polar age gradient, has a pair of well-defined dust lanes coming into the ring through the
bar, NGC 1300 and NGC 5248 have similar dust lane morphologies yet show no gradients.
Some of the poorly resolved rings, such as those in NGC 4303 or NGC 5945, do not show
gradients, but given the presence of measured age gradients in the comparably small rings
in NGC 6951 and IC 1438, the lack of spatial resolution in the former two cannot easily be
used as an argument for the lack of a gradient. But even if we ignore galaxies with either
small and badly resolved rings, or rings that do not have the classical aspect of dynamically-
induced resonance rings (NGC 613 and NGC 6503 are in the latter category), we are left
with a significant number of rings in which we should have detected gradients if present, but
which for some reason do not show any. This category includes, for instance, NGC 1300, very
similar to NGC 4314 or NGC 6951 which do show gradients (all compact well-defined rings
with around a dozen H ii regions, sitting in strong bars with well-defined sets of dust lanes);
NGC 5248, similar to NGC 4321 (complex, rich rings with many H ii regions organized into
spiral arms fragments and hosted by a relatively weak bar); NGC 7716, similar to NGC 7217
which does have a gradient; and NGC 7742. For those rings that contain a gradient we also
find no correlation between the presence of the gradient and direction of the ring rotation.
In relating the location of the youngest H ii regions to those of the contact points between
the dust lanes and ring, we have been assuming that massive SF will follow very rapidly once
the inflowing gas joins the ring. The appearance of a gradient in the ring may thus depend
on whether or not material flowing in will immediately trigger gravitational collapse, and
initiate massive SF. This in turn may be due to either a higher mass inflow rate (or possibly
increased clumpiness of the inflow) and/or an overall higher existing gas density in the ring.
Each of these possibilities will ensure that soon after a clump of inflowing gas joins the ring,
it will start to form massive stars. The alternatives, of low inflow rate and/or overall low
gas density in the ring, might lead to a more gradual increase of the gas density, without
necessarily leading to immediate gravitational collapse and massive SF on short timescales.
In this case, the positions of the H ii regions would bear no relation to the locations of the
contact points. The ring would, in such a case, merely increase its gas density gradually,
until at some stage, somewhere along the ring, an individual clump would become unstable
to collapse; therefore, no age gradients would ensue.
-- 15 --
To test this simple model, we must first consider the quantities mentioned above, gas
inflow rate and gas density in the ring. These can in principle be determined from inter-
ferometric CO measurements, but the process is not straightforward. The main issues are
the conversion factor of LCO to MH2 (the X-factor), which may well not be equal to the
canonical value in the kind of environments we consider here, and the determination of the
fractions of detected molecular gas which are in fact flowing into the ring rather than around
it (as a consequence of a "spray-type" flow, see, e.g., Regan, Vogel & Teuben 1997; Jogee et
al. 2002), or with the right physical conditions such as velocity dispersion and clumpiness to
take part in the massive SF process. So although CO interferometry has been presented for
seven of our nuclear rings in the literature (by Regan, Vogel & Teuben 1997; Sakamoto et al.
1999; Jogee et al. 2002, 2005; Helfer et al. 2003; Combes et al. 2004; Garc´ıa-Burillo et al.
2005), we will not attempt here to derive the gas inflow rate and gas density in the ring from
this collection of data, all with different observational characteristics. For two of our rings,
NGC 1530 and NGC 5248, gas inflow rates have been explicitly determined in the literature.
M = 1 M⊙ yr−1, whereas Jogee
For the former galaxy, Regan et al. (1997) give a value of
et al. (2002) determine a value of "a few" M⊙ yr−1 for the latter galaxy. Considering the
uncertainties in these values, we can say little more than that they are consistent with the
SFRs determined for our nuclear rings (see discussion below, and values in Table 1).
We could assume that the SFR of the ring can act as a tracer of the inflow rate, although
that would imply assuming, for instance, that inflowing gas is transformed into massive stars
on very short timescales. If we do this, we would expect the age gradients to occur in those
galaxies with the higher SFRs, which would be representative of higher gas inflow rates. In
contrast, lower SFRs would indicate a low inflow rate so that the gas will gradually fill the
ring, leading to SF only if and when the local gas density leads to gravitational collapse.
This will occur essentially at random and thus result in a ring with no gradients.
Kennicutt (1998) derives a relation between the Hα ionizing flux and the SFR for normal
disk star forming galaxies, where the conversion factor is computed using stellar population
synthesis modeling. Since only stars with masses larger than 10 M⊙ and lifetimes less
than 20 Myr dominate the integrated ionizing flux, the Hα emission line provides a nearly
instantaneous and linear measure of the SFR, independent of the previous star formation
history. For solar abundances and our adopted IMF (α = 2.35; Mlow = 1 M⊙ and Mup =
100 M⊙), we can use Kennicutt's (1998) equation:
SFR(M⊙ yr−1) = 7.9 × 10−42 L(Hα) (erg s−1).
to compute SFRs integrated over the entire ring, which are given in column 10 of Table 1.
Variable extinction certainly adds uncertainty in the computation of the SFR for cir-
-- 16 --
cumnuclear regions. From the B − I images, the patchiness of the extinction is very clear
in several of our galaxies (e.g., NGC 473, NGC 613, NGC 1530, NGC 5945), which reveal
only partial rings in Hα. The less obscured regions will contribute more heavily to the line
fluxes in these cases. The B − I images clearly indicate that while the dust lanes can be
traced nicely in the bar inmost cases (where there is little star formation), the appearance
of the nuclear rings is dominated by blue star-forming regions. This type of morphology
makes it difficult to precisely model the effect of the dust, but attempts have been made
for nuclear regions in particular. Osmer, Smith, & Weedman (1974) analyzed six galaxies
and derived integrated extinction values across the entire circumnuclear region ranging from
AHα ≈ 2.5 − 3 mag, while Hummel, van der Hulst & Keel (1987) found an extinction of
AHα ≈ 1.1 in the central region of NGC 1097. Kennicutt (1998) acknowledges the high un-
certainty and estimates that the errors in the SFR can be as high as ±50%. A high-resolution
survey paper by Phillips (1996) for nine circumnuclear regions confirms the potential for high
uncertainty values. Following Kennicutt, he finds that the SFRs vary by as much as a factor
of ten when applying extinctions of AHα=1.1 and AHα=3.4 mag. He concludes that circum-
nuclear SF can be a dominant factor to the overall rates of such galaxies. We thus place the
same caution that our SFRs contain potentially high uncertainties due to variable extinction
throughout the nuclear rings.
Although there are possible large uncertainties in the SFR due to extinction, using the
SFR can provide a possible link to an explanation for the existence of age gradients, via the
gas inflow rate. Figure 9 plots the integrated SFR over the entire ring for all galaxies and
distinguishes those rings that show a gradient and those that do not. There is no obvious
correlation between the SFR and the appearance of a gradient, although we cannot exclude
that such a connection does exist but is obscured by high dust extinction. The average mean
for those without gradients (SFR = 3.6 ± 1.1 M⊙ yr−1, where the uncertainty is the standard
error) is slightly larger for those with a gradient (SFR = 2.2 ± 0.7 M⊙ yr−1), but given the
large mean deviations, we cannot distinguish in a statistical sense between the two cases.
This could be due to several factors. All of the gradients, except for those in NGC 1343 and
NGC 4321, are partial gradients. That is, their gradient does not range continually from one
contact point to the next. NGC 1343 and NGC 4321, which represent the"ideal" scenario,
do have high SFRs (6.8 M⊙ yr−1 for both cases) when compared to the rest of the sample.
Computing the average SFR of the ring also is insensitive to the steepness of the gradient,
the size of the ring, the size of the individual H ii regions, or the patchiness of the ring, which
all add complexity to the interpretation of the data.
From our analysis of age gradients we thus conclude that although age gradients along
the ring are detected in more than half of the rings, all other rings show random variations
of H ii region EW, and thus age, with azimuth. We cannot pinpoint any morphological
-- 17 --
parameters that might distinguish the host galaxies and bars of those rings with from those
without gradients. Another strong conclusion from our analysis is that the kind of clear
bi-polar age gradient seen before in NGC 4321 (Ryder, Knapen, & Takamiya 2001; Allard
et al. 2006) is rare - we only see a similar bi-polar gradient in one more galaxy, NGC 1343.
It is not clear why these two galaxies show their gradients, whereas other galaxies with very
similar bar-ring systems show gradients that are not bi-polar, or no gradients at all.
6.3. Nuclear Ring and Stellar Bar Dynamics
It is agreed that the presence of a galactic bar can have significant influence on angular
momentum and re-distribution of gas, including inflow towards the central regions (e.g.,
Combes & Gerin 1985; Athanassoula 1992, 2000; Knapen et al. 1995a; Heller & Shlosman
1996; Maciejewski et al. 2002). Observed kinematics reveal that the shock-induced flows
move gas radially inwards along the bar dust lanes and tend to collect at the intersection
of the dust lanes with the ring. These contact points are thought to be approximately
perpendicular (within the plane of the disk) to the PA of the bar major axis (Regan &
Teuben 2003) but have been seen to lead the bar by as little as 60◦ in numerical models
(Heller & Shlosman 1996). The pile up of gas at these two points in the ring can be expected
to spark SF. Therefore, one could predict to find the youngest stellar H ii regions in the ring
near the contact points.
For the barred spiral galaxies in our sample, we compare the location of the youngest
H ii regions in the ring to the derived location of the contact points. We first identified the
youngest H ii region in each hemisphere, where the hemispheres are separated by the bar
PA. Taking into account the direction of flow around the ring (from our B − I images), we
calculate the PA offset between the H ii region and the contact points. Table 2 shows the
PAs of both the contact point and the youngest H ii region in each hemisphere, along with
the directional flow of the ring. We find that in two-thirds of the galaxies, one of the bar
contact points is aligned with the youngest H ii region in the associated hemisphere within
20◦ of the contact point. Two galaxies, NGC 4314 and NGC 5248, have the youngest H ii
region in both hemispheres aligned with the contact points. Our results for both of these
rings are consistent with those of Benedict et al. (2002) and Maoz et al. (2001), respectively,
who also find that the youngest stars were located at the contact points. Figure 10 shows
a histogram of the distribution of the angular offsets between the bar major axis and the
locations of the two youngest Hii regions in each ring. A peak in frequency around the
contact point locations is seen.
We next calculate the times from the bar major axis PA to the location of the contact
-- 18 --
points with those of the youngest H ii region in each hemisphere, assuming we azimuthally
move in the direction of ring flow. We compute the period of the rings using measurements
of the ring radius and the rotational velocity at the ring radius. We chose the major axis
length of each ring as derived from ellipse (see Section 3 and Table 1), and obtained the
rotational velocity values from the literature and from our own DensePak IFU data taken in
2003 & 2004 (L. M. Mazzuca et al. 2006, in prep.). Table 3 shows the major axis ring radius,
the adopted rotational velocity along with its reference, and the period for each ring. Since
we predict the contact points to be ∼90◦ from the bar major axis, we can compute their
location with respect to the bar PA in each hemisphere, and thus calculate the amount of
time it would take to travel to that location. We compare this travel time to that associated
with the location of the youngest H ii region in each hemisphere. Overall, the expected
travel times to the contact points coincide with the times computed from the bar PA to the
youngest H ii regions, which further strengthens the theory that the youngest H ii regions are
located near the contact points. We find only one H ii region, in NGC 1530, that deviates
from the general trend. In this case, the youngest region in one of the hemispheres is more
aligned with the PA of the ring than that of the bar.
The analysis described above considers merely the rotational velocity of the gas in
the ring region while deriving travel times. There is an additional contribution from the
bar, which rotates with a certain bar pattern speed (assuming that the ring co-rotates
with the bar). Estimating the relative contributions of these two rotation components,
we consider parameters for a typical ring in our sample (see Table 1) of rring = 0.5 kpc,
Ωp,bar = 40 km s−1 kpc−1, and vrot = 150,km s−1, which yields a ratio of the contributions of
circular to bar pattern speeds into the velocities around the ring of (40 × 0.5)/150. Given
that the rotational velocity is dominant, that the uncertainties in determining the bar pat-
tern speed (and, in fact, the direction of bar rotation) are significant (see Elmegreen 1996
and Knapen 1999 for reviews), and that the uncertainties from other sources, as described
above, in the determination of travel times around the ring are already rather large, we have
not pursued to derive bar pattern speeds and incorporate them into our analysis.
7. Conclusions
From a larger Hα imaging survey we identified 22 galaxies that contain a nuclear ring.
To identify their sizes and shapes, we fitted each ring with an ellipse based on the peak
isophotal intensity across the radial range of the ring. The rings vary in size from 0.2 kpc
to 1.7 kpc in major axis radius. We compare the ring morphologies to those of their host
galaxies. In all but one case the ellipticity and PA of the rings approximately match that of
-- 19 --
their respective disk, which corroborates the idea that the ring is in the same plane as the
disk. The exception is NGC 7570, which shows a distinct and self-consistent difference in
both the PA and eccentricity in the ring versus the disk.
All but one of the rings had sufficient spatial resolution for us to identify individual H ii
regions. For each H ii region, we computed the EW and converted differences between them
into relative ages and age gradients using the Starburst99 evolutionary models. Although
we adopted an instantaneous starburst model, we acknowledge that the rings may be more
representative of a hybrid episodic scenario where star formation toggles between latent and
active periods corresponding to the time it takes for gas build-up to reach critical density.
Three rings, NGC 1343, NGC 1530, and NGC 4321, contain bipolar age gradients, where
the ages increased from each of the contact points. Many of the rings in our sample exhibit
a well-defined age distribution pattern, but not throughout the entire ring. This can be a
result of the combination of bar-induced dynamics and gravitational instabilities that are
occurring on an intermittent basis. We calculated the SFR in each ring from the integrated
luminosity contributed by all of the H ii regions forming the ring, and compared those rates
to the appearance of a gradient within each ring (or lack thereof). The large uncertainties
associated with variable extinction across the rings and the large dispersion of SFR values for
both cases yields no clear correlation between the SFR and the emergence (or lack thereof)
of an age-dependent azimuthal distribution. We do see a connection between the location
of the bar contact points and the youngest H ii regions in two-thirds of the galaxies in our
sample.
acknowledgements We thank Dr. Isaac Shlosman for inflow rate discussions and Dr.
Torsten Boker for input into earlier stages of the research. We also thank the referee for the
thorough review of this paper. JHK acknowledges support from the Leverhulme Trust in
the form of a Leverhulme Research Fellowship. The WHT is operated on the island of La
Palma by the Isaac Newton Group in the Spanish Observatorio del Roque de los Muchachos
of the Instituto de Astrof´ısica de Canarias. This research has made use of the NASA/IPAC
Extragalactic Database (NED), which is operated by the Jet Propulsion Laboratory, Cal-
ifornia Institute of Technology, under contract with the National Aeronautics and Space
Administration.
-- 20 --
ǫRing P ARing
RRing
(◦)
(4)
120
160
122
135
60
25
88
135
170
115
125
141
105
192
121
146
89
130
38
135
30
133
(′′)
(5)
4.4 × 4.1
12.2 × 6.9
5.1 × 2.6
4.1 × 3.1
8.8 × 6.6
6.8 × 4.9
3.3 × 2.8
6.6 × 5.9
8.8 × 7.0
6.6 × 4.6
5.3 × 3.2
1.6 × 1.5
3.5 × 3.2
6.1 × 5.5
38.5 × 13.4
4.6 × 3.7
11 × 8.8
3.3 × 3.0
1.6 × 1.5
4.4 × 4.2
6.3 × 4.2
9.9 × 9.4
(3)
0.06
0.37
0.3
0.25
0.25
0.35
0.14
0.10
0.12
0.3
0.4
0.2
0.1
0.1
0.65
0.2
0.20
0.10
0.09
0.05
0.2
0.05
RRing
(kpc)
(6)
0.2 × 0.2
1.7 × 1.0
0.4 × 0.2
0.3 × 0.2
1.2 × 0.9
1.2 × 0.8
0.2 × 0.2
0.3 × 0.3
0.7 × 0.6
0.7 × 0.5
1.1 × 0.6
0.3 × 0.3
1.2 × 1.1
1.0 × 0.9
1.1 × 0.4
0.5 × 0.4
0.8 × 0.7
0.5 × 0.5
0.5 × 0.5
1.3 × 1.3
1.0 × 0.7
1.0 × 1.0
NGC
(1)
278
473
613
1300
1343
1530
4303
4314
4321
5248
5728
5905
5945
5953
6503
6951
7217
IC1438
7469
7570
7716
7742
Morph
Type
(2)
SAB(rs)b
SAB(r)0/a
SB(rs)bc
(R)SB(s)bc
SAB(s)b
SB(rs)b
SAB(rs)bc
SB(rs)a
SAB(s)bc
(R)SB(rs)bc
(R1)SAB(r)a
SB(r)b
SB(rs)ab
SAa
SA(s)cd
SAB(rs)bc
(R)SA(r)ab
SAB(r)a
SAB (rs)a
SBa
SAB(r)b
SA(r)b
P ADisk P Abar
SFR
(◦)
(7)
-
153
120
106
80
8
-
-
30
110
-
135
105
169
123
170
95
-
135
30
35
-
(◦)
(8)
no bar
164
120
102
82
122
10
135
153
137
33
25
10
no bar
no bar
85
no bar
124
56
30
34
no bar
( M⊙ yr−1)
(9)
0.5
2.2
2.2
0.2
6.8
3.8
1.4
0.1
-
4.2
4.0
2.6
4.4
9.9
1.5
1.4
0.6
1.3
-
1.4
3.2
4.3
Table 1: Morphological characteristics of the galaxies in the observed sample. Galaxies are
listed in order of increasing RA (Col. 1); morphological type as stated in the RC3 (Col. 2);
ellipticity of the nuclear ring as derived by using the IRAF ellipse task (Col. 3); PA of
the ring major axis, calculated counterclockwise from North to East (ellipse; Col. 4); ring
radius in arcseconds (Col. 5) and kpc (Col. 6); PA of the galactic disk from RC3 (a dash
represents a circular ring with no defined PA; Col. 7), except for NGC 1530 (Regan et al.
1996) which is based on kinematical data; PA of the bar major axis, except for NGC 1530
(Regan et al. 1996) and NGC 7469 (Davies, Tacconi, & Genzel 2004) which are based on
kinematical data (Col. 8); and the ring SFR (a dash represents lack of data necessary to
calculate the SFR; Col. 9) is derived from the total ring luminosity using the Kennicutt
(1998) Hα-SFR relation.
-- 21 --
NGC
(1)
473
613
1300
1343
1530
4303
4314
5248
5728
5905
5945
6951
IC1438
7570
7716
CP #1 H ii region #1 CP #2 H ii region #2 Rotation Diff PA #1 Diff PA#2
(2)
254
210
12
172
32
280
45
47
303
295
100
355
214
300
124
(3)
312
205
345
140
344
304
27
43
305
343
148
339
251
353
91
(4)
74
30
192
352
212
100
225
227
123
115
280
175
34
120
304
(5)
38
323
269
309
282
114
222
243
147
151
280
100
32
131
301
(6)
ccw
ccw
cw
ccw
cw
cw
cw
ccw
cw
cw
ccw
cw
ccw
cw
ccw
(7)
58
15
27
32
48
24
18
4
2
48
48
16
37
53
33
(8)
36
67
77
43
70
14
3
16
24
36
0
75
2
11
3
Table 2: Comparison of the location (i.e., PA) of the contact points (CPs) to that of the
youngest H ii region in each hemisphere, as bisected from the bar. The PAs of each contact
point, defined to occur 90◦ offset from the bar PA, are given in Col. 2 and 4; the location of
the youngest H ii region in each hemisphere is given in Col. 3 and 5 with the ring rotational
direction cited in Col 6, as observed in the B − I images; Col. 7 and 8 indicate the difference
in location between the contact points and the respective H ii region.
In most cases the
separation is small, which is conistent with the notion that the youngest H ii region (in its
respective hemisphere) should be near the location of the contact points.
NGC
(1)
278
473
613
1300
1343
1530
4303
4314
4321
5248
5728
5905
5945
5953
6503
6951
7217
IC1438
7570
7716
7742
Age
Distrib
(2)
az.gradient
az. gradient
flat
no pattern
az. gradient
az. gradient
no pattern
az. gradient
n/a
az. gradient
az. gradient
flat
flat
rad gradient
no pattern
az. gradient
az. gradient
az. gradient
rad gradient
no pattern
no pattern
-- 22 --
Vrot
( km s−1)
Ref Period CP time H ii region #1 H ii region #2
(Myr)
(Myr)
(Myr)
(Myr)
(3)
63
127
115
135
131
185
97
161
170
152
180
150
150
200
90
160
248
204
150
150
116
(4)
(a)
IFU
(b)
(c)
IFU
IFU
IFU
IFU
(d)
IFU
(e)
(f)
(g)
IFU
(h)
(c)
(i)
(j)
(g)
(g)
IFU
(5)
35
103
23
17
39
50
9
30
32
40
47
15
62
40
94
24
25
19
67
51
96
(6)
-
72
13
0.6
18
4
6
4
-
5
45
11
17
-
-
24
-
11
66
18
-
(7)
-
89
13
16
15
48
8
2
-
5
40
14
25
-
-
23
-
13
65
13
-
(8)
-
62
9
4
14
14
7
3
-
7
43
14
17
-
-
19
-
11
57
17
-
Table 3: Age distribution and H ii region times. Col. 2 cites the age distribution, which
fell into 4 categories: azimuthal gradient, radial gradient across the ring, flat, no pattern.
For those rings with host bars, the period of the ring (Col. 5) is computed based on the
rotational velocity at the ring radius (Col. 3) and the ring major axis radius (see Table 1,
Col. 5); literature references for the rotational velocites are in Col. 4: a) Knapen et al.
(2004); b) Blackman (1981); c) Knapen, Laine, & Relano (1999); d) Knapen et al. (2000);
e) Rubin (1980); f) Moiseev, Vald´es, & Chavushyang (2004); g) no reference found- assumed
a rotational velocity value of 150 km s−1, which is the average of the measured values for
the other rings (all rings in this sample have similar host galaxy morphologies and rotation
curve characteristics associated with the ring radial location) ; h) Bottema (1989) for Hβline;
i) Peterson et al. (1978); j) P. M. Treuthardt & R. J. Buta, private communication; 'IFU'
refers to our WIYN IFU data (L. M. Mazzuca et al. 2006). Travel times from the bar major
axis PA to the contact points (Col 6) and to the youngest H ii region in each hemisphere
(Col. 7 and 8, when a bar exists.
-- 23 --
A.
Individual H ii Region Parameters for the Galaxies
For each ring in the sample we present tables that contain each H ii region as identified by
SExtractor (col. 1), integrated flux (col. 2), luminosity (col. 3), EW (col 4) and log(EW)
(col. 5). Column 6 indicates if the I-band flux needed to be modified in order to compensate
for unrealistic values (i.e., negative flux values or marginally positive flux values -- see Section
4.2.1). In the cases where the I-band flux generated from SExtractor was not modified,
a value of '0' is recorded; otherwise a value of '1' is recorded. Since SExtractor detects
all H ii regions in the supplied image, extraneous H ii regions as well as the nucleus were
included in the identification process. These superfluous detections are not included in the
tables, but are instead noted in the captions for completeness sake.
-- 24 --
H ii region
Integrated Flux
(10−13 erg s−1cm−2)
Luminosity
(1039 erg s−1)
(1)
1
2
3
5
7
8
9
10
11
12
13
14
18
19
20
21
22
23
24
25
26
(2)
3.38
0.98
0.56
1.74
4.98
0.52
0.7
6.20
0.51
0.86
0.66
0.98
3.36
2.62
1.12
0.87
0.54
0.72
0.90
1.02
3.54
(3)
5.63
1.63
0.93
2.91
8.30
0.86
1.17
10.3
0.85
1.44
1.10
1.63
5.61
4.36
1.87
1.45
0.91
1.20
1.51
1.71
5.90
EW log(EW)
(A)
(4)
258
229
84
485
536
81
143
3341
(A)
(5)
2.41
2.36
1.92
2.69
2.73
1.91
2.15
3.52
0.56
2.16
2.11
2.15
2.56
2.81
2.45
2.06
1.72
2.12
2.35
2.20
3.21
4
146
128
142
365
650
284
114
53
133
222
157
1620
Flux
Correction
(6)
0
1
1
1
0
1
1
1
0
0
1
0
0
0
1
1
1
1
1
0
1
Table 4: NGC 278 - 20 H ii regions detected associated with the ring. H ii regions # 4, 6,
and 11 comprise the nucleus; H ii regions # 15, 16, and 17 are extraneous to the ring.
-- 25 --
H ii region
Integrated Flux
(10−13 erg s−1cm−2)
Luminosity
(1039 erg s−1)
(1)
1
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
25
26
28
29
30
32
33
(2)
0.33
4.01
1.42
3.34
1.88
0.62
0.35
0.37
1.03
1.15
0.42
0.22
1.32
0.11
1.13
0.60
0.32
1.60
2.08
0.83
1.55
0.17
0.32
0.10
0.22
0.19
0.26
(3)
3.49
42.6
15
35.5
20
6.57
3.73
3.93
10.9
12.3
4.44
2.37
14.1
1.20
12
6.37
3.37
17
22.1
8.85
16.5
1.83
3.36
1.05
2.32
2.02
2.79
EW log(EW)
(A)
(4)
216
9343
1643
1409
2660
608
222
262
942
1812
389
122
647
41
1722
647
200
2241
2622
1138
1261
70
192
25
120
78
107
(A)
(5)
2.33
3.97
3.22
3.15
3.42
2.78
2.35
2.42
2.97
3.26
2.59
2.09
2.81
1.62
3.24
2.81
2.30
3.35
3.42
3.06
3.10
1.85
2.28
1.41
2.08
1.89
2.03
Flux
Correction
(6)
1
1
0
0
0
1
1
1
0
1
1
1
0
1
1
1
1
0
0
1
0
1
1
1
1
1
1
Table 5: NGC 473 - 27 H ii regions detected associated with the ring. H ii regions # 23 and
24 comprise the nucleus; H ii regions # 2, 3, 27, and 31 are extraneous to the ring.
-- 26 --
H ii region
Integrated Flux
(10−13 erg s−1cm−2)
Luminosity
(1039 erg s−1)
(1)
1
3
4
5
6
7
8
9
10
12
13
(2)
26.3
13.3
3.87
2.26
8.49
1.00
2.00
0.84
0.86
11.4
5.89
(3)
96.4
48.9
14.2
8.28
31.1
3.66
7.33
3.08
3.14
41.8
21.6
EW log(EW)
(A)
(4)
179
88
189
374
26
130
221
77
368
148
87
(A)
(5)
2.25
1.95
2.28
2.57
1.42
2.12
2.34
1.88
2.57
2.17
1.94
Flux
Correction
(6)
0
0
0
0
0
0
0
0
0
0
0
Table 6: NGC 613 - 12 H ii regions detected. H ii regions # 2 and 11 are extraneous to the
ring.
H ii region
Integrated Flux
(10−13 erg s−1cm−2)
Luminosity
(1039 erg s−1)
(1)
1
2
3
4
5
6
7
8
10
11
12
13
(2)
1.34
1.34
0.21
0.05
0.40
0.16
0.15
0.18
0.70
0.51
0.45
0.39
(3)
5.69
5.67
0.88
0.23
1.71
0.67
0.63
0.75
2.95
2.17
1.91
1.64
EW log(EW)
(A)
(4)
107
56
47
7
118
33
25
37
240
155
158
56
(A)
(5)
2.03
1.75
1.67
0.85
2.07
1.52
1.39
1.57
2.38
2.19
2.20
1.74
Flux
Correction
(6)
0
0
1
1
1
1
1
1
1
1
1
0
Table 7: NGC 1300 - 12 H ii regions detected associated with the ring. H ii region # 9 is the
nucleus.
-- 27 --
H ii region
Integrated Flux
(10−13 erg s−1cm−2)
Luminosity
(1039 erg s−1)
(1)
1
2
3
4
5
6
7
8
9
10
11
12
16
17
18
19
20
21
22
23
(2)
8.24
10.2
4.73
2.14
3.88
1.36
1.44
18.9
0.36
0.65
1.37
3.95
2.55
4.61
3.22
4.12
0.72
4.84
3.77
1.52
(3)
85.8
106
49.2
22.3
40.4
14.2
15
196
3.79
6.73
14.3
41.1
26.5
48
33.6
42.9
7.53
50.4
39.3
15.8
EW log(EW)
(A)
(4)
5364
744
330
247
650
396
400
5845
33
80
401
737
968
373
1340
1339
117
2513
479
305
(A)
(5)
3.73
2.87
2.52
2.39
2.81
2.60
2.60
3.77
1.52
1.91
2.60
2.87
2.99
2.57
3.13
3.13
2.07
3.40
2.68
2.48
Flux
Correction
(6)
1
0
0
0
0
1
1
0
1
1
1
0
1
0
1
0
1
1
0
1
Table 8: NGC 1343 - 20 H ii regions detected associated with the ring. H ii regions # 13 and
14 comprise the nucleus; H ii region # 15 is extraneous to the ring.
-- 28 --
H ii region
Integrated Flux
(10−13 erg s−1cm−2)
Luminosity
(1039 erg s−1)
(1)
1
2
3
4
7
11
14
15
17
(2)
6.79
13.4
1.46
1.05
0.52
0.52
1.84
0.98
1.55
(3)
109
215
23.4
16.8
8.32
8.31
29.5
15.8
24.8
EW log(EW)
(A)
(4)
1895
630
243
232
66
854
274
1951
3123
(A)
(5)
3.28
2.80
2.38
2.37
1.82
2.93
2.44
3.29
3.49
Flux
Correction
(6)
0
0
0
0
0
1
0
1
1
Table 9: NGC 1530 - 9 H ii regions detected associated with the ring. H ii regions # 8, 9,
10, 12, 13, and 16 comprise the nucleus; H ii regions # 5 and 6 are extraneous to the ring.
H ii region
Integrated Flux
(10−13 erg s−1cm−2)
Luminosity
(1039 erg s−1)
(1)
2
3
4
5
6
7
8
(2)
25.7
17.8
1.33
6.82
1.35
9.61
1.14
(3)
71.1
49.2
3.67
18.9
3.74
26.6
3.14
EW log(EW)
(A)
(4)
99
112
367
624
74
1676
106
(A)
(5)
2.00
2.05
2.57
2.80
1.87
3.22
2.03
Flux
Correction
(6)
0
0
0
0
0
0
0
Table 10: NGC 4303 - 7 H ii regions detected associated with the ring. H ii region # 1
comprises the nucleus.
-- 29 --
H ii region
Integrated Flux
(10−13 erg s−1cm−2)
Luminosity
(1039 erg s−1)
(1)
1
2
3
4
5
7
8
9
10
11
(2)
3.32
3.78
0.39
0.23
2.13
2.82
0.12
0.54
0.42
0.46
(3)
3.74
4.25
0.44
0.26
2.40
3.18
0.14
0.61
0.48
0.52
EW log(EW)
(A)
(4)
529
2607
111
148
375
1801
344
211
1088
1288
(A)
(5)
2.72
3.42
2.05
2.17
2.57
3.26
2.54
2.32
3.04
3.11
Flux
Correction
(6)
0
0
0
0
0
0
0
0
0
0
Table 11: NGC 4313 - 10 H ii regions detected associated with the ring. H ii region # 6
comprises the nucleus.
-- 30 --
H ii region
Integrated Flux
(10−13 erg s−1cm−2)
Luminosity
(1039 erg s−1)
(1)
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
(2)
27.8
4.08
1.23
0.92
1.82
6.18
2.88
1.06
3.76
1.87
4.93
2.03
7.44
14
0.57
1.86
1.51
0.88
1.82
(3)
171
25.2
7.58
5.68
11.2
38.1
17.8
6.53
23.2
11.5
30.4
12.5
45.9
86.1
3.52
11.5
9.33
5.40
11.2
EW log(EW)
(A)
(4)
1106
259
206
1277
6790
121
222
49
6378
2275
485
539
191
965
78
153
3125
1145
179
(A)
(5)
3.04
2.41
2.31
3.11
3.83
2.08
2.35
1.69
3.80
3.36
2.69
2.73
2.28
2.98
1.89
2.19
3.49
3.06
2.25
Flux
Correction
(6)
0
0
0
1
1
0
0
0
0
0
0
0
0
0
0
0
1
1
0
Table 12: NGC 5248 - 19 H ii regions detected associated with the ring. H ii region # 1
comprises the nucleus.
-- 31 --
H ii region
Integrated Flux
(10−13 erg s−1cm−2)
Luminosity
(1039 erg s−1)
(1)
3
4
5
6
7
8
9
10
11
12
13
14
15
17
18
19
21
22
(2)
2.28
0.27
2.13
2.19
3.87
0.21
0.70
0.16
0.30
0.17
0.22
0.62
1.11
4.09
3.52
1.53
0.24
0.39
(3)
48.6
5.76
45.4
46.7
82.5
4.48
15
3.53
6.32
3.58
4.78
13.3
23.6
87.2
75
32.6
5.10
8.38
EW log(EW)
(A)
(4)
1164
163
7961
248
556
170
1516
109
479
111
226
1218
3084
211
1426
1310
182
455
(A)
(5)
3.07
2.21
3.90
2.40
2.74
2.23
3.18
2.04
2.68
2.04
2.35
3.09
3.49
2.32
3.15
3.12
2.26
2.66
Flux
Correction
(6)
0
0
1
0
0
1
1
1
1
1
1
1
1
0
0
0
1
0
Table 13: NGC 5728 - 18 H ii regions detected associated with the ring. H ii regions # 1 and
2 comprise the nucleus; H ii regions # 16 and 20 are extraneous to the ring.
-- 32 --
H ii region
Integrated Flux
(10−13 erg s−1cm−2)
Luminosity
(1039 erg s−1)
(1)
1
2
3
4
5
6
7
8
9
(2)
2.39
2.63
0.40
0.67
0.41
0.38
0.25
3.27
1.75
(3)
58.5
64.3
9.80
16.3
10
9.37
6.20
79.9
42.7
EW log(EW)
(A)
(4)
77
189
97
2870
177
154
199
87
132
(A)
(5)
1.88
2.28
1.99
3.46
2.25
2.19
2.30
1.94
2.12
Flux
Correction
(6)
0
0
0
1
0
0
0
0
0
Table 14: NGC 5905 - 9 H ii regions detected associated with the ring. No nuclear or
extraneous H ii regions detected.
H ii region
Integrated Flux
(10−13 erg s−1cm−2)
Luminosity
(1039 erg s−1)
(1)
2
3
4
(2)
3.84
3.82
0.93
(3)
249
248
60.2
EW log(EW)
(A)
(4)
15
16
29
(A)
(5)
1.17
1.21
1.46
Flux
Correction
(6)
0
0
0
Table 15: NGC 5945 - 3 H ii regions detected associated with the ring. H ii region # 1
comprises the nucleus.
-- 33 --
H ii region
Integrated Flux
(10−13 erg s−1cm−2)
Luminosity
(1039 erg s−1)
(1)
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
(2)
11.6
9.84
24.9
2.24
1.22
6.38
2.56
5.41
4.35
2.19
3.22
0.43
0.74
2.01
0.45
2.57
1.86
3.25
3.63
7.02
0.33
0.46
(3)
152
128
324
29.2
15.9
83.2
33.3
70.5
56.7
28.5
42
5.57
9.67
26.1
5.82
33.5
24.3
42.4
47.4
91.5
4.30
5.99
EW log(EW)
(A)
(4)
152
390
541
652
247
614
591
2162
784
94
1013
30
88
4
4
130
50
154
1374
866
23
48
(A)
(5)
2.18
2.59
2.73
2.81
2.39
2.79
2.77
3.33
2.89
1.97
3.01
1.48
1.94
0.61
0.55
2.12
1.70
2.19
3.14
2.94
1.37
1.68
Flux
Correction
(6)
0
0
0
1
1
0
0
1
0
0
1
1
1
0
0
0
0
0
1
0
1
1
Table 16: NGC 5953 - 22 H ii regions detected associated with the ring. H ii region # 1
comprises the nucleus.
-- 34 --
H ii region
Integrated Flux
(10−13 erg s−1cm−2)
Luminosity
(1039 erg s−1)
(1)
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
(2)
38.6
7.12
1.08
2.28
1.55
1.90
3.43
2.08
4.07
0.94
10.7
2.60
2.64
1.13
0.55
0.59
1.00
10.5
14.8
12.3
4.66
0.87
2.64
1.32
9.70
1.22
0.95
5.72
2.87
2.32
6.78
2.18
3.92
11.8
33
(3)
17.2
3.17
0.48
1.01
0.69
0.84
1.53
0.93
1.81
0.42
4.76
1.16
1.18
0.50
0.25
0.26
0.45
4.66
6.59
5.46
2.08
0.39
1.18
0.59
4.32
0.54
0.42
2.55
1.28
1.03
3.02
0.97
1.75
5.26
14.7
EW log(EW)
(A)
(4)
3686
804
45
151
52
67
224
98
307
26
1715
214
120
40
12
13
31
796
1123
2072
304
22
190
70
1271
37
21
431
148
92
808
132
323
1774
6628
(A)
(5)
3.57
2.91
1.65
2.18
1.71
1.83
2.35
1.99
2.49
1.42
3.23
2.33
2.08
1.60
1.08
1.11
1.49
2.90
3.05
3.32
2.48
1.34
2.28
1.85
3.10
1.57
1.32
2.63
2.17
1.96
2.91
2.12
2.51
3.25
3.82
Flux
Correction
(6)
0
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
0
0
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
Table 17: NGC 6503 - 84 H ii regions detected associated with the ring. H ii regions # 39,
44, and 45 comprise the nucleus; H ii regions # 1, 2, and 86 are extraneous to the ring.
-- 35 --
H ii region
Integrated Flux
(10−13 erg s−1cm−2)
Luminosity
(1039 erg s−1)
(1)
38
40
41
42
43
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
66
67
68
69
70
71
72
73
74
75
76
77
(2)
5.41
5.55
9.64
7.17
2.88
9.87
1.36
1.30
3.55
1.85
1.19
3.29
3.95
2.51
2.97
2.92
10.3
6.08
7.01
1.16
2.40
2.72
20.8
0.75
1.17
8.01
0.62
1.49
37.2
1.54
2.34
1.57
1.49
1.69
2.83
1.65
0.36
(3)
2.41
2.47
4.29
3.19
1.28
4.39
0.61
0.58
1.58
0.82
0.53
1.47
1.76
1.12
1.32
1.30
4.57
2.71
3.12
0.51
1.07
1.21
9.27
0.34
0.52
3.57
0.28
0.66
16.6
0.69
1.04
0.70
0.66
0.75
1.26
0.74
0.16
Table 17: NGC 6503 - continued.
EW log(EW)
(A)
(4)
584
570
1331
661
240
1363
51
49
299
66
40
278
337
191
129
172
1594
657
682
44
150
16
2839
19
47
1106
14
65
7731
51
110
71
49
60
174
75
10
(A)
(5)
2.77
2.76
3.12
2.82
2.38
3.13
1.71
1.69
2.48
1.82
1.60
2.44
2.53
2.28
2.11
2.24
3.20
2.82
2.83
1.64
2.18
1.19
3.45
1.28
1.67
3.04
1.13
1.81
3.89
1.71
2.04
1.85
1.69
1.78
2.24
1.87
1.00
Flux
Correction
(6)
1
1
1
1
1
1
1
1
1
0
1
1
1
1
1
1
1
1
1
1
1
0
0
1
1
1
1
1
1
1
1
1
1
1
1
1
1
-- 36 --
H ii region
Integrated Flux
(10−13 erg s−1cm−2)
Luminosity
(1039 erg s−1)
(1)
78
79
80
81
82
83
84
85
87
88
89
90
(2)
1.80
2.94
3.84
3.86
1.16
1.38
4.53
1.27
2.03
4.13
2.45
1.14
(3)
0.80
1.31
1.71
1.72
0.52
0.61
2.02
0.57
0.90
1.84
1.09
0.51
Table 17: NGC 6503 - continued.
EW log(EW)
(A)
(4)
78
236
312
219
62
81
483
45
117
356
106
35
(A)
(5)
1.89
2.37
2.49
2.34
1.79
1.91
2.68
1.65
2.07
2.55
2.03
1.55
Flux
Correction
(6)
1
1
1
0
1
1
1
1
1
1
1
1
-- 37 --
H ii region
Integrated Flux
(10−13 erg s−1cm−2)
Luminosity
(1039 erg s−1)
(1)
2
3
4
5
6
7
8
9
(2)
7.63
0.60
2.30
0.87
8.92
1.36
2.30
1.38
(3)
53
4.15
16
6.08
62
9.45
16
9.60
EW log(EW)
(A)
(4)
58
120
810
95
3034
154
103
37
(A)
(5)
1.77
2.08
2.91
1.98
3.48
2.19
2.01
1.56
Flux
Correction
(6)
0
1
1
0
0
0
0
0
Table 18: NGC 6951 - 8 H ii regions detected associated with the ring. H ii region # 1
comprises the nucleus.
-- 38 --
H ii region
Integrated Flux
(10−13 erg s−1cm−2)
Luminosity
(1039 erg s−1)
(1)
2
3
6
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
(2)
1.70
1.06
0.43
0.65
0.73
0.50
1.84
0.79
0.61
10.3
0.40
0.26
0.76
0.21
0.52
0.64
0.62
0.23
1.06
0.77
0.31
1.19
(3)
5.20
3.25
1.32
1.99
2.23
1.54
5.64
2.43
1.86
31.5
1.23
0.80
2.32
0.64
1.59
1.95
1.90
0.71
3.25
2.37
0.94
3.64
EW log(EW)
(A)
(4)
88
542
118
224
306
153
987
263
216
56
105
42
262
33
143
137
189
48
478
246
63
477
(A)
(5)
1.94
2.73
2.07
2.35
2.49
2.18
2.99
2.42
2.33
1.74
2.02
1.62
2.42
1.52
2.16
2.14
2.28
1.68
2.68
2.39
1.80
2.68
Flux
Correction
(6)
0
1
1
1
1
1
1
1
1
0
1
1
0
1
1
0
1
1
1
1
1
1
Table 19: NGC 7217 - 22 H ii regions detected associated with the ring. H ii regions # 1 and
4 comprise the nucleus; H ii regions # 5 and 7 are extraneous to the ring.
-- 39 --
H ii region
Integrated Flux
(10−13 erg s−1cm−2)
Luminosity
(1039 erg s−1)
(1)
1
2
3
4
5
6
9
10
11
12
13
(2)
0.49
1.11
0.24
0.65
0.91
0.31
2.21
3.38
1.30
0.95
0.39
(3)
6.77
15.2
3.27
8.96
12.4
4.30
30.2
46.2
17.8
13
5.31
EW log(EW)
(A)
(4)
56
1054
71
65
685
86
582
234
43
244
49
(A)
(5)
1.75
3.02
1.85
1.81
2.84
1.94
2.77
2.37
1.63
2.39
1.69
Flux
Correction
(6)
0
1
1
0
1
1
0
0
0
0
0
Table 20: IC 1438 - 11 H ii regions detected associated with the ring. H ii regions # 7 and
8 comprise the nucleus.
-- 40 --
H ii region
Integrated Flux
(10−13 erg s−1cm−2)
Luminosity
(1039 erg s−1)
(1)
1
2
3
4
5
6
7
8
9
14
15
16
17
18
19
20
21
22
(2)
0.49
0.25
0.35
0.41
0.16
0.77
0.44
0.48
1.43
0.17
0.31
0.12
0.21
0.38
0.30
0.25
0.14
0.80
(3)
12
6.11
8.46
10
3.85
18.9
10.7
11.7
35
4.26
7.49
3.04
5.18
9.42
7.45
6.08
3.38
19.6
EW log(EW)
(A)
(4)
446
234
410
439
93
915
343
374
3371
73
203
74
199
302
146
217
92
1009
(A)
(5)
2.65
2.37
2.61
2.64
1.97
2.96
2.54
2.57
3.53
1.86
2.31
1.87
2.30
2.48
2.17
2.34
1.96
3.00
Flux
Correction
(6)
0
1
1
1
1
1
1
1
1
1
1
1
1
1
0
0
1
1
Table 21: NGC 7570 - 18 H ii regions detected associated with the ring. H ii regions # 10,
11, 12, and 13 comprise the nucleus.
-- 41 --
H ii region
Integrated Flux
(10−13 erg s−1cm−2)
Luminosity
(1039 erg s−1)
(1)
1
2
3
4
5
6
7
8
10
11
12
13
14
15
17
18
19
20
21
22
23
24
25
26
29
30
32
33
34
35
(2)
3.14
1.53
1.38
0.75
0.30
0.52
0.18
0.31
0.5
0.84
0.40
0.31
1.20
0.10
2.98
0.95
3.42
0.63
0.83
0.64
2.06
0.20
0.56
0.50
0.29
0.69
0.27
0.45
3.28
0.59
(3)
42.6
20.7
18.8
10.2
4.13
7.02
2.48
4.24
6.80
11.4
5.40
4.19
16.4
1.41
40.6
13
46.5
8.57
11.3
8.71
28
2.76
7.64
6.83
3.93
9.32
3.64
6.08
44.6
8.08
EW log(EW)
(A)
(4)
1516
684
436
1259
320
1643
304
526
1478
2556
643
694
4475
100
1840
2904
20204
316
102
2039
1135
413
1146
1523
519
1100
575
1323
23884
1756
(A)
(5)
3.18
2.84
2.64
3.10
2.51
3.22
2.48
2.72
3.17
3.41
2.81
2.84
3.65
2.00
3.26
3.46
4.31
2.50
2.01
3.31
3.06
2.62
3.06
3.18
2.71
3.04
2.76
3.12
4.38
3.24
Flux
Correction
(6)
0
0
0
0
0
1
1
0
1
0
0
1
1
1
0
0
1
0
0
1
0
1
0
1
1
0
1
1
1
1
Table 22: NGC 7716 - 30 H ii regions detected associated with the ring. H ii region # 16
comprises the nucleus; H ii regions # 9, 27, 28, and 31 are extraneous to the ring.
-- 42 --
H ii region
Integrated Flux
(10−13 erg s−1cm−2)
Luminosity
(1039 erg s−1)
(1)
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
(2)
1.70
6.64
6.59
3.61
1.08
1.84
0.88
12.4
0.61
1.84
0.81
2.28
0.99
5.07
1.45
0.70
0.39
3.09
1.68
5.59
1.05
2.55
2.51
1.63
0.27
2.83
1.14
0.94
1.17
2.76
0.42
0.58
2.00
0.54
(3)
10
39.2
38.9
21.3
6.37
10.9
5.19
73.2
3.59
10.9
4.80
13.5
5.82
29.9
8.54
4.15
2.30
18.2
9.90
32.9
6.21
15
14.8
9.59
1.59
16.7
6.75
5.54
6.88
16.3
2.47
3.45
11.8
3.19
EW log(EW)
(A)
(4)
254
352
484
524
284
496
236
2057
129
696
151
1003
347
674
507
191
55
464
575
3721
317
254
274
626
27
677
342
244
234
1300
84
122
811
116
(A)
(5)
2.41
2.55
2.68
2.72
2.45
2.70
2.37
3.31
2.11
2.84
2.18
3.00
2.54
2.83
2.70
2.28
1.74
2.67
2.76
3.57
2.50
2.40
2.44
2.80
1.43
2.83
2.53
2.39
2.37
3.11
1.92
2.09
2.91
2.06
Flux
Correction
(6)
0
0
0
0
1
0
1
0
1
1
1
1
1
0
1
1
1
0
0
1
1
0
0
1
1
0
1
1
1
1
1
1
1
1
Table 23: NGC 7742 - 38 H ii regions detected associated with the ring. H ii region # 18
comprises the nucleus.
-- 43 --
H ii region
Integrated Flux
(10−13 erg s−1cm−2)
Luminosity
(1039 erg s−1)
(1)
36
37
38
39
(2)
5.52
1.22
0.38
4.75
(3)
32.6
7.19
2.25
28
EW log(EW)
(A)
(4)
3043
390
68
2633
(A)
(5)
3.48
2.59
1.83
3.42
Flux
Correction
(6)
1
1
1
1
Table 23: NGC 7742 continued.
-- 44 --
REFERENCES
Allard, E. L., Peletier, R. F., & Knapen, J. H. 2005, ApJ, 633, L25
Allard, E. L.,Knapen, J. H., Peletier, R. F., & Sarzi, M. 2006, submitted to MNRAS
Athanassoula, E. 2000, Stars, Gas and Dust in Galaxies: Exploring the Links, 221, 243
Athanassoula, E. 1992, MNRAS, 259, 345
Benedict, G. F., et al. 1993, AJ, 105, 1369
Bertin, E., Arnouts, S. 1996, A&A, S117, 393
Blackman, C. P. 1981, MNRAS, 195, 451
Bottema, R. 1989, A&A, 221, 236
Buta, R. & Combes, F., 1996, Fund. of Cosmic Phys. 17, 95
Buta, R. & Crocker, D. A. 1993, AJ, 105, 1344
Combes, F., & Gerin, M. 1985, A&A, 150, 327
Combes, F., et al. 2004, A&A, 414, 857
Davies, R. I., Tacconi, L. J., Genzel, R. 2004, ApJ, 602, 148
D´ıaz-Santos, T., Alonso-Herrero, A., Colina, L., Ryder, S. D., Knapen J. H. 2007,ApJ, 661,
149
Elmegreen, B. 1996, IAU Colloq. 157: Barred Galaxies, 91, 197
Elmegreen, B. G. 2002, Modes of Star Formation and the Origin of Field Populations, 285,
425
Elmegreen, B. 1997, RMxAC, 6, 165
Garc´ıa-Burillo, S., Combes, F., Schinnerer, E., Boone, F., & Hunt, L. K. 2005, A&A, 441,
1011
Genzel, R., Weitzel, L., Tacconi-Garman, L. E., Blietz, M., Cameron, M., Krabbe, A., Lutz,
D., & Sternberg, A. 1995, ApJ, 444, 129
Helfer, T. T., Thornley, M. D., Regan, M. W., Wong, T., Sheth, K., Vogel, S. N., Blitz, L.,
& Bock, D. C.-J. 2003, ApJS, 145, 259
-- 45 --
Heller, C. H., & Shlosman, I. 1994, ApJ, 424, 84
Heller, C. H., & Shlosman, I. 1996, ApJ, 471, 143
Hummel, E., van der Hulst, J. M., & Keel, W. C. 1987, A&A, 172, 32
Jogee, S., Shlosman, I., Laine, S., Englmaier, P., Knapen, J. H., Scoville, N., & Wilson,
C. D. 2002, ApJ, 575, 156
Jogee, S., Scoville, N., & Kenney, J. D. P. 2005, ApJ, 630, 837
Kenney, J. D. P., Wilson, C. D., Scoville, N. Z., Devereux, N. A., & Young, J. S. 1992, ApJ,
395, L79
Kenney, J. D. P., Carlstrom, J. E., & Young, J. S. 1993, ApJ, 418, 687
Kennicutt, R. C. 1998, ARA&A, 36, 189
Knapen, J.H., Beckman, J. E., Heller, C. H., Shlosman, I., de Jong, R.S. 1995a, ApJ, 454,
623
Knapen, J. H., Beckman, J. E., Shlosman, I., Peletier, R. F., Heller, C. H., & de Jong, R. S.
1995b, ApJ, 443, L73
Knapen, J. H. 1999, The Evolution of Galaxies on Cosmological Timescales, 187, 72
Knapen, J. H., Laine, S., & Relano, M. 1999, Ap&SS, 269, 605
Knapen, J. H., Shlosman, I., Heller, C. H., Rand, R. J., Beckman, J. E., & Rozas, M. 2000,
ApJ, 528, 219
Knapen, J. H., Whyte, L. F., de Blok, W. J. G., & van der Hulst, J. M. 2004, A&A, 423,
481
Knapen, J. H. 2005, A&A, 429, 141
Knapen,J.H., Mazzuca, L.M., Boker, T, Shlosman, I., Colina, L. Combes, F., Axon, D.J.
2006, A&A, 448, 489
Kormendy, J., Kennicutt, R. C. Jr. 2004 ARA&A, 42, 603
Kroupa, P. 2002, Modes of Star Formation and the Origin of Field Populations, 285, 86
Leitherer, C., et al. 1999, ApJS, 123, 3
-- 46 --
Maciejewski, W., Teuben, P. J., Sparke, L. S., & Stone, J. M. 2002, MNRAS, 329, 502
Maoz, D., Barth, A. J., Ho, L. C., Sternberg, A., & Filippenko, A. V. 2001, AJ, 121, 3048
Martinet, L. 1995, Fundamentals of Cosmic Physics, 15, 341
Moiseev, A. V., Vald´es, J. R., & Chavushyan, V. H. 2004, A&A, 421, 433
Oke,J., 1990 AJ 99,5
Osmer, P. S., Smith, M. G., & Weedman, D. W. 1974, ApJ, 192, 279
Peterson, C. J., Roberts, M. S., Rubin, V. C., & Ford, W. K. 1978, ApJ, 226, 770
Phillips, A. C. 1996, ASP Conf. Ser. 91: IAU Colloq. 157: Barred Galaxies, 91, 44
Piner, B., Stone, J., Teuben, P. 1995, ApJ,449, 508P
Pogge, R. W. 1989, ApJ, 345, 730
Regan, M. W., Teuben, P. J., Vogel, S. N., & van der Hulst, T. 1996, AJ, 112, 2549
Regan, M. W., & Teuben, P. J. 2004, ApJ, 600, 595
Regan, M. W., Vogel, S. N., & Teuben, P. J. 1997, ApJ, 482, L143
Regan, M. W., Teuben, P. 2003 APJ, 582, 723
Rubin, V. C. 1980, ApJ, 238, 808
Rubin, V. C., Kenney, J. D. P., & Young, J. S. 1997, AJ, 113, 1250
Ryder, S. D., Knapen, J. H., & Takamiya, M. 2001, MNRAS, 323, 663
Sakamoto, K., Okumura, S. K., Ishizuki, S., & Scoville, N. Z. 1999, ApJ, 525, 691
Salpeter, E. E. 1955, ApJ, 121, 161
Sarzi, M., Allard, E.L., Knapen, J. H., Mazzuca, L. M., MNRAS in press
Schwarz, M. P. 1981, ApJ, 247, 77
Shlosman, I., Begelman, M. C., & Frank, J. 1990, Nature, 345, 679
Sheth, K., Regan, M. W., Vogel, S. N., & Teuben, P. J. 2000, ApJ, 532, 221
Smith, D. A., Herter, T., Haynes, M. P., & Neff, S. G. 1999, ApJ, 510, 669
-- 47 --
Storchi-Bergmann, T., Wilson, A. S., & Baldwin, J. A. 1996, ApJ, 460, 252
Tully, R. B. 1988 British Astronomical Association Journal, 98, 6
This preprint was prepared with the AAS LATEX macros v5.2.
-- 48 --
Fig. 1. -- panels a. - l.: Ellipse fits for the nuclear rings in the sample galaxies produced by
the IRAF task ellipse, overlaid on the Hα images. The major axis lengths were optimized
based on the maximum intensity along a given contour as computed using ellipse (see
Section 3).
-- 49 --
Fig. 1. -- continued: panels m. - v.
-- 50 --
Fig. 2. -- (a) PA of the disk versus the ring major axis for those rings whose ellipticity is
greater than or equal to 0.1. (b) Ellipticity comparison between the ring and the galactic disk.
Both plots show a linear relationship between the ring and host galaxy PA and ellipticity,
and indicate that the rings are in the same plane as the disk and are nearly circular.
-- 51 --
Fig. 3. -- panels a. - l.: Results from SExtractor. Each H ii region identified is fitted using
a Kron ellipse, which calculates the total magnitude of the extended H ii region. In the cases
where the Hα boundaries are merged from one or more sources, the program deblends the
regions and indicates these fittings with a dashed ellipse. SExtractor could not identify
any H ii regions for NGC 7469 due to poor spatial resolution.
-- 52 --
-- 53 --
Fig. 4. -- Change in ring morphology for NGC 1343 based on variation of the background
box size and detection threshold value. Top row shows the resolved H ii regions based on
fixing the threshold at 2σ and varying the box size - a) box size = 4x4 pixels with 31 H ii
region detections; b) box size = 12x12 pixels with 20 detections. The center image is the
reference image: box size = 8x8 pixels with 25 detections. Bottom row shows the images
based on fixing the box size at 8x8 pixels and varying the threshold value - c) threshold
= 1.33σ with 30 detections; d) threshold = 3σ with 19 detections. The optimal solution
was derived iteratively and based on comparison of the general size, shape, and quantity
of the regions that could be easily detected by eye in the Halpha image. The grayscale is
not representative of the ages of the regions, but rather the order in which SExtractor
identified the regions.
-- 54 --
Fig. 5. -- Age distribution -- a. through h. The left panels plot the PA of each H ii region vs.
their EW. Solid diamonds represent those EWs whose I-band fluxes were not affected by the
background 3σ noise; open diamonds indicate EW where I-band fluxes needed to be increased
to compensate for the dominant background noise; open diamonds with crosses show those
H ii regions that would not have needed their flux modified if using a 2σ constraint. The
right panels picturize the rings with their identified H ii regions. Age decreases as the shade
of color changes from light to dark. Arrows are drawn to indicate the PA of the bar major
axis, as well as the rotational direction of flow in the ring (when a bar exists). The ring major
axis is indicated by two dashed lines. All rings have been rotated to a typical North-East
configuration. See Section 6.1 for details. NGC 4321 is based on Hβ analysis from Allard et
al. (2006).
-- 55 --
-- 56 --
-- 57 --
Fig. 6. -- Hα EW vs.
(1999; Starburst99 curves) for an
instantaneous and a continuous burst of star formation with a Salpeter IMF between Mlow =
1 M⊙ and Mup = 100 M⊙. A solar metallicity of Z = 0.02 is used.
time following Leitherer et al.
-- 58 --
time following Leitherer et al.
Fig. 7. -- Hα EW vs.
(1999; Starburst99 curves) for an
instantaneous burst of star formation (left panel) and continuous burst (right panel) for
varying IMF slopes, upper limit masses, and metalicities. The thick solid line represents
the model prediction we adopted for IMF between Mlow = 1 M⊙ and Mup = 100 M⊙, IMF
slope= 2.35, and Z = 0.02. Long dashed lines represent a slope of 3.3 and Mup = 100 M⊙
for Z=0.04 and Z=0.008. Short dashed lines represent a slope of 2.35 and Mup = 30 M⊙ for
Z=0.04 and Z=0.008.
-- 59 --
Fig. 8. -- panels a. through l.: B -I images with indicators of the bar PA and contact point
(CP) PA. The CP PA is defined as a 90◦ offset from the bar PA. In those images with a
bar, the blue line (the longer line if viewing in black and white) extends to approximately
half the bar major axis size (and traverses the entire FOV in some cases where the ring was
magnified such that the bar cannot be fully seen); the the green line (shorter line drawn to
the approximate ring diameter) represents the CP PAs, which are assumed to be 90 ◦ offsets
of the bar PA.
-- 60 --
-- 61 --
Fig. 9. -- Integrated SFR of the rings. The SFR is compared to the presence (left panel)
or absence (right panel) of an age gradient. The mean SFR values are 2.2±0.7 M⊙ yr−1
(uncertainty is standard error) for the galaxies with and 3.6±1.1 M⊙ yr−1 for those without
gradients.
Fig. 10. -- Histogram showing the distribution of angular offsets between the bar major axis
PA and the locations of the youngest Hii region in each hemisphere, which is bisected by
the bar. A Gaussian fit to the distribution is overlaid, and shows a peak at an offset of 90◦,
which is consistent with the location of the contact points. See Table 2 for more details.
|
0708.2576 | 2 | 0708 | 2007-08-24T02:09:54 | The Nature of the Hard-X-Ray Emitting Symbiotic Star RT Cru | [
"astro-ph"
] | We describe Chandra High-Energy Transmission Grating Spectrometer observations of RT Cru, the first of a new sub-class of symbiotic stars that appear to contain white dwarfs (WDs) capable of producing hard X-ray emission out to greater than 50 keV. The production of such hard X-ray emission from the objects in this sub-class (which also includes CD -57 3057, T CrB, and CH Cyg) challenges our understanding of accreting WDs. We find that the 0.3 -- 8.0 keV X-ray spectrum of RT Cru emanates from an isobaric cooling flow, as in the optically thin accretion-disk boundary layers of some dwarf novae. The parameters of the spectral fit confirm that the compact accretor is a WD, and they are consistent with the WD being massive. We detect rapid, stochastic variability from the X-ray emission below 4 keV. The combination of flickering variability and a cooling-flow spectrum indicates that RT Cru is likely powered by accretion through a disk. Whereas the cataclysmic variable stars with the hardest X-ray emission are typically magnetic accretors with X-ray flux modulated at the WD spin period, we find that the X-ray emission from RT Cru is not pulsed. RT Cru therefore shows no evidence for magnetically channeled accretion, consistent with our interpretation that the Chandra spectrum arises from an accretion-disk boundary layer. | astro-ph | astro-ph | Draft version October 27, 2018
Preprint typeset using LATEX style emulateapj v. 08/22/09
7
0
0
2
g
u
A
4
2
]
h
p
-
o
r
t
s
a
[
2
v
6
7
5
2
.
8
0
7
0
:
v
i
X
r
a
THE NATURE OF THE HARD-X-RAY EMITTING SYMBIOTIC STAR RT CRU
Instituto de Astronomia, Geof´ısica e Ciencias Atmosfericas, Universidade de Sao Paulo, Rua do Matao 1226, Cid. Universitaria, Sao
G. J. M. Luna
Paulo, Brazil 05508-900
and
J. L. Sokoloski
Columbia Astrophysics Lab. 550 W120th St., 1027 Pupin Hall, Columbia University, New York, New York 10027, USA
Draft version October 27, 2018
ABSTRACT
We describe Chandra High-Energy Transmission Grating Spectrometer observations of RT Cru,
the first of a new sub-class of symbiotic stars that appear to contain white dwarfs (WDs) capable
of producing hard X-ray emission out to greater than 50 keV. The production of such hard X-ray
emission from the objects in this sub-class (which also includes CD −57 3057, T CrB, and CH Cyg)
challenges our understanding of accreting WDs. We find that the 0.3 -- 8.0 keV X-ray spectrum of RT
Cru emanates from an isobaric cooling flow, as in the optically thin accretion-disk boundary layers
of some dwarf novae. The parameters of the spectral fit confirm that the compact accretor is a WD,
and they are consistent with the WD being massive. We detect rapid, stochastic variability from the
X-ray emission below 4 keV. The combination of flickering variability and a cooling-flow spectrum
indicates that RT Cru is likely powered by accretion through a disk. Whereas the cataclysmic variable
stars with the hardest X-ray emission are typically magnetic accretors with X-ray flux modulated at
the WD spin period, we find that the X-ray emission from RT Cru is not pulsed. RT Cru therefore
shows no evidence for magnetically channeled accretion, consistent with our interpretation that the
Chandra spectrum arises from an accretion-disk boundary layer.
Subject headings: binary stars: general -- white dwarf: accretion -- X-rays
1. INTRODUCTION
Symbiotic stars are interacting binaries in which a hot,
compact star accretes from the wind of a red-giant com-
panion. Although a few symbiotics contain neutron-
star accretors (e.g.,GX 1+4, 4U 1700+24, 4U 1954+31,
IGR J16194-2810; Davidsen et al. 1977; Masetti et al.
2002; Galloway et al. 2002; Masetti et al. 2007a,b), the
accreting compact object is usually a white dwarf (WD).
Typical binary separations are on the order of AU, with
orbital periods on the order of a few hundred days to a
few decades (Kenyon 1986; Belczy´nski et al. 2000). Sym-
biotics can thus be thought of as very large cousins of
cataclysmic variables (CVs). The accretion rate onto the
WD appears to be high enough in most symbiotic sys-
tems that accreted material is burned quasi-steadily in a
shell on the WD surface, producing a high UV luminos-
ity (Sokoloski et al. 2001; Orio et al. 2007). Although
accretion disks are likely to exist around the WDs in
some symbiotics (Livio 1997; Sokoloski & Kenyon 2003),
there is little direct evidence for these disks. Finally, the
red-giant wind produces a dense nebula that surrounds
the system.
Most symbiotic stars with detectable X-ray emis-
sion display soft, or supersoft, thermal X-ray spectra.
As in the supersoft X-ray sources, the lowest-energy
X-rays could emanate directly from material burning
quasi-steadily on the WD surface (Jordan et al. 1994;
Orio et al. 2007). Based on a survey of symbiotics with
ROSAT , Muerset et al. (1997) proposed that symbiotics
be classified according to the hardness of their X-ray
Electronic address: [email protected]
Electronic address: [email protected]
spectra. They labeled sources with supersoft spectra
α-types, sources with the slightly harder spectra likely
to arise from collision of the red-giant and white-dwarf
winds β-types, and systems with the hardest spectra
that might be indicative of neutron stars γ-types. Al-
though more recent observations using the broader X-ray
coverage and greater sensitivity of Chandra and XMM-
N ewton have shown that some symbiotics do not fit into
the simple α/β/γ classification scheme (e.g., Z And and
o Ceti; Sokoloski et al. 2006; Karovska et al. 2005), most
still appear to produce primarily soft X-rays (E < 3
keV).
With the advent of the sensitive hard X-ray detec-
tors on the Swif t and INTEGRAL satellites, a new pic-
ture has emerged. Some symbiotic stars can produce
X-ray emission out to greater than 50 keV. Such hard X-
ray emission has so far been detected from 4 symbiotics
thought to harbor WDs -- RT Cru (Chernyakova et al.
2005; Bird et al. 2007), T CrB (Tueller et al. 2005;
Luna et al. 2007), CH Cyg (Mukai et al. 2007) and CD -
57 3057 (Masetti et al. 2006; Bird et al. 2007). Although
the origin of this hard X-ray emission is not known, there
are some underlying similarities between these hard X-
ray emitting symbiotics. Unlike most other symbiotic
stars, they display a high incidence of optical flickering.
They also tend to have low optical line strengths, indi-
cating that they are often only "weakly" symbiotic. Both
the visibility of the optical flickering (which in most sym-
biotics is overwhelmed by reprocessed shell-burning emis-
sion; Sokoloski 2003) and the weakness or low-ionization-
state of the optical lines suggest that quasi-steady shell
burning is not taking place in these objects, either be-
cause: 1) the WD is more massive, or 2) the accre-
2
Luna & Sokoloski
tion rate is lower than in other symbiotics. Hard X-
ray emission might therefore be a proxy for high WD
mass. In fact, at least one of the hard-X-ray symbiotics,
T CrB, is a recurrent nova and contains a high-mass WD
(Hachisu & Kato 2001). Finally, jet production appears
to be more common in flickering symbiotics, and one
of the 4 hard X-ray symbiotics -- CH Cyg -- regularly
produces jets (Taylor et al. 1986; Karovska et al. 1998;
Crocker et al. 2001, 2002). Other WDs in hard X-ray
symbiotics might thus also harbor jets.
Cieslinski et al. (1994) classified RT Cru as a symbi-
otic star based on its optical spectrum. They noted
that the lack of strong high-ionization emission lines and
the very weak forbidden emission lines make the optical
spectrum similar to that of T CrB. They detected opti-
cal flickering in the U band with a time scale of a few
tens of minutes. Except for GX 1+4 (Jablonski et al.
1997; Chakrabarty & Roche 1997), none of the neutron-
star containing symbiotic stars produce optical flickering
(e.g. Sokoloski et al. 2001), which is common in CVs,
or show Balmer or He II emission lines (Masetti et al.
2007b). The presence of optical flickering and Balmer
and He II emission lines from RT Cru (Cieslinski et al.
1994) suggests that it therefore contains an accreting WD
rather than a neutron star. Reddening estimates from
optical spectra and infrared magnitudes (coupled with
the assumption that the radius of the M5 III red giant is
about 0.5 AU; van Belle et al. 1999) suggest that RT Cru
is between 1.5 and 2 kpc away (J. Miko lajewska, private
communication).
In 2003 and 2004, the IBIS instrument on board IN-
TEGRAL detected hard X-ray emission extending out
to ∼100 keV from the source IGR J12349-6434, which
Chernyakova et al. (2005) found to have a 20-60 keV flux
density of ∼3 mCrab. Masetti et al. (2005) suggested
an association between IGR J12349-6434 and RT Cru,
which observations with the Swif t satellite later con-
firmed (Tueller et al. 2005). The long-term optical light
curve from the AAVSO indicates that at the time of
the INTEGRAL observations, RT Cru was in an optical
bright state; it brightened from 13.5 to 11.5 mag some-
time between 1998 and 2000. Between 2000 and 2005,
the optical brightness slowly decreased to approximately
12.1 mag. The short (4.7 ks) Swift observation of 2005
August showed that between 2003 and 2005, the hard
X-ray flux also decreased.
In this paper, we describe Chandra High Energy
Transmission Grating (HETG) observations of RT Cru,
the first member of a new class of hard X-ray emitting
symbiotic WDs. We detail the observations and data re-
duction in §2 and the results from spectral and timing
analyses in §3. In §4, we discuss our interpretation of the
observations, which confirm that the accreting compact
object in RT Cru is a WD and provide some of the most
direct evidence to date for an accretion disk around a
wind-fed WD in a symbiotic system. In this section, we
also discuss the implications of a system that can accel-
erate particles to relativistic speeds and produce X-ray
emission out to greater than 50 keV being powered by
an accreting WD. We summarize our conclusions in §5.
2. OBSERVATIONS AND DATA REDUCTION
vation of RT Cru using the HETG (Canizares et al. 2005)
and the ACIS-S detector (ObsId 7186, start time 10:21:12
UT). We requested the DDT observation to attempt to
catch RT Cru in the optical bright state that appeared
to be associated with hard X-ray emission. We used the
HETG instrument because the Swift XRT observation
of 2005 August hinted at several possible emission-line
complexes. The data were collected in Timed Exposure
mode, in which the CCD chips were read out every 2.54 s.
The data were telemetered back to earth in Faint mode,
which conveys photon arrival times, event amplitudes,
and additional information for evaluating the validity
of each event. We reduced the data according to stan-
dard procedures using the software package CIAO 3.31.
We extracted a spectrum from the undispersed light (the
zeroth-order spot, which fell on the S3, back-illuminated
chip) using a circular extraction region with a radius of 6′′
centered on the source coordinates: α = 12h 34m 43.74s
and δ = −64◦ 33′56.0′′. To obtain the background for the
zeroth-order light, we extracted photons from a source-
free circular region on the same CCD. We grouped the
spectrum, which is shown in Fig. 1, to have at least 50
counts per bin. The average zeroth-order source count
rate was 0.11 c s−1.
For the dispersed light from both of the HETG sets of
gratings -- the Medium Energy Grating (MEG) and the
High Energy Grating (HEG) -- we extracted spectra from
each of the m = ±1, ±2, and ±3 orders individually (us-
ing the CIAO software tool dmtype2split). To obtain the
background for the dispersed light, we extracted counts
from rectangular regions on either side of the spectral
image. The count rate in the HEG and MEG m ± 1
orders was 0.042 c s−1 and 0.034 c s−1, respectively. Al-
though the dispersed spectral orders contained too few
counts to produce a high signal-to-noise-ratio spectrum
of lines spanning the full energy range of the instrument,
the combined m = ±1 spectrum (grouped at twice the
full width at half maximum of 0.012A) provided good-
quality data in the region around the Fe Kα emission-line
complex. We therefore used the the zeroth-order spec-
trum for continuum fitting and the dispersed (m = ±1)
spectra primarily for analysis of the Fe lines. The HEG
and MEG m = ±2 and ±3 spectra helped confirm the
Fe line identifications. For spectral fitting of both the
zeroth and higher order photons, we used the standard
software packages Xspec (Arnaud 1996) v12.3.0 and ISIS
(Houck 2002). The background contributed less than 1%
of the total extracted dispersed and undispersed light.
We generated light curves in the energy bands 0.3 --
4.0 keV and 4.0 -- 8.0 keV by extracting counts (with
CIAO) from a region containing the zeroth-order spot
and the m = ±1 dispersed orders of both the HEG and
MEG. Since we estimated there to be only ∼70 back-
ground counts in this extraction region during the course
of the observation (compared to more than 9000 source
counts), we did not background-subtract the light curves.
At high count rates, two or more photons can arrive
close enough together in time that they appear to be
a single event. This "pileup" phenomenon can cause a
spectrum to become distorted and the count rate to be
reduced. To confirm that the zeroth order spectrum was
On 2005 October 19, the Chandra X-ray Observatory
performed a 50.1 ks Director's Discretionary Time obser-
1 Chandra Interactive Analysis of Observations
(CIAO),
http://cxc.harvard.edu/ciao/.
The Hard-X-Ray Symbiotic RT Cru
3
not affected by pileup, we divided the number of counts
at the peak of the point spread function of the undis-
persed light by the number of 2.54-s frames in the obser-
vation to obtain an upper limit on the number of counts
per pixel per frame. The resulting 0.08 counts per pixel
per frame is well below the 1 count per pixel per frame
where pileup can become important (Harris et al. 2004).
Moreover, the average count rate in the zeroth order was
significantly less than one count per frame time, indicat-
ing that the light curve was not significantly distorted
by the saturation that can occur at higher count rates
(i.e., higher pileup fractions). The pileup fraction in the
higher-order spectrum was negligible.
3. ANALYSIS AND RESULTS
3.1. Spectral Analysis
To model the X-ray spectrum, we first consider simple,
single-component continuum models. We fitted these
models to the binned 0.3 -- 8.0 keV zeroth-order spec-
trum (above 8.0 keV, the noise rises and the quantum
efficiency drops sharply). Absorbed single-component
emission models such as a thermal plasma, powerlaw, or
blackbody (plus Gaussian lines) do not produce accept-
able fits. Even if we include complex absorption, such as
an absorber that only partially covers the source, a pow-
erlaw distribution of absorbers (as seen in some magnetic
CVs; e.g., Done & Magdziarz 1998), or a"warm" ionized
absorber, single-component emission models still do not
produce acceptable fits.
Including an additional broad-band emission compo-
nent improves the fitting results. The spectrum is for-
mally well fitted with a highly absorbed, optically thin
thermal plasma (Mekal model in Xspec), plus a moder-
ately absorbed non-thermal powerlaw component. Since
there is some degeneracy between the amount of absorp-
tion and the powerlaw index, we determine the powerlaw
index by fixing the plasma temperature, T , and fitting
the spectrum above 4 keV, where absorption is relatively
unimportant. The resulting photon index is Γ = 1.051.29
0.78,
where Γ is given by dFN /dE = K(E/E0)−Γ, dFN /dE
is the photon flux density, E0 is 1 keV, K is the nor-
malization constant, and the superscripts and subscripts
are 90% confidence upper and lower limits, respectively.
The photon index is not sensitive to the value to which
we fix T . We determined the remaining model param-
eters by fixing Γ to 1.1 and letting the other parame-
ters vary. The thermal plasma has kT = 8.612.4
6.1 keV
and an absorbing column nH = 9.312.5
7.2 × 1023 cm−2.
The powerlaw component has an absorbing column of
6.3 × 1022 cm−2 and a normalization K =
nH = 7.38.7
0.98 × 10−3 photons cm−2 s−1 keV−1. To obtain an
1.22.9
acceptable fit to the line emission, the model required
abundances of about 0.3 times solar (using the abun-
dances of Anders & Grevesse 1989). Because of the large
column density absorbing the thermal emission in this
model, the powerlaw component dominates below ap-
proximately 4 keV, and the thermal emission dominates
above 4 keV. We also obtained a formally acceptable
fit with a highly absorbed powerlaw and a moderately
absorbed thermal plasma.
In that case, the thermal
plasma dominated the low-energy portion of the spec-
trum, and the powerlaw dominated the high-energy part
of the spectrum.
Finally, we consider the isobaric cooling-flow model
that has worked well for both magnetic and non-magnetic
CVs. In this model (Mkcflow; Mushotzky & Szymkowiak
1988), gas is assumed to radiatively cool from a high
post-shock temperature under conditions of constant
pressure. Such an isobaric cooling flow has a differen-
tial emission-measure distribution that is a flat function
of temperature. The gas is also assumed to be optically
thin. This cooling flow model provides a good fit to the
data. We find that the initial post-shock temperature is
quite high. Although our 0.3 -- 8.0 keV spectrum does not
allow a high-confidence determination of this parameter,
the formal 90% confidence lower limit for the initial tem-
perature is kTmax = 55 keV. The minimum cooling-flow
temperature is consistent with the smallest value allowed
by the Xspec Mkcflow model (kTmin = 80.8 eV), indicat-
ing that the gas does indeed remain optically thin as it
cools. Allowing the differential emission measure to have
a powerlaw distribution (with the Cemkl model in Xspec)
did not improve the fit. As with the two-component con-
tinuum model, the cooling-flow model requires significant
absorption. The abundances may also be sub-solar (0.3
times solar, using the abundances of Anders & Grevesse
1989). The best absorption model consists of both a pho-
toelectric absorber that fully covers the source and an-
other that only partially covers it. Table 1 lists the best-
fit parameters for this model, which, as we discuss in §4,
we believe provides the best description of the Chandra
spectrum of RT Cru.
In the first-order spectrum, we detect the iron-line
complex spanning roughly 6.4 -- 7.0 keV. Fig. 2 shows
the region around the iron-line complex in the combined
MEG and HEG first-order (m = ±1) spectrum. Because
of the large absorption and the resulting low count rate at
low energies, we are not sensitive to lines such as OVIII
(∼ 19 A) and Ne X (∼ 12 A) that have been seen in
HETG observations of some other accreting WDs (e.g.,
Pandel et al. 2005; Mukai et al. 2003). For the Fe lines,
we use a simple powerlaw to establish a continuum level
and three Gaussian profiles to fit the Fe Kα, H-like Fe,
and He-like Fe lines, respectively. To avoid the possible
introduction of errors from misalignment of the HEG and
MEG spectra, we use only the combined HEG first-order
(m = ±1) spectrum for computation of the equivalent
widths (EWs). Table 2 lists the line-center energies and
EWs. Although we do not have enough counts to use the
recombination, intercombination, and forbidden compo-
nents of the H- or He-like Fe lines as density diagnostics,
the observed line strengths and EWs confirm that the
source is surrounded by a large amount of neutral mate-
rial and that the abundances might be slightly sub-solar.
The Fe Kα EW of 108 eV is consistent with that ex-
pected for a source inside a cloud of cold material with
the NH of ∼ 1023 cm−2 (Inoue 1985), as we found from
the continuum fitting.
3.2. Timing Analysis
Examining time series binned at 508.208 s and
4065.664 s (i.e., 200 and 1600 times the frame time, re-
spectively), we detected significant aperiodic, flickering-
type variations on time scales of minutes to hours in the
0.3 -- 4.0 keV emission. Figure 3 shows the 508-s binned
time series (light curves) in the energy ranges 0.3 -- 4.0 keV
and 4.0 -- 8.0 keV. The fractional amplitude of the stochas-
4
Luna & Sokoloski
tic variations appears to be largest in the 0.3 -- 4.0 keV
energy range. In the 508-s binned 0.3 -- 4.0 keV time se-
ries, the ratio of measured fractional rms variation, s, to
that expected from Poisson fluctuations alone, sexp, is
1.96 (s = 36.6% and sexp = 18.7%). We detect the 0.3 --
4.0 keV variability with even greater statistical signifi-
cance in the 4065-s binned time series; the ratio s/sexp in
this case is 3.35 (s = 22.1% and sexp = 6.6%). In the 4.0 --
8.0 keV energy range, the 508-s binned time series has
s/sexp = 1.36 (where s = 16.6% and sexp = 12.2%), and
the 4065-s binned time series has s/sexp = 1.42 (s = 6.1%
and sexp = 4.3%).
We do not detect any periodic flux modulations. We
are theoretically sensitive to an oscillation with fractional
amplitude:
graphs that follow). Comparing this luminosity with that
expected from accretion, (1/2)(GM M /R) & LX , the ra-
dius of the accreting compact object is:
R . 3.2 × 108 cm(cid:18) M
1.3M⊙(cid:19)
M
1.8 × 10−9M⊙ yr−1! ,
(3)
where R and M are the radius and mass of the accre-
tor, respectively, and M is the rate of accretion through
the boundary layer. The radius is that of a WD. The
Chandra X-ray spectrum therefore confirms that the
compact object is a WD.
To determine whether the Chandra-band X-ray emis-
sion is indeed from an accretion-disk boundary layer, we
consider the rapid variability. Rapid flickering typically
emanates from an accretion region close to a compact ob-
ject. Our detection of flickering therefore suggests that
the X-ray emission detected from RT Cru by Chandra is
powered by accretion. This accretion could proceed via a
wind-fed accretion disk, magnetic accretion columns, or
Bondi-Hoyle type direct-impact of the accreting material
onto the WD. While the two-component (thermal plasma
plus powerlaw) model provides a formally acceptable fit
to the data, it is difficult to construct an interpretation
of this model that is consistent with the rapid flicker-
ing from accretion onto a WD. The isobaric cooling-flow
spectral model, on the other hand, 1) provides a good fit
to the data, 2) has been successfully applied to both the
boundary layer emission from non-magnetic CVs (e.g.,
Mukai et al. 2003; Pandel et al. 2005) and the accretion
columns of magnetic CVs (e.g., Cropper et al. 1998), and
3) provides a natural context for the flickering from ac-
cretion.
Most CVs with X-ray emission as hard as that
which INTEGRAL and Swif t/BAT have detected from
RT Cru have magnetic fields strong enough to channel
the accretion flow into accretion columns (B ∼ 105 --
106 G, where B is the magnetic field strength at the
WD surface). Of the 8 CVs detected at energies greater
than ∼ 50 keV with Swif t/BAT, all but SS Cyg are
likely magnetic accretors (Barlow et al. 2006). In these
systems, the hard X-rays come from hot gas behind the
stand-off shock in the accretion column. Magnetic CVs
typically have X-ray oscillations with pulsation ampli-
tudes of tens of percent at the WD spin period, which
is usually less than an hour (Warner 1995). Since typi-
cal symbiotic-star accretion rates are higher than typical
CV accretion rates, if the WD in RT Cru was strongly
magnetic and in spin equilibrium, the spin period would
probably be either comparable to or faster than those
in CVs. Given our sensitivity to oscillations with peri-
ods less than an hour, we therefore should have detected
a spin modulation if RT Cru was magnetic. In fact, the
power spectrum has no statistically significant peaks. We
conclude that RT Cru is probably not a magnetic accre-
tor. We thus favor the picture in which the X-ray emis-
sion from RT Cru detected by Chandra is from a cooling
flow in an accretion-disk boundary layer.
4.2. Implications
The parameters of the cooling-flow fit to the boundary-
layer emission provide the accretion rate as well as infor-
mation about the WD. From the normalization param-
A ≈ 2 (Ctotα)−1/2(cid:18)ln(cid:20)
9, 400(cid:19)−1/2
= 0.08(cid:18) Ctot
1 − δ
1 − (1 − ǫ)1/nf req(cid:21)(cid:19)1/2
0.77(cid:17)−1/2
(cid:16) α
,
(1)
(2)
where Ctot is the total number of counts in the observa-
tion (ignoring the small number of background counts,
which have a negligible effect), δ and ǫ are small num-
bers related to the chance that a noise power in the power
spectrum will exceed the detection threshold (both taken
to be 0.05), nf req is the number of frequencies searched
(nf req = 1644), and α has an average value of 0.77 and
depends upon the location of the signal frequency in
the frequency bin (see, e.g., van der Klis 1989; Sokoloski
1999). We are therefore sensitive to oscillations with
fractional amplitudes of ∼ 8% in regions of the power
spectrum dominated by white noise, which in this case
consisted of frequencies greater than ∼ 1.4 mHz. In this
analysis, we binned the time series in 15-s bins (i.e., 6
times the frame time). We were therefore sensitive to os-
cillations with periods as short as 30 s, and most sensitive
to oscillations with periods between 30 s and 12 m.
Consistent with the presence of flickering in the light
curves, the power spectrum rises at frequencies below 1.4
mHz. At these low frequencies, the power spectrum has
a powerlaw index (i.e., slope on a log-log plot) of about
−1. This "1/f -noise" at low frequencies reduces our sen-
sitivity to oscillatory signals with periods greater than
approximately 12 m. Since we expect the minimum os-
cillation amplitude to which we are sensitive to increase
roughly as the square root of the rising average broad-
band power as we go to lower frequencies (i.e., longer
periods), the oscillation amplitude required for detection
increases gradually from &8% to &15% as we move from
periods of 12 m to 1 hr. Taking into account the underly-
ing broadband power, as well as the number of frequency
bins searched, we did not detect any statistically signifi-
cant oscillations in any portion of the power spectrum.
4. DISCUSSION
4.1. Interpretation of the Chandra Observations
To estimate the radius of the accreting compact object,
we take the unabsorbed 0.3 -- 8.0 keV X-ray luminosity,
LX, to be either approximately equal to, or a rough lower
limit to, the emission from an accretion-disk boundary
layer (we justify the assumption that the Chandra X-
ray emission emanates from a boundary layer in the para-
The Hard-X-Ray Symbiotic RT Cru
5
eter of the cooling-flow model, the accretion rate onto
M = 1.8 × 10−9 M⊙ yr−1 (d/2 kpc)2 (see Ta-
the WD is
ble 1). Since quasi-steady nuclear burning is frequently
present on the WD surface, symbiotic stars should have
accretion rates that are on average higher than those in
CVs. The accretion rate we have found for RT Cru is
consistent with this picture. It is also, however, just low
enough that we expect the boundary layer to remain op-
tically thin; Narayan & Popham (1993) find that for a
1 M⊙ WD, the boundary layer remains optically thin for
accretion rates below 3 × 10−9 M⊙ yr−1.
The parameters of the cooling-flow fit also indicate
that the WD radius is small, suggesting that the WD
could be quite massive. Taken at face value, the ra-
dius constraint would imply that the WD mass is at
least 1.3 M⊙. The high upper cooling-flow tempera-
ture, kTmax ≥ 55 keV (see Table 1), supports the conclu-
sion that the accretor is a massive WD. The relationship
between kTmax and WD mass is due to the fact that
that the Kepler velocity (v2
K = GM/R) is greater in the
deep potential well of a more massive WD. Since the
boundary-layer material is shock heated, and the initial
post-shock temperature (Tmax) is proportional to veloc-
ity squared, Tmax increases with WD mass. Alterna-
tively, if we equate the amount of energy available per
particle, (1/2)µmpv2
K, where µ is the mean molecular
weight and mp is the mass of a proton, with the energy re-
leased per particle in an isobaric cooling flow, (5/2)kTmax
(Pandel et al. 2005), we see that kTmax ∝ v2
K ∝ GM/R.
Although the determination of kTmax from X-ray emis-
sion below 8 keV is highly uncertain, we can still ask
what such a high kTmax would imply if it is confirmed
by an instrument with greater high-energy sensitivity
(such as SUZAKU). For their sample of 9 non-magnetic
CVs, Pandel et al. (2005) found the upper cooling-flow
temperature was roughly consistent with the expected
kTmax = (3/5) kTvir, where Tvir is the virial temperature
(defined by (3/2)kTvir = (1/2)µmpv2
K). Using the WD
mass-radius relationship of Hansen & Kawaler (1994), a
maximum cooling-flow temperature of kTmax > 55 keV
implies M & 1.3 M⊙.
The high absorbing columns for both the partial-
covering and fully covering absorber, as well as the cover-
ing factor of >0.7 for the partial-covering absorber (see
Table 1), indicate that the X-ray source is highly ob-
scured at this epoch. Since the X-ray emission region
is small, the absorber that only partially covers the X-
ray source must also be small. A possible source of this
partially covering absorber is an accretion structure such
as an accretion disk seen almost edge-on (as in OY Car;
Pandel et al. 2005). We assume that the fully covering
component of the absorption comprises both interstel-
lar absorption (1.1×1022cm−2 from NASA/IPAC IRSA)
and intrinsic absorption. The column density of this ab-
sorber is probably high because the WD orbits within the
strong, dense stellar wind from the red giant. RT Cru
has an orbital period of ∼ 450 d (J. Miko lajewska, pri-
vate communication), and therefore a binary separation
on the order of an AU. This separation puts the WD well
within the dense region of the red-giant wind. Month-
time-scale variations in the absorption (not correlated
with the orbital period) have been detected by Swif t
(Kennea et al. 2007), suggesting that either the red-giant
wind is clumpy, that the mass loss rate in the wind of the
red giant is variable, or the accretion structure or struc-
tures partially blocking the WD boundary layer must be
unstable.
4.3. The Nature of RT Cru
As shown by the INTEGRAL detection in 2003-2004
(Chernyakova et al. 2005), RT Cru can at times produce
X-ray emission out to greater than 60 keV. From the
broad-band flux densities reported by Chernyakova et al.
(2005), the hard X-ray spectrum in 2003 - 2004 appears
consistent with a powerlaw with photon index Γ = 2.7
(energy index α = 1.7), and the 16-100 keV luminos-
ity at that time was approximately 10 L⊙ (d/2kpc)2. By
2005, however, the 16-100 keV luminosity had dropped,
and the hard X-ray spectrum was closer to thermal
(Kennea et al. 2007). To better appreciate the proper-
ties of the unusual accreting WD in RT Cru, we briefly
explore the possible sources of the powerlaw hard X-ray
emission observed by INTEGRAL in 2003-2004. In par-
ticular, we consider direct synchrotron emission, inverse-
Compton (IC) scattering from a thermal distribution of
electrons, and IC scattering from a non-thermal distri-
bution of electrons.
IC scattering from a non-thermal
distribution of electrons turns out to be the most likely
option.
One way to generate a powerlaw energy spectrum
is with synchrotron emission from a powerlaw distri-
bution of relativistic electrons moving in a magnetic
field. Most of the synchrotron emission from an elec-
tron with Lorentz factor γ is emitted at a frequency
νsync = (0.3/2π)(3 sin α/2)γ2
sync(qB/mec), where α is
the pitch angle, q is the electron charge, B is the mag-
netic field strength, me is the electron mass, and c is the
speed of light (Rybicki & Lightman 1979). Since we did
not detect X-ray pulsations from RT Cru, the magnetic
field at the surface of the WD is probably not strong
enough to disrupt the accretion flow, and therefore less
than ∼ 104 G. Thus, even near the surface of the WD,
Lorentz factors γsync of a few times 104 to 105 would be
needed to produce significant direct synchrotron emission
at 60 keV. We can estimate the maximum Lorentz fac-
tor to which electrons will be accelerated by setting the
diffusive shock acceleration time scale equal to the syn-
chrotron cooling time scale. Following the approach of
Markoff et al. (2001), we equate the synchrotron loss rate
to a conservative acceleration rate (see Jokipii 1987) to
find a maximum Lorentz factor to which electrons can be
accelerated of γmax ≈ 2700(B/104 G)−1/2(ξ/100)−1/2,
where ξ is the ratio of the diffusive scattering mean free
path to the gyroradius. Thus γmax is one to two orders
of magnitude below γsync. Moreover, whereas significant
power from a distribution of synchrotron-emitting rela-
tivistic electrons would also be expected at energies be-
low 15 keV, the 16-100 keV X-ray luminosity of ∼ 10 L⊙
during 2003 and 2004 is already close to the total en-
ergy budget available from accretion. Finally, to gen-
erate the X-ray spectral index observed by INTEGRAL
from direct synchrotron emission, one would need a dis-
tribution of electrons that is much steeper than expected
from standard theories of particle acceleration in shocks
(Ellison et al. 1990).
It is therefore unlikely that the
hard X-ray emission from RT Cru was due to direct syn-
6
chrotron emission.
As an aside, we note that the production of significant
synchrotron emission at radio wavelengths would not re-
quire such high Lorentz factors. The ratio of synchrotron
power to IC power from an electron is equal to the ratio of
the energy density in the magnetic field to energy density
in photons (Rybicki & Lightman 1979). Examining this
ratio as a function of distance from the WD, we expect a
similar amount of power from direct synchrotron and IC
scattering near the surface of the WD, where the B field
could be strong (B ∼ 104 G). We therefore suggest that
the next time RT Cru is in a powerlaw hard X-ray state
like the one detected by INTEGRAL in 2003-2004, radio
observations be performed to look for radio synchrotron
emission.
Hard X-rays can be produced with much lower Lorentz
factors through IC scattering. For photons scattering off
of a thermal distribution of electrons, Reynolds & Nowak
(2003) give a relation between the observed photon in-
dex Γ and the Compton y parameter, which is related
to the factor by which the average photon energy in-
creases. A steep photon index of Γ > 2 indicates a y
parameter less than 1,
in which case there is no sig-
nificant up-scattering. Therefore, since RT Cru had a
steep photon index of Γ ≈ 2.7, scattering off of a ther-
mal distribution of electrons could not have been re-
sponsible for the powerlaw hard X-ray emission. Non-
thermal, relativistic electrons must have been involved
in producing the powerlaw spectrum observed by INTE-
GRAL. RT Cru thus contains a white dwarf that can
accelerate electrons to relativistic speeds. Three other
systems that contain WDs that also generate relativis-
tic electrons are CH Cyg, RS Oph, and possibly R Aqr,
all of which have jets (Rupen et al. 2007; O'Brien et al.
2006; Crocker et al. 2001; Nichols et al. 2007).
Assuming now that the powerlaw hard X-ray spectrum
was due to IC scattering from a powerlaw distribution of
relativistic electrons, we can estimate the location of the
scattering electrons. Since the strength of a dipole B field
falls like 1/r3, whereas the energy density in the photon
field only falls like 1/r2, IC scattering will dominate as
one moves farther from the WD. If we ask how far away
from the radiation source the IC region should be to give
us an IC cooling time on the order of a year (the approx-
imate duration of the powerlaw hard X-ray state), we
find that it should be a few tenths of an AU away. Given
the orbital period for this systems, that could put the
IC emission region either between the WD and the red
giant, as in a colliding-winds region, above the WD disk
in a corona, or at the base of a jet. A model consisting
of IC scattering off of relativistic electrons in a corona
or at the base of a jet reproduces the observed hard X-
ray emission well in X-ray binaries (Markoff et al. 2005).
The steep powerlaw index, which is similar to that seen in
the steep powerlaw state in microquasars, could be due
Luna & Sokoloski
to a low scattering optical depth (Rybicki & Lightman
1979).
5. CONCLUSIONS
We have observed the first symbiotic star with X-ray
emission out to greater than 60 keV with the HETG
on Chandra. The stochastic variability and cooling-flow
type spectrum suggest that RT Cru is powered by ac-
cretion onto a WD through a disk with an optically thin
boundary layer. The accretion rate is near the top of the
range for which the boundary layer can remain optically
thin. The high initial temperature of the cooling flow,
and the high luminosity given the accretion rate from the
spectral fit suggest that the WD in RT Cru could be quite
massive. More generally speaking, it would be difficult to
get such hard X-ray emission from accretion onto a low-
mass WD with a shallow potential well and low Kepler
velocity. Given the nature of the powerlaw hard X-ray
emission previously observed by INTEGRAL, it therefore
appears that the accreting, non-magnetic WD in RT Cru
is able to generate a non-thermal, powerlaw distribution
of electrons and very hard X-ray emission through IC
scattering. Three other systems in which WDs can accel-
erate electrons to relativistic speeds (RS Oph, CH Cyg,
and possibly R Aqr; Rupen et al. 2007; O'Brien et al.
2006; Crocker et al. 2001; Nichols et al. 2007) all have
jets. Radio observations of RT Cru during the next pow-
erlaw hard X-ray state like that observed by INTEGRAL
in 2003 and 2004 could play an important role in diag-
nosing this system, and determining the extent to which
some symbiotic stars might constitute the nanoquasar
analog to microquasars (Zamanov & Marziani 2002).
We thank the referee, Marina Orio,
for comments
and suggestions which improved the final quality of
this article. We also thank K. Mukai, F. Paerels, T.
Maccarone, and A. J. Bird for useful discussions, R.
Lopes de Oliveira and D. Huenemoerder for help with
the data analysis, and S. Markoff for comments on the
manuscript. G.J.M. Luna acknowledges support from
CNPq (process 0141805/2003-0) and FAPESP (process
02/08816-5). J.L.S. is supported by an NSF Astron-
omy and Astrophysics Postdoctoral Fellowship under
award AST-0302055. Support for this work was pro-
vided by the National Aeronautics and Space Admin-
istration through Chandra Award Numbers DD5-6034X
and NNX06AI16G issued by the Chandra X-ray Obser-
vatory Center, which is operated by the Smithsonian As-
trophysical Observatory for and on behalf of the National
Aeronautics Space Administration under contract NAS8-
03060. We acknowledge with thanks the variable star
observations from the AAVSO International Database
contributed by observers worldwide and used in this re-
search.
REFERENCES
Arnaud K. A., 1996, Astronomical Data Analysis Software and
Systems V, eds. Jacoby G. and Barnes J., p17, ASP Conf.
Series volume 101.
Anders E. & Grevesse N., 1989, Geochimica et Cosmochimica
Acta 53, 197
Barlow, E. J., Knigge, C., Bird, A. J., J Dean, A., Clark, D. J.,
Hill, A. B., Molina, M., & Sguera, V. 2006, MNRAS, 372, 224
Bird, A. J., et al. 2007, ApJS, 170, 175
Belczy´nski, K., Mikolajewska, J., Munari, U., Ivison, R. J.,
Friedjung, M., 2000, A&AS, 146,407
Canizares, C. et al. 2005, 117, 1144
Chakrabarty, D., & Roche, P. 1997, ApJ, 489, 254
Chernyakova M., Courvoisier T. J., Rodriguez J., Lutovinov A.,
2005, The Astronomer's Telegram, 519,1
The Hard-X-Ray Symbiotic RT Cru
7
Cieslinski D., Elizalde F., Steiner J. E., 1994, A&A, 106, 243
Crocker M. M., Davis R. J., Eyres S. P. S., Bode M. F., Taylor A.
R., Skopal A., Kenny H. T., 2001, MNRAS, 326,781
Crocker M. M., Davis R. J., Spencer R. E., Eyres S. P. S., Bode
M. F., Skopal A., 2002, MNRAS, 335, 1100
Cropper M., Ramsay G., Wu K., 1998, MNRAS, 292, 222
Davidsen, A., Malina, R., & Bowyer, A. 1977, ApJ, 211, 866
Done, C., & Magdziarz, P. 1998, MNRAS, 298, 737
Ellison, D. C., Reynolds, S. P., & Jones, F. C. 1990, ApJ, 360, 702
Galloway, D. K., Sokoloski, J. L., & Kenyon, S. 2002, ApJ, 580,
1065
Houck, J. C.,High Resolution X-ray Spectroscopy with
XMM-Newton and Chandra, Proceedings of the international
workshop held at the Mullard Space Science Laboratory of
University College London, Holmbury St Mary, Dorking,
Surrey, UK, October 24 - 25, 2002, Ed. Branduardi-Raymont,
G., published electronically and stored on CD., meeting
abstract
Hachisu, I., & Kato, M. 2001, ApJ, 558, 323
Hansen, C. J., & Kawaler, S. D. 1994, Stellar Interiors. Physical
Principles, Structure, and Evolution, XIII, Springer-Verlag
Berlin Heidelberg New York
Harris D. E., Mossman A. E., Walker R. C., 2004, ApJ, 615, 161
Inoue H., 1985, Sp. Sci. Rev., 40, 317
Jablonski, F. J., Pereira, M. G., Braga, J., & Gneiding, C. D.
1997, ApJ, 482, L171
Jokipii, J. R. 1987, ApJ, 313, 842
Jordan S., Muerset U. & Warner K., 1994, A& A, 283, 475
Karovska, M., Carilli, C. L., Mattei, J. A., 1998, JAVSO, 26,97
Karovska M., Schlegel E., Hack W., Raymond J., Wood B. E.,
2005, ApJ, 623,137
Kenyon, S. J., The symbiotic stars, 1986, Cambridge University
Press
Kennea, J., Mukai, K., et al. 2007 in prep.
Luna, G. J. M., Sokoloski J. L., Mukai K., Costa, R. D. D., in
prep.
Livio M., 1997, in: Accretion Phenomena and Related Outflows,
IAU Colloquium 163, ASP Conf. Series, Vol. 121, D. T.
Wickramasinghe, L. Ferrario & G. V. Bicknell, eds., p. 845
Makishima, K. 1986, in The Physics of accretion onto Compact
Objects, eds. K. O. Mason, M. G. Watson & N. E. White,
Springer Verlag Berlin, p249
Markoff, S., Falcke, H., & Fender, R. 2001, A&A, 372, L25
Markoff, S., Nowak, M. A., & Wilms, J. 2005, ApJ, 635, 1203
Masetti, N., et al. 2002, A&A, 382, 104
Masetti, N., Bassani L., Bird A. J., Bazzano A., 2005, The
Astronomer's Telegram, 528
Masetti N, Bassani L., Dean A. J., Ubertini P., Walter R., 2006,
The Astronomer's Telegram, 715
Masetti, N., et al. 2007, A&A, 464, 277
Masetti, N., et al. 2007, A&A, 470, 331
Mushotzky R. F. & Szymkowiak A. E., 1988, Cooling Flows in
Clusters and Galaxies ed. A. C. Fabian, p. 53
Mukai K., Kinkhabwala A., Peterson J. R., Kahn S. M., Paerels
F., 2003, ApJ, 586, 77
Mukai K., Ishida M., Kilbourne C., Mori H., Terada Y., Chan K.,
Soong Y., 2007, PASJ, 59, 177
Muerset U., Wolff B., Jordan S., 1997, A& A, 319, 201
Narayan, R., & Popham, R. 1993, Nature, 362, 820
Nichols, J. S., DePasquale, J., Kellogg, E., Anderson, C. S.,
Sokoloski, J., & Pedelty, J. 2007, ApJ, 660, 651
O'Brien, T. J., et al. 2006, Nature, 442, 279
Orio M., Zezas A., Munari U., Siviero A., Tepedelenlioglu E.,
2007, ApJ, 661, 1105
Pandel D., C´ordova F. A., Mason K. O., Priedhorsky W. C.,
2005, ApJ, 626, 396
Reynolds, C. S., & Nowak, M. A. 2003, Phys. Rep., 377, 389
Rupen, M. J., Mioduszewski, A. J., & Sokoloski, J. L. 2007,
submitted to ApJ
Rybicki, G. B., & Lightman, A. P. 1979, New York,
Wiley-Interscience, 1979, p. 179
Sokoloski, J. L. 1999, Ph.D. Thesis, U.C. Berkeley
Sokoloski, J. L. 2003, Astronomical Society of the Pacific
Conference Series, 303, 202
Sokoloski J. L., Bildsten L., Ho W. C. G. 2001, MNRAS, 326, 553
Sokoloski J., Kenyon S. J., 2003, ApJ, 584, 1021
Sokoloski J. L., Kenyon S. J., Espey B. R., Keyes C. D.,
McCandliss S. R., Kong A. K. H., Aufdenberg J. P., Filippenko
A. V., Li W., Brocksopp C., et al., 2006, ApJ, 636,1002
Taylor A. R., Seaquist E. R., Mattei J. A., 1986, Nature, 319, 38
Tueller J., Gehrels N., Mushotzki R. F., Markwardt C. B.,
Kennea J. A., Burrows D. N., Mukai K., Sokoloski J., 2005,
The Astronomer's Telegram, 591
Tueller J., Barthelmy S., Burroes D., Falcone A., Gehrels N.,
Grupe D., Kennea J., Markwardt C. B., Mushotzky R. F.,
Skinner G. K., 2005, The Astronomer's Telegram, 669
van Belle, G. T., et al. 1999, AJ, 117, 521
van der Klis, M. 1989,in Timing Neutron Stars (Kluwer Academic
Publishers), Ogelman, H., & van den Heuval, E. P. J. eds., p27
Warner, B. 1995, Cataclysmic Variable Stars (Cambridge
University Press: Cambridge)
Zamanov, R., & Marziani, P. 2002, ApJ, 571, L77
8
Luna & Sokoloski
Cooling-flow model parameters.
TABLE 1
Parameter
Value (Min, Max)a
M b (10−9 M⊙ yr−1)
kTmax (keV)
nH : Full covering (1022 cm−2)
nH : Partial covering (1022 cm−2)
Covering Fraction
Abundance (w.r.t. Solarc)
FX
LX
d (10−12 erg cm−2 s−1)
d (1034 erg s−1)
1.8 (1.6, 2.0)
80 (56,
· · · )
8.2 (7.7, 8.8)
65 (52, 78)
0.74 (0.69, 0.79)
0.30 (0.02, 0.45)
9.1 (7.5, 10.2)
3.1 (2.4, 3.9)
Note. -- The model consists of optically thin thermal
emission from an isobaric cooling flow with absorbers that
both fully cover and partially cover the source, plus a
Gaussian line.
a 90% confidence upper and lower limits.
b Accretion rate onto the compact object.
c Using solar abundance of Anders & Grevesse (1989).
d FX is the absorbed 0.3 -- 8.0 keV flux, and LX is the
unabsorbed 0.3 -- 8.0 keV luminosity (d = 2 kpc).
The Hard-X-Ray Symbiotic RT Cru
9
TABLE 2
Iron lines.
Fe XXV
Fe XXVI
Fe Kα
Line center (keV)a:
EWb (eV):
6.9466.959
6.933
6.6936.706
6.677
6.3796.399
6.358
72
60
108
a Super- and subscripts represent 90% confidence upper
and lower limits, respectively.
b Gaussian-fit equivalent widths. EW uncertainties are on
the order of 10 -- 15%.
Luna & Sokoloski
10
1
−
V
e
k
1
−
s
s
t
n
u
o
C
2
χ
0.02
0.01
5×10−3
2×10−3
2
0
−2
2
Energy (keV)
5
Fig. 1. -- Undispersed (zeroth order) spectrum. The top panel shows the spectrum with the absorbed, isobaric cooling-flow model over-
plotted. The bottom panel shows residuals with respect to this model (in units of χ2, where χ2 is shorthand for the difference between the
data and the model, squared, divided by the variance, with the sign of the difference between the data and the model).
Fig. 2. -- The iron-line complex from the combined HEG and MEG first-order (m = ±1) spectrum. The best fit model of a powerlaw
plus three Gaussian emission lines is over-plotted. The bottom panel shows the residuals, in the same units as Fig. 1.
The Hard-X-Ray Symbiotic RT Cru
11
Fig. 3. -- Chandra light curves for RT Cru, with a bin size of 508.28 s. The light curves include the undispersed light as well as the
counts from the HEG and MEG m = ±1 orders. The top and bottom panels show the flux as a function of time in the energy ranges
0.3 -- 4.0 keV and 4.0 -- 8.0 keV, respectively. The 0.3 -- 4.0 keV emission is clearly variable on time scales of minutes to hours.
|
astro-ph/9611107 | 4 | 9611 | 1997-10-21T23:14:37 | A Universal Density Profile from Hierarchical Clustering | [
"astro-ph"
] | We use high-resolution N-body simulations to study the equilibrium density profiles of dark matter halos in hierarchically clustering universes. We find that all such profiles have the same shape, independent of halo mass, of initial density fluctuation spectrum, and of the values of the cosmological parameters. Spherically averaged equilibrium profiles are well fit over two decades in radius by a simple formula originally proposed to describe the structure of galaxy clusters in a cold dark matter universe. In any particular cosmology the two scale parameters of the fit, the halo mass and its characteristic density, are strongly correlated. Low-mass halos are significantly denser than more massive systems, a correlation which reflects the higher collapse redshift of small halos. The characteristic density of an equilibrium halo is proportional to the density of the universe at the time it was assembled. A suitable definition of this assembly time allows the same proportionality constant to be used for all the cosmologies that we have tested. We compare our results to previous work on halo density profiles and show that there is good agreement. We also provide a step-by-step analytic procedure, based on the Press-Schechter formalism, which allows accurate equilibrium profiles to be calculated as a function of mass in any hierarchical model. | astro-ph | astro-ph |
A Universal Density Profile from Hierarchical Clustering
Julio F. Navarro 1
Steward Observatory, University of Arizona, Tucson, AZ, 85721, USA.
Carlos S. Frenk 2
Physics Department, University of Durham, Durham DH1 3LE, England.
Simon D.M. White 3
Max Planck Institut fur Astrophysik, Karl-Schwarzschild Strasse 1, D-85740, Garching, Germany.
ABSTRACT
We use high-resolution N-body simulations to study the equilibrium density profiles
of dark matter halos in hierarchically clustering universes. We find that all such
profiles have the same shape, independent of halo mass, of initial density fluctuation
spectrum, and of the values of the cosmological parameters. Spherically averaged
equilibrium profiles are well fit over two decades in radius by a simple formula originally
proposed to describe the structure of galaxy clusters in a cold dark matter universe.
In any particular cosmology the two scale parameters of the fit, the halo mass and its
characteristic density, are strongly correlated. Low-mass halos are significantly denser
than more massive systems, a correlation which reflects the higher collapse redshift
of small halos. The characteristic density of an equilibrium halo is proportional to
the density of the universe at the time it was assembled. A suitable definition of
this assembly time allows the same proportionality constant to be used for all the
cosmologies that we have tested. We compare our results to previous work on halo
density profiles and show that there is good agreement. We also provide a step-by-step
analytic procedure, based on the Press-Schechter formalism, which allows accurate
equilibrium profiles to be calculated as a function of mass in any hierarchical model.
Subject headings: cosmology: theory – dark matter – galaxies: halos – methods:
numerical
1Bart J. Bok Fellow. E-mail: [email protected]
2E-mail: [email protected]
3E-mail: [email protected]
– 2 –
1.
Introduction
It is twenty-five years since the discovery that galaxies are surrounded by extended massive
halos of dark matter. A variety of observational probes – disk rotation curves, stellar kinematics,
gas rings, motions of globular clusters, planetary nebulae and satellite galaxies, hot gaseous
atmospheres, gravitational lensing effects – are now making it possible to map halo mass
distributions in some detail. These distributions are intimately linked to the nature of the dark
matter, to the way halos formed, and to the cosmological context of halo formation.
Insight into these links came first from analytic studies. Building on the early work of Gunn &
Gott (1972), similarity solutions were obtained by Fillmore & Goldreich (1984) and Bertschinger
(1985) for the self-similar collapse of spherical perturbations in an Einstein-de Sitter universe.
Such solutions necessarily resemble power laws in the virialized regions. Hoffman & Shaham (1985)
and Hoffman (1988) extended this analysis by considering open universes, and by modeling as
scale-free spherical perturbations the ob jects which form by hierarchical clustering from power-law
initial density perturbation spectra (P (k) ∝ kn ). They argued that isothermal structure (ρ ∝ r−2 )
should be expected in an Einstein-de Sitter universe if n ≤ −2, and that steeper profiles should be
expected for larger n and in open universes.
Despite the schematic nature of these arguments, their general predictions were verified as
numerical data became available from N -body simulations of hierarchical cosmologies. Power-law
fits to halo density profiles in a variety of simulations all showed a clear steepening as n increases
or the density of the universe decreases (Frenk et al. 1985, 1988; Quinn et al. 1986; Efstathiou et
al. 1988; Zurek, Quinn & Salmon 1988; Warren et al. 1992; Crone, Evrard & Richstone 1994).
An apparent exception was the work of West et al. (1987), who found that galaxy cluster density
profiles show no clear dependence on n.
Significant departures from power-law behaviour were first reported by Frenk et al. (1988),
who noted that halo profiles in cold dark matter (CDM) simulations steepen progressively with
increasing radius. Efstathiou et al. (1988) found similar departures – at odds with the analytic
predictions – in their simulations of scale-free hierarchical clustering. They also noted that these
departures were most obvious in their best resolved halos. Similar effects were noted by Dubinski
& Carlberg (1991) in a high resolution simulation of a galaxy-sized CDM halo. These authors
found their halo to be well described by a density profile with a gently changing logarithmic slope,
specifically that proposed by Hernquist (1990).
In earlier papers of this series we used high-resolution simulations to study the formation
of CDM halos with masses spanning about four orders of magnitude, ranging from dwarf galaxy
halos to those of rich galaxy clusters (Navarro, Frenk & White 1995, 1996). This work showed
that the equilibrium density profiles of CDM halos of all masses can be accurately fit over two
decades in radius by the simple formula,
ρ(r)
ρcrit
=
δc
(r/rs )(1 + r/rs )2 ,
(1)
– 3 –
where rs is a scale radius, δc is a characteristic (dimensionless) density, and ρcrit = 3H 2/8πG is
the critical density for closure. This profile differs from the Hernquist (1990) model only in its
asymptotic behaviour at r ≫ rs (it tends to r−3 instead of r−4 ). Power-law fits over a restricted
radial range have slopes which depend on the range fitted, steepening from −1 near the center to
−3 at large r/rs .
This similarity between CDM halos of widely differing mass is surprising in view of the strong
dependence on power spectrum shape reported in earlier studies. The effective slope of the CDM
power spectrum varies from neff ≈ −2 on galactic scales to neff ≈ −1 on cluster scales, so one
might have expected shallower profiles in galaxy halos than in clusters. In fact, the opposite is
true; low-mass halos are denser, i.e. have higher values of δc , than high-mass halos. This property
reflects the higher collapse redshift of the smaller systems. Power-law fits actually yield steeper
slopes for less massive halos both when carried out at a fixed radius and when carried out at a
fixed fraction of the virial radius (see Figures 3 and 4 of Navarro, Frenk & White 1996).
We argue below that the apparent relationship between profile shape and initial power
spectrum seen in earlier work results from systematic differences in the characteristic density of
the halos chosen when comparing different models. This interpretation is reinforced by the recent
work of Cole & Lacey (1996) and Tormen, Bouchet & White (1996), who find that the profiles
of massive halos formed from power-law spectra are well described by eq. 1. They also confirm
the strong correlation between δc and halo mass seen in our earlier work, and they find that, at
a given mass, halos are denser when n is larger. Thus, the spectral index n seems to control the
exact relationship between characteristic density and halo mass, rather than the effective slope of
the density profile.
It is clear from this discussion that a comprehensive study of this problem needs to consider
the role of at least three factors: the halo mass, the power spectrum of initial density fluctuations,
and the values of the cosmological parameters. In this paper we present the results of a large
set of N-body simulations specifically designed to address these issues. We consider a variety of
hierarchical clustering models, including CDM and power-law initial fluctuation spectra, as well
as different values of the cosmological parameters Ω0 and Λ. In each cosmology we study halos
spanning a large range in mass, carefully choosing numerical parameters so that all systems are
simulated with comparable numerical resolution.
The plan of the paper is as follows. Our numerical experiments are described in §2. In §3 we
present our results and in §4 we discuss them in the context of earlier work. In §5 we summarize
our main conclusions. An appendix lays out the formulae necessary to calculate analytically the
density profile of an equilibrium halo of any mass in any hierarchical cosmology.
2. The Numerical Experiments
– 4 –
2.1. The Cosmological Models
We analyze the structure of dark matter halos in 8 different cosmologies. Five are Einstein-de
Sitter (Ω0 = 1) models with various power spectra: the standard biased CDM spectrum (SCDM
model: Ω0 = 1, h = 0.5, σ8 = 0.63) and four power-law or “scale-free” spectra with indices n = 0,
−0.5, −1, and −1.5. Two further models also have power-law spectra (n = 0 and −1) but in an
open universe (Ω0 = 0.1). The last model we consider is a low-density CDM model with a flat
geometry (CDMΛ: Ω0 = 0.25, Λ = 0.75, h = 0.75, σ8 = 1.3). (Here and throughout this paper we
express the cosmological constant Λ in units of 3H 2
0 , so that a low-density universe with a flat
geometry has Ω0 + Λ = 1. We also adopt the standard convention of writing the present Hubble
constant as H0 = 100 h km s−1 Mpc−1 .)
The normalization of the CDM models is specified by σ8 , the rms mass fluctuation in spheres
of radius 8 h−1 Mpc. Because of self-similarity, the normalization of the scale-free models is
arbitrary. The evolutionary state of these models may be fully specified by a single parameter,
the current value of the “nonlinear mass”, M⋆ (z ). This mass scale is defined by requiring that the
variance of the linear overdensity field at redshift z = 0, smoothed with a top-hat filter enclosing
a mass M = M⋆ , should equal the square of the critical density threshold for spherical collapse by
0 (M⋆ (z )) = δ2
redshift z : ∆2
crit(z , Ω0 , Λ). (See the Appendix for details on the computation of δcrit).
This definition provides a “natural” way to scale the scale-free simulations to physical units and
to compare different cosmological models. In scale-free models the mass scale defined by M⋆ (z ) is
the only physical scale, and therefore the structure of halos can depend on mass only through the
ratio M /M⋆ .
2.2. The Simulations
Large cosmological N–body simulations are required to simulate the evolution of dark matter
halos in their full cosmological context. However, such simulations are not generally well suited
to explore a large range of halo masses. This is because systems of differing mass formed in
a single simulation are resolved to differing degrees. More massive systems are better resolved
because they contain more particles and because the gravitational softening is a smaller fraction
of the virial radius. These systematic differences can introduce insidious numerical artifacts in
the mass trends that we wish to investigate. We circumvent this problem by using the procedure
outlined by Navarro, Frenk & White (1996). Halos are first identified in cosmological N-body
simulations of large periodic boxes and then resimulated individually at higher resolution. During
the resimulation the remainder of the original simulation is treated only to the accuracy needed
to model tidal effects on the halo of interest. The advantage of this procedure is that numerical
parameters can be tuned so that all halos are simulated with comparable resolution. Its main
disadvantage is that only one halo is modeled per simulation, so that many simulations are needed
to compile a representative halo sample.
– 5 –
2.2.1. The cosmological simulations
The cosmological simulations were carried out using the P3M code of Efstathiou et al.
(1985). The desired initial power spectrum was generated by using the Zel’dovich approximation
to displace particles from a uniform initial load. The uniform load we used was either a “glass”
configuration (White 1996) or a cubic grid. For simulations with power-law power spectra, the
amplitude of the initial displacements was chosen by setting the power of the perturbed density
field to the white-noise level at the Nyquist frequency of the particle grid. These simulations
followed 106 particles on a 1283 mesh and were stopped when the nonlinear mass, M⋆ , corresponded
to 1, 000-2, 000 particles. We identify this time with the present (z = 0). Since clustering evolves
faster for more negative values of n, an expansion factor of 9.5 was sufficient for n = −1.5 whereas
an expansion factor of 90 was required for n = 0 (Ω0 = 1). Open models require even longer
integrations, an expansion factor exceeding 150 for n = 0 and Ω0 = 0.1. These simulations used
the time-stepping and numerical scheme described in Efstathiou et al (1988). (Note, however, that
the definition of M⋆ in that paper differs from the one we use here.)
We ran two P3M simulations for each of our CDM models. The SCDM runs followed
643 particles in periodic boxes of 180 h−1 and 15 h−1 Mpc and were stopped when σ8 = 0.63
(M⋆ ≈ 1.6 × 1013h−1 M⊙ ), the time which we identify with the present. The CDMΛ runs followed
106 particles in boxes of 140 h−1 and 46.67 h−1 Mpc, until σ8 = 1.3 (M⋆ ≈ 4.1 × 1013 h−1 M⊙ ).
2.2.2. The individual halo simulations
Halos to be resimulated at higher resolution were selected randomly at z = 0 from a list of
clumps identified using a friends-of-friends group finder with linking length set to 10% of the mean
interparticle separation. We chose masses in the range 0.1–10M⋆ for the power-law models and
0.01–100M⋆ for the CDM models. (The mass range is larger for the CDM models because in this
case halos were chosen from two parent cosmological simulations with different box sizes.) Because
we are interested the structure of equilibrium halos we were careful not to choose for analysis any
halo which is far from virial equilibrium. In practice, we analyze each resimulated halo at the
time between redshifts 0.05 and 0 when it is closest to dynamical equilibrium, defined as the time
when the ratio of kinetic to potential energy is closest to 0.5 for material within the virial radius.
[Throughout this paper, we measure halo masses, M200 , within a virial radius, r200 , defined as the
radius of a sphere of mean interior density 200 ρcrit . Halo circular speeds, V200 = (GM200 /r200 )1/2 ,
are also measured at this radius unless otherwise specified. Numerical experiments show that for
Ω = 1 this radius approximately separates the virialized and infall regions (Cole & Lacey 1996).
For convenience we continue to use these definitions when Ω0 6= 1.]
Once a halo is chosen for resimulation, the particles within its virial radius are traced back
to the initial conditions, where a small box containing all of them is drawn. This box is filled
with ∼ 323 particles on a cubic grid which are then perturbed using the waves of the original
– 6 –
P3M simulation, together with extra high-frequency waves added to fill out the power spectrum
between the Nyquist frequencies of the old and new particle grids. The regions beyond the
“high-resolution” box are coarsely sampled with a few thousand particles of radially increasing
mass in order to account for the large-scale tidal fields present in the original simulation.
This procedure ensures the formation of a clump similar in all respects to the one selected in
the P3M run, except for the improved numerical resolution. The size of the high-resolution box
scales naturally with the total mass of each system and, as a result, all resimulated halos have
about the same number of particles within the virial radius at z = 0, typically between 5, 000
and 10, 000. The extensive tests presented in Navarro, Frenk & White (1996) indicate that this
number of particles is adequate to resolve the structure of a halo over approximately two decades
in radius. We therefore choose the gravitational softening, hg , to be 1% of the virial radius in all
cases. (This is the scale-length of a spline softening; see Navarro & White 1993 for a definition.)
The tree-code carries out simulations in physical, rather than comoving, coordinates, and uses
individual timesteps for each particle. The minimum timestep depends on the maximum density
resolved in each case, but was typically 10−5H −1
0 .
As discussed in Navarro, Frenk & White (1996), numerically convergent results require
that the initial redshift of each run, zinit , should be high enough that all resolved scales in the
initial box are still in the linear regime. In order to satisfy this condition, we chose zinit so that
the median initial displacement of particles in the high-resolution box was always less than the
mean interparticle separation. Problems with this procedure may arise if zinit is so high that the
gravitational softening (which is kept fixed in physical coordinates) becomes significantly larger
that the mean initial interparticle separation. We found this to be a problem only for the smallest
masses, M ∼<M⋆ , in the n = 0, Ω0 = 0.1 model. In this case the initial redshift prescribed by the
median displacement condition is zinit > 700 and the gravitational softening is then a significant
fraction of the initial box. This can affect the collapse of the earliest progenitors of these systems
and so introduce spurious effects. We therefore limit our investigation to M ∼>M⋆ in this particular
cosmological model. Further tests of the effects of particle number, timestep size, and gravitational
softening are given by Tormen et al. (1996). Their results confirm that the numerical parameters
chosen here are adequate to give stable and accurate results.
3. Results
3.1. Time Evolution
Figure 1 illustrates the time evolution of halos of different mass selected from the Ω0 = 1,
n = −1 series. Time runs from top to bottom and mass increases from left to right. The box size
in each column is chosen so as to contain always approximately the same number of particles. The
redshifts of each snapshot have been chosen so that the nonlinear mass M⋆ increases by factors of
4 from z2 to z1 and from z1 to z0 = 0. This figure illustrates convincingly that low-mass halos
complete their assembly earlier than more massive systems.
– 7 –
3.2. Density Profiles
Figure 2 shows spherically averaged density profiles at z = 0 for one of the least and one of
the most massive halos for each set of cosmological parameters. These halos span almost four
orders of magnitude in mass in the case of the CDM models, and about two orders of magnitude
in mass in the power-law models. Radial units are kpc for CDM models (scale at the top), and are
arbitrary in the power-law panels. Density is in units of 1010M⊙/kpc3 in the CDM models and
in arbitrary units in the others. Solid lines are fits to each halo profile using eq. 1. This simple
formula provides a good fit to the structure of all halos over about two decades in radius, from the
gravitational softening (indicated by arrows in Figure 2) to about the virial radius. The quality
of the fit is essentially independent of halo mass or cosmological model and implies a remarkable
uniformity in the equilibrium structure of dark matter halos in different hierarchical clustering
models.
The solid and dashed lines in Figure 3 show the profile fits of Figure 2 but with the radius
scaled to the virial radius of each halo. This scaling removes the intrinsic dependence of size
on mass (more massive halos are bigger) and allows a meaningful comparison between halos of
different mass. From the definition of virial radius, the “concentration” of a halo, c = r200 /rs , and
the characteristic density, δc , are linked by the relation
δc =
200
3
c3
[ln(1 + c) − c/(1 + c)]
.
(2)
Thus at given halo mass (specified by M200 ), there is a single free parameter in eq. 1, which may
be expressed either as the characteristic density, δc , or as the concentration parameter, c. This free
parameter varies systematically with mass; Figure 3 shows that c and δc decrease with increasing
halo mass.
A universal density profile implies a universal circular velocity profile, Vc(r) = (GM (r)/r)1/2 .
This is illustrated in Figure 4, where we plot Vc -profiles for the same systems shown in Figure 2.
As in Figure 3, radii are plotted in units of the virial radius; circular speeds have been normalized
to the value at the virial radius, V200 . The circular velocity curve implied by eq. 1 is
V200 (cid:19)2
(cid:18) Vc (r)
where x = r/r200 is the radius in units of the virial radius. Circular velocities rise near the
center, reach a maximum (Vmax ) at xmax ∼ 2/c, and decline near the virial radius. More centrally
concentrated halos (higher δc or higher c) are characterized by higher values of Vmax /V200 . The
dashed lines in Figure 4 show plots of eq. 3 with parameter values derived from the fits to the
density profiles of Figure 2. The dotted lines are fits using a Hernquist (1990) model constrained
ln(1 + cx) − (cx)/(1 + cx)
ln(1 + c) − c/(1 + c)
(3)
=
1
x
,
– 8 –
to match the location of the maximum of the Vc -curve. The two fits are indistinguishable near
the center, but the Hernquist model underestimates Vc near the virial radius. This disagreement
becomes more pronounced in lower mass systems, for which δc and Vmax /V200 are larger.
3.3. The mass dependence of halo structure
The mass-density dependence pointed out above is further illustrated in Figure 5, where we
plot δc versus mass (expressed in units of M⋆ ) for all the systems in each series. An equivalent
plot, illustrating the mass dependence of the concentration, c, is shown in Figure 6. (The panel
on the upper-left, corresponding to the SCDM model, is equivalent to Figure 7 of Navarro, Frenk
& White 1996). The characteristic density of a halo increases towards lower masses in all the
cosmological models considered. This result supports the idea that the M200 -δc relation is a direct
result of the higher redshift of collapse of less massive systems, and suggests a simple model to
describe the mass-density relation. This model assigns to each halo of mass M (identified at z = 0)
a collapse redshift, zcoll (M , f ), defined as the time at which half the mass of the halo was first
contained in progenitors more massive than some fraction f of the final mass. With this definition,
zcoll can be computed simply using the Press-Schechter formalism (e.g. Lacey & Cole 1993),
δcrit (zcoll ) − δ0
0 (M )) (cid:19) =
erfc(cid:18)
crit
q2(∆2
0 (f M ) − ∆2
where ∆2
0 (M ) is the linear variance of the power spectrum at z = 0 smoothed with a top-hat filter
of mass M , δcrit (z ) is the density threshold for spherical collapse by redshift z , and δ0
crit = δcrit (0).
[This definition can be extended to halos identified at any redshift z0 by replacing δ0
crit by δcrit (z0 )
in eq. 4.] Assuming the characteristic density of a halo to be proportional to the density of the
universe at the corresponding zcoll then implies
(4)
1
2
,
δc (M f ) = C Ω0 (1 + zcoll (M , f ))3
(5)
where C is a proportionality constant which might, in principle, depend on f and on the power
spectrum.
We will see below that f ≪ 1 is needed for this argument to give a good fit to our simulation
0 (f M ) ≫ ∆2
data. In this limit ∆2
0 (M ) and eq. 4 reduces to
δcrit (zcoll ) = δ0
crit + C ′∆0 (f M ),
(6)
where C ′ ≈ 0.7. For f ≪ 1, δcrit (zcoll ) ≫ δ0
crit for all masses in the range of interest, so that
δcrit(zcoll ) ∝ ∆0 (f M ). Since M⋆ (zcoll ) is defined by ∆0 (M⋆ (zcoll )) = δcrit (zcoll ), this equation
implies that the characteristic density of a halo is proportional to the mean density of the universe
at the time when M⋆ ≈ f M , ie. when the characteristic non-linear mass is a fixed small fraction
of the final halo mass. For scale-free models this implies δc ∝ M −(n+3)/2 , the same scaling that
links M⋆ (z ) and the mean cosmic density at redshift z .
– 9 –
Figure 5 shows the correlations predicted from eq. 5 for three values of the parameter f : 0.5,
0.1, and 0.01. The value of the proportionality constant, C (f ), is chosen in each case in order to
match the results of the Einstein-de Sitter simulations for M = M⋆ . These values are given in
Table 1. The same values of C (f ) are used to plot the curves in the panels corresponding to the
low-density models. Some interesting results emerge from inspection of Figure 5 and Table 1.
(i) The agreement between the mass-density dependence predicted by eq. 5 and the results of
the Einstein-de Sitter simulations improves for smaller values of f . This is also true for the
low-density models. Once C (f ) is fixed by matching the results of the Einstein-de Sitter
models, the same value of C (f ) provides a good match to the low density models only if f ∼<0.01.
Interestingly, for f = 0.01 approximately the same value of the proportionality constant,
C ≈ 3 × 103 , seems to fit all our simulations.
(ii) The characteristic density of M⋆ halos decreases systematically for more negative values of
the spectral index n. At M = M⋆ , SCDM halos are the least dense in our Ω0 = 1 series, less
concentrated still than those corresponding to n = −1.5. This is consistent with the general
trend because, according to eqs. 4 and 5, the characteristic density of a halo of mass M⋆ is
controlled by the shape of the power spectrum on scales ∼ f M⋆ . This is about ∼ 1011M⊙ for
f ≈ 0.01 and the effective slope of the CDM spectrum on this mass scale is neff ∼ −2.
(iii) For the power-law models with n = 0 and −1 the characteristic density at a given M /M⋆
increases as Ω0 decreases. Such a trend is plausible since we expect the collapse redshift of
halos of a given mass to increase as Ω0 decreases. On the other hand, halos formed in the
low-density CDMΛ universe are actually less dense than those formed in the standard biased
CDM model because δc depends not only on collapse redshift but also on Ω0 (see eq. 5).
Although reducing Ω0 increases the collapse redshift, the increase in δc from the (1 + zcoll )3
factor can be outweighed by the change in Ω0 . In the CDMΛ model the two effects can combine
to give a reduction in δc as Ω0 decreases. (We remind the reader that δc is defined relative to
the critical density rather than the mean density.)
(iv) Each halo has a characteristic maximum circular speed, Vmax (see eq. 3), which is strongly
correlated with its mass. This is shown in Figure 7, where we also plot least squares fits of the
form M200 ∝ V α
max to the data in each series. Consistent with the trends shown in Figures
5 and 6, the correlation steepens as n increases; we find α ∼ 3.2 – 3.3 for CDM models and
α > 5 for n = 0. Note also that the correlations are extremely tight; the rms scatter in
log M200 is less than ∼ 0.1 in all cases. This is in part a consequence of the generally good
fit of the density profile of eq. 1. The ratio Vmax /V200 increases only logarithmically with the
central concentration of the halo, and changes only by about a factor of two as δc varies by
four orders of magnitude between 103 and 107 . As a result, the M200 -Vmax relation does not
deviate much from the M200 -V200 relation which, by definition, has zero scatter. This has
important consequences for the expected tightness of empirical correlations between mass and
characteristic velocity, such as the Tully-Fisher relation. We intend to return to this issue in
– 10 –
future work.
The results in Figures 5–7 support our conclusion that the characteristic density of a halo
is controlled mainly by the mean matter density of the universe at a suitably defined time of
collapse. One important test of this interpretation is to measure zcoll directly in the simulations
and to compare the result to eq. 5. To do this we identify clumps at every output time using our
friends-of-friends group-finder with linking length set to 20% of the current mean interparticle
separation. We then trace the particles in the most massive clump identified at z = 0 (which
typically has a mean overdensity of ∼ 200) and, at each redshift, add up the total mass in clumps
which contain any of these particles and which are individually more massive than 10% of the final
mass. We identify zcoll with the redshift at which this mass first exceeds half of the final mass.
This is roughly equivalent to the analytic procedure outlined in eq. 4 for f = 0.1. (We decided to
use f = 0.1 rather than f = 0.01 because the smaller value results in very high collapse redshifts,
often before the first output in the simulation.) The main difference is that some of the mass from
the high redshift progenitors ends up outside the virial radius at z = 0. This causes a slight bias
of the measured collapse redshifts towards higher values than the Press-Schechter predictions.
The correlation between δc/Ω0 and (1 + zcoll )3 obtained using this procedure is shown in
Figure 8. All halos in Einstein-de Sitter models are shown with filled circles; those in open
universes with open circles; and those in the CDMΛ model with starred symbols. The solid line is
the relation predicted by eq. 5 for C = 5 × 103 . This is clearly an excellent approximation to the
results of the numerical simulations, and confirms that the mean matter density of the universe
at the time of collapse is the main factor determining the characteristic density of a halo. Note
that the value of the proportionality constant is slightly lower than the values given in Table
1 for f = 0.1. This difference compensates for the slight bias towards higher collapse redshifts
introduced by our numerical procedure.
A summary of our results is presented in Figure 9. The panels on the left compile the fits
(for f = 0.01) to the mass dependence of δc and c in the Einstein-de Sitter models. The panels
on the right compare the Einstein-de Sitter results with those for low-density models. As noted
above, the typical density of M⋆ -halos increases with n. However, the difference between models
becomes less pronounced at higher masses, and is almost negligible at M ∼> 10M⋆ . Halos of a given
M /M⋆ in low density universes can have either lower or higher characteristic densities than their
Einstein-de Sitter counterparts, depending on the competing effects of the collapse redshift and
the value of Ω0 . Note that all the curves in Figure 9 use the same value, f = 0.01, and essentially
the same value of the proportionality constant C ≈ 3 × 103 (see eq. 5). Thus, once calibrated
for an Einstein-de Sitter model, it is possible to apply eqs. 4 and 5 to predict the characteristic
density of halos formed in other hierarchically clustering models. In the Appendix we provide a
detailed description of how to compute δc (M ) numerically for a variety of cosmologies.
– 11 –
3.4. The scatter in the correlations
We examine now the origin of the scatter in the correlations presented above. In particular,
we explore whether at fixed mass the dispersion in the measured values of δc is due to variations
either in the collapse redshift or in the global angular momentum of the system. As shown by
Figures 10 and 11, the bulk of the scatter in δc at a given M /M⋆ can be attributed to small
differences in the redshift of collapse. Figure 10 shows that the δc -deviations from the solid-line fits
in Figure 5 (i.e. those for f = 0.01) correlate strongly with deviations in the redshift of collapse.
(The latter are measured from least-square fits to the M200 vs (1 + zcoll ) correlations measured
directly from the simulations.) Furthermore, the two residuals seem to correlate just as expected
from eq. 5, i.e. ∆ log δc = 3∆ log(1 + zcoll ). (We note that the magnitude of this scatter may not
be fully representative of the dispersion in halo properties corresponding to each cosmological
model, since the sample has been selected so as to minimize departures from equilibrium.) This
relation is indicated by the solid line and is seen to reproduce very well the trend observed in
Figure 10. Figure 11 shows the same δc -residuals of Figure 10, but now as a function of the scatter
in the mass–rotation parameter (M200 vs λ) correlation. (The rotation parameter is defined by
λ = J E 1/2 /GM 5/2 , where J is the angular momentum and E is the binding energy of the halo.
The median λ in our series is ∼ 0.04, in good agreement with previous studies.) We find no
discernible correlation between M /M⋆ and λ or between the δc - and λ-residuals (Figure 11). This
provides further evidence supporting our contention that the redshift of collapse is the primary
factor determining the characteristic density of a halo.
4. Comparison with previous work
The main conclusion of our study, that the shape of halo density profiles is independent of
cosmological context, appears to contradict previous work on this sub ject. As discussed in §1,
a strong dependence of the slope of the density profile on the spectral index n and the density
parameter Ω0 has been established by a number of analytic and numerical studies. We now show
that our results actually include and extend those of previous workers, and we offer an attractive
explanation for some discrepancies found in the literature.
We first address the claim by Quinn et al. (1986), Efstathiou et al (1988), Zurek et al. (1988),
and Warren et al. (1992) that halo density profiles steepen with increasing values of the spectral
index n in an Einstein-de Sitter universe. This claim is based on the discovery that circular
velocity profiles of galaxy-sized halos are relatively flat for n ≈ −2 (or for SCDM; Frenk et al.
1985) but decline progressively faster at large r for larger n. Figure 12 shows that we find the
same trend if we analyze our data in the same fashion as these authors. In this figure we plot, in
linear units, the Vc -profile of an M⋆ halo for different values of n. Each curve has been computed
using eq. 3 and the values of c obtained from the fits presented in Figure 9. Since all these halos
have the same mass, they also have the same virial radius and circular speed, r200 and V200 ,
– 12 –
respectively, which we have used as normalizing factors. The linear units in Figure 12 obscure the
fact that the form of the curves is the same in all cases, and encourage one to conclude, as did
previous authors, that halo rotation curves steepen with increasing n. In fact this trend is present
because M⋆ halos collapse earlier, and so are denser, for larger values of n.
We note that the trend with n in Figure 12 is sensitive to the choice of halo mass. Had we
used halos with M < M⋆ the trend would have been stronger. On the other hand, very massive
halos (M ∼>10 M⋆ ) have similar characteristic densities, irrespective of n (see Figure 9). Thus,
had we plotted very massive halos in Figure 12 we would have concluded that the density profile
depends only weakly on n (at least for Ω0 = 1). This is reminiscent of the claim by West et al.
(1987) that the structure of galaxy cluster halos is independent of n in Einstein-de Sitter universes.
Our analysis suggests an explanation for this apparently discrepant result. By considering only
the most massive systems in their simulations, West et al. focused on a mass range where the
dependence of δc on n is minimal, and so concluded (correctly) that cluster profiles depend very
weakly on initial power spectrum.
A similar explanation accounts for the weak dependence of halo density profile on spectral
index n and on Ω0 reported by Crone et al. (1994). These authors chose to combine the 35 most
massive clumps in each of their cosmological simulations in order to produce an “average” density
profile for each value of n and Ω0 . They then fitted power laws of the form ρ(r) ∝ rγ to these
profiles in the radial range corresponding to density contrasts between 100 and 3000. For Ω0 = 1,
this procedure yielded γ values that decrease from −2.2 to −2.5 as n increases from −2 to 0 (see
the curve labeled “EdS” in their Figure 4). Our results show this weak trend to be a consequence
of the averaging procedure they adopted. The shape of the halo mass function depends strongly
on n and is increasingly peaked around M ≈ M⋆ for larger values of n. Thus, combining the 35
most massive clumps in a simulation results in different “effective” masses for different values of
n, with a bias towards larger values of M /M⋆ for more negative values of n. Applying the same
selection procedure as Crone et al. to our own cosmological simulations we find that the median
mass of these halo ensembles increases from ∼ M⋆ for n = 0 to ∼ 6M⋆ for n = −1.5. As illustrated
in Figure 13, fitting power-laws to the density profiles of these “average” halos results in a weak
steepening with n similar to the trend reported by Crone et al. The slopes of γ ∼ −3 which they
found for low density models are also easily understood, since in these cases they fit a radial range
which extends well outside our nominal virial radius r200 .
Our results are also in agreement with recent work by Cole & Lacey (1996) and Tormen et
al. (1996), who analyzed the structure of massive halos (M > M⋆ ) in scale-free universes. The
general correlations with mass which they report (more massive halos tend to be less centrally
concentrated than less massive halos) agree well with the results we present in §3. At a given halo
mass, however, Cole & Lacey’s halos are significantly less concentrated than ours. For example,
for M⋆ halos they find c ∼ 12 and ∼ 17 for n = −1 and 0, respectively (Ω0 = 1), compared with
c ∼ 17 and ∼ 30 in our simulations (see connected symbols in the lower-right panel of Figure 9).
On the other hand, the results of Tormen et al. (1996) are in very good agreement with ours;
– 13 –
they find c ∼ 10 for a 10M⋆ halo formed in an n = −1, Ω0 = 1 universe, which compares well with
c ∼ 9 that we find for a similar system.
As discussed in detail by Tormen et al, the discrepancy between our results and those of Cole
& Lacey is likely to be due in part to the poorer numerical resolution of their simulations. Indeed,
their simulations used gravitational softenings that are 2-3 times larger than ours and timesteps
which are also significantly longer than the typical values in our simulations. Both of these
effects can artificially lower the central concentration of a halo. A concurrent factor may be the
averaging procedure adopted by Cole & Lacey. These authors constructed average density profiles
by co-adding all halos of similar mass identified in their cosmological simulations, regardless of
their dynamical state. This sample contains a number of unrelaxed systems and ongoing mergers
where substructure and double centers can bias the results of the fitting procedure towards lower
concentrations.
We can test directly for the effects of these various factors by applying the averaging procedure
of Cole & Lacey to halos identified in our own P 3M cosmological simulations (see §2.2.1) and
comparing with the results of the individual halo runs. We restrict this comparison to the most
massive halos in each run, M ∼> 10M⋆ , since these systems have a large number of particles
and comparable numerical resolution to that of Cole & Lacey. The comparison confirms that
the central concentration of halos can be significantly underestimated as a result of the factors
mentioned above. The magnitude of the effect is sensitive to n and Ω0 ; the concentration c can be
underestimated by up to a factor of ∼ 3 for n = 0 and Ω0 = 0.1, but by less than ∼ 1.5 for n = −1
and Ω0 = 1. We conclude that the disagreement between our results and those of Cole & Lacey is
likely to be the result of the combined effect of their halo selection and averaging procedures and
their poorer numerical resolution.
5. Discussion
Our results suggest that equilibrium dark halos formed through dissipationless hierarchical
clustering have density profiles with a universal shape which does not depend on their mass,
on the power spectrum of initial fluctuations, or on the cosmological parameters Ω0 and Λ. It
appears that mergers and collisions act during halo formation as a “relaxation” mechanism to
produce an equilibrium which is largely independent of initial conditions. This mechanism must
operate rapidly since the similarity between profiles extends to the virial radius. These properties
are characteristic of the “violent relaxation” process proposed by Lynden-Bell (1967) to explain
the regularities observed in the structure of elliptical galaxies, and it is interesting that our
universal profile is similar to the Hernquist profile which gives a good description of elliptical
galaxy photometry. The two differ significantly only at large radii, perhaps because ellipticals
are relatively isolated systems whereas dark halos are not. Syer & White (1996) suggest that a
universal profile may be understood as a fixed point for a process of repeated mergers between
unequal ob jects. Their analysis of this process predicts a dependence on initial power spectrum
– 14 –
which seems stronger, however, than that seen in our numerical data.
Our simulations suggest that the density profile of an isolated equilibrium halo can be
specified quite accurately by giving two parameters; halo mass and halo characteristic density.
Furthermore, in any particular hierarchical model these parameters are related in such a way that
the characteristic density is proportional to the mean cosmic density at the time when the mass of
typical nonlinear ob jects was some fixed small fraction of the halo mass. The characteristic density
thus reflects the density of the universe at the collapse time of the ob jects which merge to form
the halo core. With this interpretation we are able to fit the mass-density relation for equilibrium
halos in all the cosmological models we have considered. In addition, it is possible to calculate the
mass-density relation in any other hierarchical model (see Appendix). It is difficult to imagine
a simpler situation – halos of all masses in all hierarchical cosmologies look the same, and their
characteristic densities are just proportional to the cosmic density at the time they “formed”.
Our results extend the radial range over which dark halo structure can reliably be determined
to more than two decades. The central regions of our models have densities of order 106 ρcrit ,
comparable to those in the luminous parts of galaxies and in the central cores of galaxy clusters.
As a result a variety of direct observational tests of our predictions are available. In Navarro
et al. (1996) we discussed several of these in the context of the standard CDM cosmogony –
rotation curves of giant and dwarf galaxies, satellite galaxy dynamics, hot gaseous atmospheres
around galaxies and in clusters, strong and weak gravitational lensing – all provide interesting
constraints. We will not pursue these issues here, however, because they would take us too far
from our primary goal, namely the presentation of a simple and apparently general theoretical
result: hierarchical clustering leads to a universal halo density profile just as it leads to universal
distributions of halo axial ratios and halo spins; none of these properties depends strongly on
power spectrum, on Ω, or on Λ [see Cole & Lacey (1996) and references therein]. Comparison
of the predicted halo structure with observation should, of course, provide strong constraints on
the parameters which define particular hierarchical cosmogonies, and perhaps on the hierarchical
clustering paradigm itself. We expect to come back to these issues in future work.
We are grateful for the hospitality of the Institute for Theoretical Physics of the University
of California at Santa Barbara, where some of the work presented here was carried out. JFN is
also grateful for the hospitality of the Max Planck Institut fur Astrophysik in Garching, where
this pro ject was started. In addition he would like to acknowledge useful discussions with Cedric
Lacey. Special thanks are due to Shaun Cole for making the data shown in Figure 9 available in
electronic form. This work was supported in part by the PPARC and by the NSF under grant
No. PHY94-07194 to the Institute for Theoretical Physics of the University of California at Santa
Barbara.
– 15 –
A.
Step-by-step calculation of the density profile of a dark matter halo
In this Appendix we describe in detail the calculation of the parameters that specify the
density profile of a dark halo of mass M . This calculation is applicable to Einstein-de Sitter
(Ω0 = 1, Λ = 0), open (Ω0 < 1, Λ = 0), and flat (Ω0 + Λ = 1) universes. We provide approximate
fitting formulae that are valid for power-law and CDM initial density fluctuation spectra.
A halo of mass M identified at z = z0 can be characterized by its virial radius,
h−1M⊙ (cid:19)1/3(cid:18) Ω0
Ω(z0 ) (cid:19)−1/3
r200 = 1.63 × 10−2(cid:18) M
or by its circular velocity,
(1 + z0 )−1 h−1kpc,
(A1)
r200 (cid:19)1/2
Ω(z0 ) (cid:19)1/2
= (cid:18) r200
V200 = (cid:18) GM
h−1kpc (cid:19) (cid:18) Ω0
The density profile of this system is fully specified by its characteristic density δc and is given by
(see eq.1)
(1 + z0 )3/2 km/s.
(A2)
(A3)
3H 2
0
8πG
ρ(r) =
(1 + z0 )3 Ω0
Ω(z0 )
δc
cx(1 + cx)2 ,
where x = r/r200 and c is the concentration parameter, a function of δc given in eq. 2. The
corresponding circular velocity profile, Vc (r), is given by
V200 (cid:19)2
(cid:18) Vc (r)
using the concentration c as a parameter.
ln(1 + cx) − (cx)/(1 + cx)
ln(1 + c) − c/(1 + c)
1
x
=
,
(A4)
The characteristic density is determined by the collapse redshift zcoll , which is given by (see
eq. 4)
=
= 1 +
δcrit(zcoll )
δcrit (z0 )
δ0
crit(Ω(zcoll ), Λ)
δ0
crit (Ω(z0 ), Λ)
D(z0 , Ω0 , Λ)
D(zcoll , Ω0 , Λ)
0.477
δcrit(z0 ) q2(∆2
0 (f M ) − ∆2
0 (M ))
Here D(z , Ω0 , Λ) is an Ω-dependent linear growth factor that can be written as
D(z , Ω0 , Λ) =
where we have used the following auxiliary definitions,
1/(1 + z )
F1 (w)/F1 (w0 )
F2 (y)F3 (y)/F2 (y0 )F3 (y0 )
if Ω0 = 1 and Λ = 0,
if Ω0 < 1 and Λ = 0,
if Ω0 + Λ = 1,
w0 =
1
Ω0
− 1,
w =
w0
1 + z
,
(A5)
(A6)
(A7)
(A8)
– 16 –
3
u
+
F1 (u) = 1 +
ln [(1 + u)1/2 − u1/2 ],
3(1 + u)1/2
u3/2
y0 = (2w0 )1/3 ,
y0
1 + z
(u3 + 2)1/2
u3/2
u′3 + 2 (cid:19)3/2
F3 (u) = Z u
0 (cid:18) u′
A good numerical approximation to the Ω-dependence of the critical threshold for spherical
collapse is given by
F2 (u) =
du′ .
and
y =
,
,
(A13)
(A9)
(A10)
(A11)
(A12)
0.15 (12π)2/3
0.15 (12π)2/3 Ω0.0185
0.15 (12π)2/3 Ω0.0055
if Ω = 1 and Λ = 0,
if Ω0 < 1 and Λ = 0,
if Ω0 + Λ = 1,
crit (Ω, Λ) =
δ0
which can be used to compute δcrit (z0 ) = δ0
crit(Ω(z0 ))/D(z0 , Ω0 , Λ). Finally, eq. A5 requires
∆2
0 (M ), the variance of the power spectrum on mass scale M , extrapolated linearly to z = 0. In
the case of a power-law spectrum of initial density fluctuations, P (k) ∝ kn , this is simply
M⋆ (z = 0) (cid:19)−(n+3)/6
crit(cid:18) M
0 (M ) = δ0
∆2
where we have normalized the spectrum by M⋆ (z = 0), the present nonlinear mass. A CDM
spectrum is usually normalized by σ8 , the rms mass fluctuations within a sphere of radius 8h−1
Mpc, and its variance can be approximated by,
,
(A14)
(A15)
where we have used the following definitions,
∆0 (M ) = σ8F4 (M8 )/F4 (Mh ),
M8 = 6.005 × 1014 (h Ω0 )3 ,
Mh = (cid:18) M
h−1M⊙ (cid:19)h3Ω2
0
and
F4 (u) = A1u0.67 [1 + (A2u−0.1 + A3u−0.63 )p ]1/p ,
with A1 = 8.6594 × 10−12 , A2 = 3.5, A3 = 1.628 × 109 , and p = 0.255.
(A16)
(A17)
(A18)
(A19)
Eqs. A6-A19 can be used to solve eq. A5 and find the collapse redshift, zcoll , corresponding to
a halo of mass M . As noted when discussing Fig. 5, we recommend using f = 0.01 when solving
eq. A5, since this value seems to reproduce well the results of all our numerical experiments.
– 17 –
Once zcoll (M ) has been found, the characteristic density of the halo (expressed in units of the
critical density at z = z0 ) can be computed from (see eq.5 and Table 1),
1 + z0 (cid:19)3
δc (M , z0 ) ∼ 3 × 103 Ω(z0 )(cid:18) 1 + zcoll
where we have assumed f = 0.01. A FORTRAN subroutine that implements the procedure
described here and returns δc (M , z0 ) for all the cosmologies we discuss is available from the authors
upon request.
(A20)
,
– 18 –
P (k)
CDM
CDMΛ
Ω0
1.0
1.0
1.0
0.25
0.25
0.25
Λ
0.0
0.0
0.0
0.75
0.75
0.75
n = −1.5
n = −1.0
n = −0.5
n = 0.0
1.0
1.0
1.0
1.0
1.0
1.0
0.1
0.1
0.1
1.0
1.0
1.0
1.0
1.0
1.0
0.1
0.1
0.1
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
f
0.5
0.1
0.01
0.5
0.1
0.01
0.5
0.1
0.01
0.5
0.1
0.01
0.5
0.1
0.01
0.5
0.1
0.01
0.5
0.1
0.01
0.5
0.1
0.01
C (f )
1.75 × 104
7.44 × 103
3.41 × 103
1.75 × 104
7.44 × 103
3.41 × 103
3.00 × 104
1.08 × 104
3.15 × 103
5.00 × 104
1.38 × 104
2.50 × 103
5.00 × 104
1.38 × 104
2.50 × 103
1.25 × 105
2.81 × 104
2.81 × 103
4.00 × 105
6.66 × 104
4.00 × 103
4.00 × 105
6.66 × 104
4.00 × 103
Table 1: Parameters in eq. 5 used to plot the fits in Figure 5.
– 19 –
REFERENCES
Bertschinger, E. 1985, ApJS, 58, 39.
Cole, S.M., & Lacey, C., 1996, MNRAS, in press (astro-ph 9510147).
Crone, M., Evrard, A.E., & Richstone, D.O. 1994, ApJ, 434, 402.
Dubinski, J. & Carlberg, R. 1991, ApJ, 378, 496.
Efstathiou, G.P., Davis, M., Frenk, C.S., & White, S.D.M. 1985, ApJS, 57, 241.
Efstathiou, G.P., Frenk, C.S., White, S.D.M., & Davis, M. 1988, MNRAS, 235, 715.
Fillmore, J.A., & Goldreich, P. 1984, ApJ, 281, 1.
Frenk, C.S., White, S.D.M., Efstathiou, G.P., and Davis, M. 1985, Nature, 317, 595.
Frenk, C.S., White, S.D.M., Davis, M., and Efstathiou, G.P. 1988, ApJ, 327, 507.
Gunn, J., & Gott, J.R. 1972, ApJ, 176,1.
Hernquist, L. 1990, ApJ, 356, 359.
Hoffman, Y. 1988, ApJ, 328, 489.
Hoffman, Y., & Shaham, J. 1985, ApJ, 297, 16.
Lacey, C.G., & Cole, S.M. 1993, MNRAS, 262, 627.
Lynden-Bell, D., 1967, MNRAS, 136, 101.
Navarro, J.F., & White, S.D.M. 1993, MNRAS, 265, 271.
Navarro, J.F., Frenk, C.S., & White, S.D.M. 1995, MNRAS, 275, 720.
Navarro, J.F., Frenk, C.S., & White, S.D.M. 1996, ApJ, 462, 563.
Quinn, P.J., Salmon, J.K., & Zurek, W.H. 1986, Nature, 322, 329.
Syer, D., & White, S.D.M. 1996, MNRAS, submitted (astro-ph 9611065).
Tormen, G., Bouchet, F., & White, S.D.M., 1996, MNRAS, submitted (astro-ph 9603132).
Warren, M.S., Quinn, P.J., Salmon, J.K., & Zurek, W.H. 1992, ApJ, 399, 405.
West, M.J., Dekel, A., & Oemler, A., 1987, ApJ, 316, 1.
White, S.D.M. 1996, in Les Houches Session LX: Cosmology and Large-scale Structure, ed.
Schaeffer, R. Silk, J., Spiro, M., Zinn-Justin, J., North Holland, p. 349.
Zurek, W.H., Quinn, P.J., & Salmon, J.K. 1988, ApJ, 330, 519.
– 20 –
This preprint was prepared with the AAS LATEX macros v4.0.
– 21 –
Fig. 1.— Particle plots illustrating the time evolution of halos of different mass in an Ω0 = 1,
n = −1 cosmology. Box sizes of each column are chosen so as to include approximately the same
number of particles. At z0 = 0 the box size corresponds to about 6 × r200 . Time runs from top to
bottom. Each snapshot is chosen so that M⋆ increases by a factor of 4 between each row. Low mass
halos assemble earlier than their more massive counterparts. This is true for every cosmological
scenario in our series.
– 22 –
Fig. 2.— Density profiles of one of the most and one of the least massive halos in each series. In
each panel the low-mass system is represented by the leftmost curve. In the SCDM and CDMΛ
models radii are given in kpc (scale at the top) and densities are in units of 1010M⊙/kpc3 . In all
other panels units are arbitrary. The density parameter, Ω0 , and the value of the spectral index, n
is given in each panel. Solid lines are fits to the density profiles using eq. (1). The arrows indicate
the value of the gravitational softening. The virial radius of each system is in all cases two orders
of magnitude larger than the gravitational softening.
– 23 –
Fig. 3.— The fits to the density profiles of Figure 2, scaled to the virial radius, r200 , of each
system and to the critical density of the universe at z = 0. Solid and dashed lines correspond to the
low- and high-mass systems, respectively. Note that low-mass systems are denser than high-mass
systems near the center, indicating that the characteristic density of a halo increases as the halo
mass decreases.
– 24 –
Fig. 4.— The circular velocity profiles of the halos shown in Figure 2. Radii are in units of the
virial radius and circular speeds are normalized to the value at the virial radius. The thin solid line
shows the data from the simulations. All curves have the same shape: they rise near the center
until they reach a maximum and then decline at the outer edge. Low mass systems have higher
maximum circular velocities in these scaled units because of their higher central concentrations.
Dashed lines are fits using eq.(3). The dotted lines are the fit to the low-mass halo in each panel
using a Hernquist profile. Note that this model fits rather well the inner regions of the halos, but
underestimates the circular velocity near the virial radius.
– 25 –
Fig. 5.— The correlation between the mass of a halo and its characteristic density. Masses are
given in units of the nonlinear mass scale M⋆ (see text for a definition). Densities are relative to
the critical value. Three curves are shown in each panel for different values of the parameter f (see
eq. 5). The fits are normalized to intersect at M200 = M∗ in the case Ω = 1. This normalization
is then used for the low-density models (Ω0 < 1). Note that for f = 0.01 this procedure results in
good fits to the results of the simulations in all cases.
– 26 –
Fig. 6.— As Figure 5, but for the concentration parameter c.
– 27 –
Fig. 7.— The mass dependence of the maximum circular velocity of a halo. Velocity units are
arbitrary in the power-law panels and km s−1 in the CDM models (scale at top). Mass is in units
of 1010M⊙ for CDM halos and in units of M⋆ in the other panels. Power law fits of the form
M ∝ V α
max are shown. The value of α and the rms scatter in the mass about the fit are indicated
in each panel. Note that the M -Vmax dependence steepens for larger values of the spectral index
n. The effect of the cosmological parameters on α seems to be rather small.
– 28 –
Fig. 8.— The characteristic density of all halos in our series as a function of the redshift at which
half of the final mass is in collapsed progenitors more massive than 10% of the final mass. Solid
circles correspond to all our runs with Ω0 = 1, open circles to our runs with Ω0 = 0.1 and Λ = 0,
and starred symbols to the CDMΛ runs (Ω0 = 0.25, Λ = 0.75). The solid line shows the “natural”
scaling, δc ∝ Ω0 (1 + z )3 expected if the characteristic density of a halo is directly proportional to
the mean matter density of the universe at the time of collapse.
– 29 –
Fig. 9.— A summary of the mass dependence of δc and c in our different cosmological models.
Mass is given in units of the nonlinear mass scale M⋆ . Curves are labeled in the upper plots and
the same line types are used in the bottom plots. The symbols in the lower-right panel show the
correlation between c and mass found by Cole & Lacey (1996), with open squares for n = −1,
Ω0 = 1 and solid squares for n = 0, Ω0 = 1. These results should be compared to the lower dashed
and lower dot-dashed curves in the same panel, respectively. Because of their poorer numerical
resolution, Cole & Lacey’s halos are significantly less concentrated than the ones in our study.
– 30 –
Fig. 10.— The δc -deviation from the fits shown in Figure 5 (for f = 0.01) plotted versus the
deviation from the mean relation between the collapse redshift, 1 + zcoll (Figure 8), and the mass of
a system. Note that systems assembled earlier (later) than the average tend to have characteristic
densities above (below) the mean. The correlation between the δc and 1 + zcoll residuals follows
the “natural” scaling, ∆ log δc = 3∆ log(1 + zcoll ), as shown by the solid line. Only the results
corresponding to Ω0 = 1 are shown.
– 31 –
Fig. 11.— The δc -residuals in Figure 5 relative to the solid curve fit (f = 0.01) plotted versus the
residuals in the M200 –λ correlation. This figure shows that the scatter in δc at a given mass cannot
be attributed to the effects of rotation. All halos with Ω0 = 1 are included in this plot.
– 32 –
Fig. 12.— Circular velocity profiles of M⋆ halos formed in simulations with different power spectra
in an Einstein-de Sitter universe. The curves were computed using eq. 3 and the values of the
concentration c obtained from the fits in Figure 9. Radii are in units of the virial radius and
circular speeds in units of the circular velocity at the virial radius. Note that, because of the use
of linear units, the fact that all curves have the same shape (eq. 3) is not immediately apparent.
– 33 –
Fig. 13.— A comparison between our density profiles and the power-law fits of Crone et al. (1994).
The region fitted by these authors, corresponding to densities between 100 and 3000, is shown by
the dotted lines. Over this region our results (solid lines) and the power-law parameterization
adopted by Crone et al.
(dashed lines) are consistent. The profiles shown by the solid curves
correspond to M = 6M∗ (n = −1.5), M = 4M∗ (n = −1), M = 2M∗ (n = −0.5), and M = M∗
(n = 0), as discussed in the text.
|
astro-ph/0501138 | 1 | 0501 | 2005-01-08T02:09:18 | The IMF Long Ago and Far Away: Faint Stars in the Ursa Minor dSph Galaxy | [
"astro-ph"
] | The dwarf spheroidal galaxy in Ursa Minor is apparently dark-matter dominated, and is of very low surface brightness, with total luminosity only equal to that of a globular cluster. Indeed its dominant stellar population is old and metal-poor, very similar to that of a classical halo globular cluster in the Milky Way Galaxy. However, the environment in which its stars formed was clearly different from that in the globular clusters in the Milky Way Galaxy -- what was the stellar IMF in this external galaxy a long time ago? The fossil record of long-lived, low-mass stars contains the luminosity function, derivable from simple star counts. This is presented here. The mass function requires a robust mass-luminosity relation, and we describe the initial results to determine this, from our survey for eclipsing low-mass binaries in old open clusters. The massive star IMF at early times is constrained by elemental abundances in low-mass stars, and we discuss the available data. All data are consistent with an invariant IMF, most probably of Salpeter slope at the massive end, with a turnover at lower masses. | astro-ph | astro-ph |
THE IMF LONG AGO AND FAR AWAY:
Faint Stars in the Ursa Minor dSph Galaxy
Rosemary F.G. Wyse
Johns Hopkins University
Department of Physics & Astronomy
[email protected]
Abstract
The dwarf spheroidal galaxy in Ursa Minor is apparently dark-matter
dominated, and is of very low surface brightness, with total luminosity
only equal to that of a globular cluster.
Indeed its dominant stellar
population is old and metal-poor, very similar to that of a classical halo
globular cluster in the Milky Way Galaxy. However, the environment
in which its stars formed was clearly different from that in the globular
clusters in the Milky Way Galaxy -- what was the stellar IMF in this
external galaxy a long time ago? The fossil record of long-lived, low-
mass stars contains the luminosity function, derivable from simple star
counts. This is presented here. The mass function requires a robust
mass-luminosity relation, and we describe the initial results to deter-
mine this, from our survey for eclipsing low-mass binaries in old open
clusters. The massive star IMF at early times is constrained by elemen-
tal abundances in low-mass stars, and we discuss the available data. All
data are consistent with an invariant IMF, most probably of Salpeter
slope at the massive end, with a turnover at lower masses.
Introduction
The stellar IMF at high redshift is of great importance for a wide range
of astrophysical problems, such as the ionization and enrichment of the
intergalactic medium, the extragalactic background light, the visibility
of galaxies and the rate at which baryons are locked-up into stars and
stellar remnants. There are two complementary approaches to the de-
termination of the IMF long ago and far away: one is to observe directly
high redshift objects, and attempt inferences on the stellar IMF from the
integrated spectrum and photometry, while the second approach anal-
yses the fossil record in old stars at low redshift. The characterization
of the stellar IMF in external galaxies, compared to that in the Milky
1
2
Way, is a crucial step in deciphering the important physical processes
that determine the distribution of stellar masses under a range of differ-
ent physical conditions. The low mass stellar IMF at the high redshifts
at which these stars formed is directly accessible through star counts,
plus a mass-luminosity relation. The high mass IMF at these high red-
shifts is constrained by the chemical signatures in the low mass stars
that were enriched by the supernovae from the high mass stars. I will
discuss both ends of the IMF at high redshift, in an external galaxy.
1.
Extremely Old Stars
Simulations of galaxy formation within the framework of the 'concor-
dance' (Λ)CDM cosmology agree that the first stars form within struc-
tures that are less massive than a typical L∗ galaxy today (e.g. Kauff-
mann, White & Guiderdoni 1993; Cole et al. 2000). Large galaxies form
hierarchically, through the merging and assimilation of such smaller sys-
tems. Satellite galaxies of the Milky Way are survivors of this merging
(e.g. Bullock, Kravtsov & Weinberg 2000). The stars that formed at
early times are found, at the present day, throughout large galaxies, and
also in satellite galaxies. Environments with little subsequent star for-
mation are the best places to find and study old stars -- the stellar halo
of the Milky Way, and a few of the dwarf spheroidal satellite galaxies.
1.1
The Ursa Minor Dwarf Spheroidal Galaxy
The dwarf spheroidal galaxy in Ursa Minor (UMi dSph), like all mem-
bers of its morphological class (Gallagher & Wyse 1994), has extremely
low surface brightness, with a central value of only ∼ 25.5 V mag/sq. arc-
sec, or ∼ 2.5 L⊙/pc−2. The total luminosity is in the range 2 − 4 ×
105 LV,⊙ (Kleyna et al. 1998; Palma et al. 2003), equal to that of a lu-
minous Galactic globular cluster. Again similar to a globular cluster,
the Ursa Minor dSph contains little or no gas and has apparently not
formed a significant number of stars for ∼ 12 Gyr (e.g. Hernandez et
al. 2000; Carrera et al. 2002), or since a redshift z >
∼ 2. The metallicity
distribution of the stars is narrow, with a mean of [Fe/H] ∼ −1.9 dex
and a dispersion of ∼ 0.1 dex (e.g. Bellazzini et al. 2002). The stellar
line-of-sight velocity dispersion is ∼ 10 km/s (e.g. Wilkinson et al. 2004),
sufficiently large that, unlike globular clusters, equilibrium models have
>
∼ 50 (M/L)V,⊙, perhaps as high as
a large mass-to-light ratio, (M/L)V
∼ 108 M⊙ (Wilkinson et
several hundred in solar units if the mass is >
al. 2004), indicating a non-baryonic dark halo. Non-equilibrium models
are rather contrived and themselves fail to explain the data (e.g. Wilkin-
son et al. 2004). Models of the evolution of dwarf spheroidals are by no
The IMF Long Ago
3
means well developed, but very likely the stars formed in an environ-
ment rather different than that of globular cluster stars, or of current
star-forming regions in the disk of the Milky Way.
The distance of the Ursa Minor dSph is only ∼ 70 kpc, close enough
that a determination of the luminosity function of low-mass main se-
quence stars through star counts is feasible, particularly using the Hub-
ble Space Telescope. The (unusually) simple stellar population of this
dwarf spheroidal -- essentially of single age, single metallicity -- makes
the derivation of the luminosity function from star counts a robust pro-
cedure. This determines the low-mass stellar IMF at redshifts of >
∼ 2.
High-resolution spectroscopy of the luminous evolved stars is also possi-
ble, yielding elemental abundances which constrain the high-mass stellar
IMF that enriched the low-mass stars we observe.
2.
The Faint Stellar Luminosity Function and
Mass Function at Redshift >
∼
The (very) dominant stellar population in the UMi dSph is very simi-
lar to that of a classical Galactic halo globular cluster. The most robust
constraint on the low-mass stellar IMF is then obtained by a direct com-
parison between the faint stellar luminosity functions of the UMi dSph
and of representative globular clusters of the same age and metallicity,
such as M15 and M92, observed in the same bandpasses, same telescope
and detector. With the same stellar populations, this is a comparison
between mass functions, and differences may be ascribed to variations
in the low-mass IMF.
We therefore obtained deep images with the Hubble Space Telescope
in a field close to the center of the UMi dSph, using STIS as the primary
instrument (optical Long Pass filter), with WFPC2 (V606 and I814 filters)
and NICMOS (NIC2/H-band) in parallel. The WFPC2 filters matched
those of extant data for M15 and M92; we obtained our own STIS/LP
and NIC2/H-band data for M15. Similarly exposed data for an 'off'
field, at 2 -- 3 tidal radii from the centre of UMi dSph, were also acquired.
The detailed paper presenting the results from the full dataset is Wyse
et al. (2002); preliminary results from a partial WFPC2 dataset were
presented in Feltzing, Wyse & Gilmore (1999).
The images are not crowded and standard photometric techniques
were used to derive the luminosity functions. For the WFPC2 data,
the luminosity functions were based only upon stars (unresolved ob-
jects) that lie close to the well-defined UMi dSph main sequence locus
in the colour-magnitude diagram (CMD; see Figure 1). The 'off' field
CMD confirmed little contamination from Galactic stars. The STIS lu-
4
minosity function was based on one band only, and we employed various
approaches to background subtraction. The NICMOS data served only
to exclude a hypothetical population of very red stars and will not be
discussed further here. The extant WFPC2 data for the globular clusters
(Piotto et al. 1997) are from fields at intermediate radii within the clus-
ters, where effects of dynamical evolution on the mass function should
be minimal.
Figure 1
(2002).
From Wyse et al.
Colour-
magnitude diagram for all UMi dSph
stars from the three wide-field cameras on
WFPC2. The full curves delineate the se-
lection criteria for stars to be included in
the luminosity function. The error-bars in
each magnitude bin are shown at the right.
The well-defined narrow main sequence is
the main feature. The few blue stars close
to the turn-off are probably blue stragglers,
rather than younger stars. The distribu-
tion of stars across the main sequence is
asymmetric, to the red, and is consistent
with a normal population of binary stars.
2.1
Comparisons with Globular Clusters
The comparisons with the WFPC2 colour-magnitude based V-band
and I-band luminosity functions are shown in Figure 2. We adopted
0.5 mag bins to have reasonable numbers in each bin, and to mini-
mize effects of e.g. reddening and distance moduli uncertainties. The
50% completeness limits for the UMi dSph data are equivalent to ab-
solute magnitudes of M606 = +9.1 and M814 = +8.1, which using the
Baraffe et al. (1997) models both correspond to masses of M ∼ 0.3 M⊙.
The STIS/LP data provide an independent check, by both a direct LP-
luminosity function comparison between M15 and UMi dSph, and a de-
rived STIS-based I-band luminosity function. All show that the globular
cluster stars and the UMi stars have indistinguishable faint luminosity
functions, down to an equivalent mass limit of ∼ 0.3 M⊙ (see Wyse et
al. 2002 for details).
We employed various statistical tests to quantify the agreement of the
various datasets -- e.g. STIS-derived I-band vs. WFPC2 I-band etc:
♥ Linear, least-square fits to the (log) counts as a function of apparent
magnitude, using various ranges of magnitude and differing bin choices,
consistently found agreement to better than 2σ.
The IMF Long Ago
5
Based on figures in Wyse et al. (2002). Comparisons between the
Figure 2.
completeness-corrected Ursa Minor luminosity functions (histograms; 50% complete-
ness indicated by the vertical dotted line) in the V-band (left panel) and the I-band
(right panel) and the same for M92 (filled circles) and M15 (open triangles) (both
taken from Piotto et al. 1997, renormalized and shifted to the same distance as the
Ursa Minor dSph). The luminosity functions for the globular clusters and the dwarf
spheroidal galaxy are indistinguishable.
♥ Kolmogorov-Smirnov tests on the unbinned data for a variety of
magnitude ranges; the results depend on systematics such as the relative
distance moduli, but again there is general agreement to better than 5%
significance level.
♥ χ−square tests were carried out on the binned data, using a variety
of bin centers (maintaining 0.5 mag bin widths) and magnitude ranges
and again agreement to better than 5% significance level.
The main result is that the underlying mass functions of low-mass
stars in Galactic halo globular clusters and in the external galaxy the
UMi dSph are indistinguishable. This is a comparison between two
different galaxies, and systems of very different baryonic densities and
dark matter content.
3.
Low-Mass Stellar Mass Functions
Adopting the Baraffe et al. (1997) models, the 50% completeness lim-
its for the luminosity functions of the stars in the UMi dSph correspond
to ∼ 0.3 M⊙, and the mass function may be fit by a power law, with
slope somewhat flatter than the Salpeter (1955) value, over the range
we test, of 0.3 <
<
∼ 0.8. This is consistent with the solar neigh-
bourhood mass function over this mass range, and indeed the universal
mass function that appears to be the conclusion of this meeting.
∼ M/M
⊙
However, the light-to-mass transformation is not robustly defined for
K/M dwarfs, especially as a function of age and metallicity. Calibration
of this is best achieved by analysis of low-mass stars in detached eclipsing
binary systems, and we have recently undertaken a photometric survey
6
of open clusters to identify candidate low-mass binary systems to be
followed up with spectroscopy for radial velocity curves; this forms the
PhD thesis of Leslie Hebb at Johns Hopkins University.
Our sample consists of six open clusters of known age and metallicity
(from the brighter turn-off stars), old enough to have low-mass stars on
∼ 2 × 108 yr, with the oldest being ∼ 4 Gyr. We
the main sequence, age >
used both the Wide Field Camera on the 2.5m Isaac Newton Telescope
and the Mosaic Camera on the Kitt Peak 4m telescope, each of which
provide a field of view of ∼ 35′ × 35′. The observing strategy we adopted
was designed to enable the detection of a 0.05 mag amplitude eclipse in
a target 0.3 M⊙ star, monitored on timescales of fraction of an hour,
hours and days. The low probability of eclipse means that populous
clusters must be observed for many days. We expect our survey to find
3 -- 5 low-mass eclipsing systems. The details of the survey are presented
in Hebb, Wyse & Gilmore (2004).
The phase-folded lightcurve for a candidate M-dwarf eclipsing system in
Figure 3.
the open cluster M35, with a period of just over 1 day. The crosses show all differential
photometry measurements for this object collected in January 2002, January 2003 and
February 2003. The solid line is a simple sine-wave fit to the measurements taken
during the primary eclipse. The phase of the secondary eclipse is marked by the same
sine function fit to the primary eclipse, but with half the amplitude.
The imaging data has all been acquired, differential photometry ob-
tained and we are now analyzing the derived light curves. An example
of the light curve of a candidate eclipsing low-mass system is shown in
Figure 3; the candidate was identified by applying a box-fitting algo-
rithm to the photometric time series. Our photometry, plus infrared
data from 2MASS, is consistent with this being an M-dwarf system. We
have applied for follow-up spectroscopic data, together with higher time-
sampling photometric data, for this system and for our other candidates.
4.
High-Mass Stellar Mass Functions
The high-mass stellar mass function long ago can be constrained by
the elemental abundances in the long-lived low-mass stars that formed
The IMF Long Ago
7
from gas that was enriched by the Type II supernovae from the massive
stars of a previous generation (see e.g. review of Wyse 1998). Interpreta-
tion of the pattern of elemental abundances is easiest for low-mass stars
that formed early in a star-formation event, and were enriched by only
massive stars. Most of the dwarf spheroidal companions to the Milky
Way have had extended star formation, and so are expected to show
evidence in the elemental abundances for the incorporation of iron from
long-lived Type Ia supernovae. The Ursa Minor dSph is the best candi-
date for having had a sufficiently short duration of star formation that
a significant fraction of its low-mass stars formed prior to the onset of
Type Ia supernovae in sufficient numbers to be noticed in the chemical
elemental abundances; this timescale is uncertain, but likely to be of the
order of 1 -- 2 Gyr.
The elemental mix produced by a generation of massive stars depends
on the massive-star mass function, because the yields of a given Type
II supernova depends on its mass. In particular, the α-element yields
(nuclei formed by successive addition of a helium nucleus) vary more
strongly with progenitor mass than does the iron yield (e.g. Figure 1
of Gibson 1998). There appears to have been surprisingly good mixing
at early times, at least in the stellar halo of the Milky Way (see the
remarkably low scatter in the ratio of [α/Fe] at [Fe/H] <
∼ − 2.5 in the
sample analysed by Cayrel et al. 2004), so that a well-defined value of
[α/Fe] is produced by a generation of massive stars of given IMF. This
is seen as the 'Type II plateau' in [α/Fe] for metal-poor Galactic stars.
The available elemental abundance data for a handful of individual
stars in the UMi dSph are consistent with the same value for the Type II
plateau as seen in stars of the Milky Way (Shetrone et al. 2001), with
some downturn for more metal-rich stars, as expected if there is an age
spread of 1 -- 2 Gyr and an age-metallicity relationship. The simplest
interpretation is that the high mass IMF was the same in the UMi dSph
as in the Galaxy -- and that IMF is a power-law with Salpeter (1955)
slope.
Most of the stars in the other dwarf spheroidals have low values of
[α/Fe], consistent with a standard -- Salpeter -- IMF for massive stars and
an extended star formation history, as implied by their colour-magnitude
diagram (see e.g. Venn et al. 2004).
5.
Conclusions
The fossil record in low-mass stars at the present time allows the
derivation of the stellar IMF at high redshift. The low-mass luminosity
function is accessible through star counts, most robustly in a system
8
with a simple stellar population. We have found that the low-mass IMF
is invariant between globular clusters in the halo of the Milky Way and
an external galaxy, the dwarf spheroidal in Ursa Minor. The underlying
mass function is apparently the same as that for present-day star forma-
tion in the local disk of the Milky Way. The low-mass IMF is remarkably
invariant, over a broad range of metallicities, age, star-formation rate,
baryonic density, dark matter content -- indeed most of the parameters
that a priori one might have expected to be important in determining
the masses of stars. The high-mass IMF is also apparently independent
of these parameters. This invariance is particularly surprising if the
Jeans mass plays an important role.
Acknowledgments
I thank my colleagues and collaborators Sofia Feltzing, Jay Gallagher,
Gerry Gilmore, Leslie Hebb, Mark Houdashelt and Tammy Smecker-
Hane for their contributions to the results described here. I would also
like to thank the tireless organizers of this stimulating meeting for invit-
ing me.
References
Baraffe, I., Chabrier, G., Allard, F. & Hauschildt, P. 1997, A&A, 327, 1054
Bullock, J., Kravtsov, A. & Weinberg, D. 2000, ApJ, 539, 517
Bellazzini, M., Ferraro, F., Origlia, L., Pancino, E. et al. 2002, AJ, 124, 3222
Carrera, R., Aparicio, A., Martinez-Delgado, D. & Alonso-Garcia, J. 2002, AJ, 123,
3199
Cayrel, R. et al. 2004, A&A, 416, 1117
Cole, S., Lacey, C., Baugh, C. & Frenk, C. 2000, MNRAS, 319, 168
Feltzing, S., Wyse, R.F.G. & Gilmore, G. 1999, ApJL, 516, 17
Gallagher, J.S. & Wyse, R.F.G. 1994, PASP, 106, 1225
Gibson, B. 1998, ApJ, 501, 675
Hebb, L., Wyse, R.F.G. & Gilmore, G. 2004, AJ, December issue (astro-ph/0409289)
Hernandez, X., Gilmore, G. & Valls-Gabaud, D. 2000, MNRAS, 317, 831
Kauffmann, G., White, S.D.M. & Guiderdoni, B. 1993, MNRAS, 264, 201
Kleyna, J., Geller, M., Kenyon, S., Kurtz, M. & Thorstensen, J. 1998, AJ, 115, 2359
Palma, C., Majewski, S., et al. 2003, AJ, 125, 1352
Piotto, G., Cool, A. & King, I. 1997, AJ, 113, 1345
Salpeter, E.E. 1955, ApJ, 121, 161
Shetrone, M., Cot´e, P. & Sargent, W. 2001, ApJ, 548, 592
Venn, K. et al. 2004, AJ, 128, 1177
Wilkinson, M., Kleyna, J., Evans, W., Gilmore, G., Irwin, M. & Grebel, E. 2004,
ApJL, 611, 21
Wyse, R.F.G. 1998, in The Stellar IMF, ASP Conf series 142, eds Gilmore & Howell,
p89
Wyse, R.F.G. et al. 2002, New Ast, 7, 395
|
astro-ph/0404287 | 1 | 0404 | 2004-04-14T20:15:36 | The Low Quiescent X-Ray Luminosity of the Neutron Star Transient XTE J2123-058 | [
"astro-ph"
] | We report on the first X-ray observations of the neutron star soft X-ray transient (SXT) XTE J2123-058 in quiescence, made by Chandra and BeppoSAX, as well as contemporaneous optical observations. In 2002, the Chandra spectrum of XTE J2123-058 is consistent with a power-law model, or the combination of a blackbody plus a power-law, but it is not well-described by a pure blackbody. Using the interstellar column density, the power-law fit gives photon index of 3.1 (+0.7,-0.6) and indicates a 0.3-8 keV unabsorbed luminosity of 9(+4,-3)E31 (d/8.5 kpc)^2 ergs/s (90% confidence errors). Fits with models consisting of thermal plus power-law components indicate that the upper limit on the temperature of a 1.4 solar mass, 10 km radius neutron star with a hydrogen atmosphere is kT_eff < 66 eV, and the upper limit on the bolometric luminosity is L_infinity < 1.4E32 ergs/s, assuming d = 8.5 kpc. Of the neutron star SXTs that exhibit short (< 1 year) outbursts, including Aql X-1, 4U 1608-522, Cen X-4, and SAX J1810.8-2609, the lowest temperatures and luminosities are found for XTE J2123-058 and SAX J1810.8-2609. From the BeppoSAX observation of XTE J2123-058 in 2000, we obtained an upper limit on the 1-10 keV unabsorbed luminosity of 9E32 ergs/s. Although this upper limit allows that the X-ray luminosity may have decreased between 2000 and 2002, that possibility is not supported by our contemporaneous R-band observations, which indicate that the optical flux increased significantly. Motivated by the theory of deep crustal heating by Brown and co-workers, we characterize the outburst histories of the five SXTs. The low quiescent luminosity for XTE J2123-058 is consistent with the theory of deep crustal heating without requiring enhanced neutron star cooling if the outburst recurrence time is >~ 70 years. | astro-ph | astro-ph |
DRAFT VERSION NOVEMBER 18, 2018
Preprint typeset using LATEX style emulateapj v. 7/15/03
THE LOW QUIESCENT X-RAY LUMINOSITY OF THE NEUTRON STAR TRANSIENT XTE J2123 -- 058
JOHN A. TOMSICK1, DAWN M. GELINO1, JULES P. HALPERN2, PHILIP KAARET3
Draft version November 18, 2018
ABSTRACT
We report on the first X-ray observations of the neutron star soft X-ray transient (SXT) XTE J2123 -- 058 in
quiescence, made by the Chandra X-ray Observatory and BeppoSAX, as well as contemporaneous optical
observations. In 2002, the Chandra spectrum of XTE J2123 -- 058 is consistent with a power-law model, or
the combination of a blackbody plus a power-law, but it is not well-described by a pure blackbody. Using the
interstellar value of NH, the power-law fit gives Γ = 3.1+0.7
- 0.6 and indicates a 0.3 -- 8 keV unabsorbed luminosity
of (9+4- 3) × 1031 (d/8.5 kpc)2 ergs s- 1 (90% confidence errors). Fits with models consisting of thermal plus
power-law components indicate that the upper limit on the temperature of a 1.4 M⊙, 10 km radius neutron star
with a hydrogen atmosphere is kTeff < 66 eV, and the upper limit on the unabsorbed, bolometric luminosity is
L∞ < 1.4 ×1032 ergs s- 1, assuming d = 8.5 kpc. Of the neutron star SXTs that exhibit short (<1 year) outbursts,
including Aql X-1, 4U 1608 -- 522, Cen X-4, and SAX J1810.8 -- 2609, the lowest temperatures and luminosities
are found for XTE J2123 -- 058 and SAX J1810.8 -- 2609. From the BeppoSAX observation of XTE J2123 -- 058
in 2000, we obtained an upper limit on the 1 -- 10 keV unabsorbed luminosity of 9 × 1032 ergs s- 1. Although
this upper limit allows that the X-ray luminosity may have decreased between 2000 and 2002, that possibility
is not supported by our contemporaneous R-band observations, which indicate that the optical flux increased
significantly. Motivated by the theory of deep crustal heating by Brown and co-workers, we characterize the
outburst histories of the five SXTs. The low quiescent luminosity for XTE J2123 -- 058 is consistent with the
theory of deep crustal heating without requiring enhanced neutron star cooling if the outburst recurrence time
is ∼> 70 years.
Subject headings: accretion, accretion disks -- stars: individual (XTE J2123 -- 058, SAX J1810.8 -- 2609) --
stars: neutron -- X-rays: stars
1. INTRODUCTION
Accreting neutron stars can be found in high-mass (HMXB)
or low-mass (LMXB) X-ray binary systems. The majority of
HMXBs have transient X-ray emission. Their outburst spec-
tra are relatively hard and X-ray pulsations from these highly
magnetized (B ∼ 1012 G) neutron stars are typically detected.
A wide variety of X-ray behaviors are seen for neutron star
LMXBs, but, in general, the lack of X-ray pulsations from
most (but not all) of these systems, and the emission of type I
X-ray bursts from some, suggest that they harbor neutron stars
with relatively low magnetic field strengths (B ∼ 108- 9 G).
During ≈ 33 years of X-ray observations, some sources (e.g.,
Sco X-1, Cyg X-2) have maintained luminosities approaching
the Eddington limit of ≈ 1038 ergs s- 1, while others are able
to maintain persistent luminosities several orders of magni-
tude lower (Wilson et al. 2003). In addition, there is a class of
transient neutron star LMXBs for which the luminosity varies
from a substantial fraction of Eddington to quiescent levels
typically near 1032- 33 ergs s- 1. In outburst, these systems have
relatively soft spectra compared to the HMXBs, and are com-
monly grouped with black hole transients as soft X-ray tran-
sients (SXTs).
In quiescence, most neutron star SXTs exhibit X-ray en-
ergy spectra with a component that is typically fitted well
by a blackbody, suggesting that the origin of this compo-
1 Center for Astrophysics and Space Sciences, Code 0424, University
jtom-
of California at San Diego, La Jolla, CA, 92093, USA (e-mail:
[email protected])
2 Columbia Astrophysics Laboratory, Columbia University, 550 West
120th Street, New York, NY 10027
3 Harvard-Smithsonian Center for Astrophysics, 60 Garden Street, Cam-
bridge, MA, 02138, USA
nent is thermal emission from the surface of a cooling neu-
tron star. Although a pure blackbody often provides a good
fit to the spectrum, unphysical neutron star radii near 1 km
are inferred unless an atmosphere is modeled (Rutledge et al.
1999).
In addition to the thermal component, the energy
spectra often contain a second component that has a power-
law shape. The brightest and best studied systems in this
class, Cen X-4 and Aql X-1, usually display both compo-
nents (Rutledge et al. 2001, 2002a; Campana & Stella 2003;
Campana et al. 2004). However, other systems may be dom-
inated by the thermal component, such as MXB 1659 -- 29
(Wijnands et al. 2003b) and sources X-5 and X-7 in the glob-
ular cluster 47 Tucanae (Heinke et al. 2003), or by the power-
law component, such as SAX J1808.4 -- 3658 (Campana et al.
2002) and EXO 1745 -- 248 (Wijnands et al. 2003a). Although
theories for the thermal component, such as the deep crustal
heating model of Brown, Bildsten & Rutledge (1998), are rel-
atively well-developed and are being tested with observa-
tions, the origin of the power-law component is not under-
stood beyond suggestions that it may be related to accretion
onto the neutron star magnetosphere (Campana et al. 1998)
or a putative pulsar wind colliding with infalling matter from
the companion star (Tavani 1991).
In addition to our lack
of understanding of the power-law component, questions re-
main about the mass accretion rate in quiescence, the ori-
gin of rapid (100 -- 10,000 s) variability (Rutledge et al. 2002a;
Campana et al. 2004), and the origin of variability in the ther-
mal component on longer time scales (Rutledge et al. 2002a).
Another important question is if quiescent observational prop-
erties correlate with other known differences between neutron
star SXTs, such as whether they are millisecond X-ray pulsars
(during outbursts) or not, whether the systems are in the field
2
J.A. Tomsick et al.
or in globular clusters, and whether their X-ray outbursts are
long (years to decades) or short (weeks to months).
indicating that
Here, we report on X-ray and optical observations of the
field neutron star SXT XTE J2123 -- 058 taken during qui-
escence. XTE J2123 -- 058 had its only detected X-ray out-
burst in 1998 June-August (Levine, Swank & Smith 1998;
Tomsick et al. 1999), and we focus on observations made
with Chandra, BeppoSAX, and optical telescopes 2 -- 4 years
after the outburst. During the outburst,
the Rossi X-ray
Timing Explorer (RXTE) detected type I X-ray bursts and a
pair of kHz quasi-periodic oscillations (Homan et al. 1999;
Tomsick et al. 1999),
the system contains
a rapidly rotating neutron star. However, coherent X-ray
pulsations were not found. The 6 hr binary orbital pe-
riod and the fact that the binary inclination of the sys-
tem is relatively high were established from optical mod-
ulation and the presence of partial eclipses in the opti-
cal light curve (Tomsick et al. 1999; Soria, Wu & Galloway
1999; Zurita et al. 2000; Shahbaz et al. 2003). The high
Galactic latitude (b = - 36◦) and low extinction have allowed
for detailed optical studies of XTE J2123 -- 058 in quiescence
even though the source is rather faint at its relatively large
distance (8.5 ± 2.5 kpc). The optical observations show that
XTE J2123 -- 058 consists of a K7 V star on or close to the
main sequence and a neutron star for which mass determi-
nations of 1.5 ± 0.3 M⊙ (Tomsick et al. 2001; Casares et al.
2002; Tomsick et al. 2002) and 1.04 -- 1.56 M⊙ (Shahbaz et al.
2003) have been obtained. The focus of this paper is the first
X-ray study of XTE J2123 -- 058 in quiescence.
2. OBSERVATIONS AND ANALYSIS
We observed XTE J2123 -- 058 with Chandra on UT 2002
November 13 (ObsID 2709), using the Advanced CCD Imag-
ing Spectrometer (ACIS) with the target placed on one of
the back-illuminated ACIS chips (ACIS-S3). For our anal-
ysis, we used the "level 2" event list produced by the stan-
dard data processing with ASCDS version 6.9.2 using Cal-
ibration Data Base (CALDB) version 2.17. Light curves
using counts from the full field-of-view do not show any
background flares, allowing us to use the data from the full
17,706 s integration. Using the Chandra Interactive Analysis
of Observations (CIAO) version 3.0 software routine wavde-
tect (Freeman et al. 2002), we searched for sources on the S3
chip in the 0.3 -- 8 keV energy band. We detected 22 sources
with counts between 5 and 108 per source, using a detec-
tion threshold of 10- 6, including a 24 count source at R.A.
= 21h23m14s.54, decl. = - 05◦47′53′′.2 (equinox J2000, un-
certainty 0′′.6). This position is consistent with the target's
optical position (Tomsick et al. 1999), and we conclude that
this source is the quiescent X-ray counterpart of XTE J2123 --
058.
We also report on a BeppoSAX observation of the XTE
J2123 -- 058 field made on UT 2000 May 12 -- 13. We produced
1 -- 10 keV images using the data from the two Medium En-
ergy Concentrator/Spectrometer (MECS) units that were op-
erational during the observation (units 2 and 3). We also pro-
duced a 0.1 -- 10 keV image using data from the Low Energy
Concentrator/Spectrometer (LECS). We obtained a MECS
exposure time of 46,340 s and a LECS exposure time of
17,080 s. To search for sources, we convolved each of the
three images with a two-dimensional Gaussian with a width
(σ) of 5 pixels (40′′) in both directions. Only one source
was clearly detected in both of the MECS images at R.A. =
21h22m50s.6, decl. = - 05◦45′09′′ (equinox J2000, uncertainty
∼ 1′), and this source, which we call SAX J2122.8 -- 0575, is
also present in the LECS image. It is clear that SAX J2122.8 --
0575 is not XTE J2123 -- 058 as they are separated by 6′.6. We
conclude that XTE J2123 -- 058 was not detected during the
BeppoSAX observation, and we derive an upper limit on its
X-ray flux below.
Finally, we obtained R-band images on three occasions
close to the times of the X-ray observations. As shown in
Table 1, we observed XTE J2123 -- 058 using the 2.4 m Hiltner
telescope of the MDM Observatory on 2000 July 24, about
2 months after the BeppoSAX observation. We also observed
XTE J2123 -- 058 with the Shane 3 m telescope of Lick Ob-
servatory about 2 months before the Chandra observation and
again at MDM about 2 weeks after the Chandra observation.
For both MDM runs, we used the same SITe 2048 × 2048
pixel, thinned, back-illuminated CCD with a spatial scale of
0′′.275 per 24 µm pixel, and an R filter that is very close to
Harris R. 4 At Lick, we used the Prime Focus Camera, with
a SITe 2048 × 2048 pixel, thinned CCD with a spatial scale
of 0′′.296 per 24 µm pixel, and a Kron-Cousins R filter. In
all, we obtained thirteen 600 -- 700 s exposures (see Table 1),
and we reduced the images using standard IRAF 5 routines.
For XTE J2123 -- 058, we carried out the photometry with the
IRAF package phot and used two calibrated reference stars
with R magnitudes of 19.47 and 19.51.
In 2002, the con-
ditions at Lick were photometric, and we observed Landolt
(1992) standards to obtain the calibration. We note that this
calibration is about 0.1 magnitudes brighter than the calibra-
tion previously obtained using MDM observations from 1998
reported in Tomsick et al. (1999).
3. ENERGY SPECTRUM AND SOURCE LUMINOSITY
We used the CIAO software routine psextract to produce
the ACIS energy spectrum for XTE J2123 -- 058 and to create
the appropriate instrument response matrix for the spectrum.
The software used CALDB 2.23 to create the response matrix,
and we included a correction for the time-dependence of the
ACIS response. The source spectrum included counts from a
circular region with a 5 pixel (about 2′′.5) radius, and we es-
timated the background level using counts from a source-free
annulus around the target position. The source spectrum con-
sists of 24 counts, and we estimate a background level of 0.7
counts in the extraction region. We produced a "light curve"
of the source in six time bins of ≈ 3000 s. Each bin contains
between 2 and 6 counts, which is consistent with a constant
flux; however, the low count rate does not allow us to place
tight constraints on the possible amplitude of variability. The
small number of counts also indicates that χ2 statistics are
not appropriate for spectral analysis, as the assumption of a
Gaussian probability distribution in each spectral bin is not
met. Thus, we carried out our spectral analysis by minimiz-
ing the Cash statistic (Cash 1979), which is appropriate in
cases where the assumption of a Poisson probability distribu-
tion in each spectral bin is valid. We fitted the spectrum using
XSPEC 11.2.
The spectra of other neutron star SXTs are typically well-
described by a blackbody, a power-law, or both components
with interstellar absorption. We began by fitting the spec-
4 See http://www.astro.lsa.umich.edu/obs/mdm/technical/filters for the ex-
act transmission curve.
5 IRAF (Image Reduction and Analysis Facility) is distributed by the Na-
tional Optical Astronomy Observatories, which are operated by the Asso-
ciation of Universities for Research in Astronomy, Inc., under cooperative
agreement with the National Science Foundation.
XTE J2123 -- 058 in Quiescence
a
b
10-2
10-3
10-4
10-5
10-6
10-7
1
-
V
e
k
1
-
s
s
t
n
u
o
C
)
1
-
s
2
-
m
c
V
e
k
(
E
F
E
1
Energy (keV)
10
FIG. 1. -- Chandra/ACIS energy spectrum of XTE J2123 -- 058. Top: folded
through the detector response. Bottom: unfolded. In each panel, the solid
line is a power-law fit to the data (using Cash statistics as described in the
text) with NH fixed at the interstellar value. The power-law photon index is
- 0.6, and the unabsorbed 0.3 -- 8 keV luminosity is (9+4- 3) × 1031 ergs s- 1. A
3.1+0.7
blackbody model (dashed line), by itself, does not provide a good description
of the spectrum.
trum for XTE J2123 -- 058 with an absorbed power-law, and
the results are given in Table 2. Once we found the best-fitted
parameters by minimizing the Cash statistic, we determined
the quality of the fit and the 90% confidence errors on the
parameters by producing and fitting 10,000 simulated spec-
tra. We used the best-fitted parameters from the fits to the
actual data as input to the simulations. Using simulations to
determine the parameter errors is necessary only because of
the small number of counts in the spectrum. We tested our
method by producing spectra with 2 -- 3 times as many counts,
and we found that the errors produced by calculating changes
in the Cash statistic (the standard technique) match the val-
ues we obtain using simulations. In addition to the power-law
model, we used simulations to determine the errors on param-
eter for all other fits presented in this work. For the power-law
model, we obtain NH = (7.0+40.0
- 7.0 ) × 1020 cm- 2 for the column
density and Γ = 3.1+2.8
- 0.8 for the photon index. The fit using
a blackbody model is significantly worse than the power-law
model as indicated by the fact that for 97% of the simulated
spectra we obtained a better Cash statistic than the one ob-
tained when fitting the actual spectrum (compared to 69% for
the power-law model).
We refitted the spectrum with the same spectral mod-
els but with NH fixed to the interstellar column density.
The interstellar NH comes from the AV measurement of
Hynes et al. (2001), which gives NH = (6.6 ±2.7) ×1020 cm- 2,
and the total Galactic H I value of 5.7 × 1020 cm- 2 from
Dickey & Lockman (1990). We adopt a value of 6 × 1020
cm- 2. For the power-law alone, we obtain Γ = 3.1+0.7
- 0.6 and an
- 0.4) × 10- 14 ergs cm- 2 s- 1,
unabsorbed 0.3 -- 8 keV flux of (1.1+0.5
3
which corresponds to a luminosity of (9+4- 3) × 1031 ergs s- 1
at a distance of 8.5 kpc. The best fitted blackbody temper-
ature is 240+60
- 50 eV, but the quality of the blackbody fit is even
worse (98.7%) with NH fixed. Figure 1 shows the spectrum
(rebinned for clarity, although the Cash fits are performed
without rebinning) along with the best fitted power-law and
blackbody models with NH fixed. It is clear that the curvature
of the blackbody model is too large, underpredicting the ob-
served spectrum below 0.7 keV and above 2 keV. This, along
with the measurements of the fit quality from the simulations,
indicates that the spectrum is well-described by a power-law
model or the combination of a blackbody and a power-law
model, but it is not well-described by a blackbody alone.
Although a power-law alone provides an adequate descrip-
tion of the ACIS spectrum, we performed additional fits to
obtain limits on the temperature and luminosity of the puta-
tive thermal component that is expected to be emitted from
the surface of the neutron star. We fitted the spectrum with a
model consisting of a power-law component along with ther-
mal emission from an atmosphere composed of hydrogen.
The latter component was modeled using the "Neutron Star
Atmosphere" (NSA) model of Zavlin, Pavlov & Shibanov
(1996). We assumed a neutron star mass of 1.4 M⊙, which
is consistent with the measured mass, and a radius of 10 km.
With these parameters fixed, the remaining free parameters in
the NSA model are the temperature (kTeff) and the distance.
For XTE J2123 -- 058, there are several arguments that lead to
distance estimates in the range 5 -- 15 kpc (Tomsick et al. 1999;
Homan et al. 1999; Zurita et al. 2000), but the most reliable
estimates come from optical observations of the source in qui-
escence. Previous estimates include 8 ± 3 kpc (Zurita et al.
2000), 8.5 ± 2.5 kpc (Tomsick et al. 2001), and 9.6 ± 1.3 kpc
(Casares et al. 2002). Here, we adopt a range of 6 -- 11 kpc. We
performed the NSA plus power-law fits with the distance fixed
to three values spanning this range (6, 8.5, and 11 kpc). As
shown in Table 3, the 90% confidence upper limits on kTeff are
57, 66, and 73 eV, respectively, for these three distances. Us-
eff(1 - 2GM/Rc2), where R = 10 km and M
ing L∞ = 4πR2σT 4
= 1.4 M⊙, upper limits on the unabsorbed luminosity from the
NSA component as seen by a distant observer are 8.0 × 1031,
1.4 × 1032 and 2.1 × 1032 ergs s- 1 for distances of 6, 8.5, and
11 kpc, respectively. While these values are bolometric lumi-
nosities, it should be noted that we cannot rule out the pos-
sibility that there is thermal emission at energies below the
Chandra bandpass.
Although XTE J2123 -- 058 was not detected by BeppoSAX
in 2000 May, we can calculate upper limits on its luminos-
ity if we assume that the energy spectrum was similar to that
measured by Chandra. To calculate the upper limits, we as-
sume the simplest model of an absorbed power-law with a
photon index of 3.1 and the interstellar column density. For
the MECS (units 2 and 3 combined), we measure a 1 -- 10 keV
count rate of 1.684 ks- 1 in a circle of radius 2′ centered on the
XTE J2123 -- 058 position. This is not significantly higher than
the expected background rate (using a blank-sky pointing) of
1.675 ks- 1. These values imply a 3-σ upper limit on the count
rate from XTE J2123 -- 058 of 0.676 ks- 1, which corresponds
to an unabsorbed flux < 1.02 × 10- 13 ergs cm- 2 s- 1, and a 1 --
10 keV luminosity < 9 × 1032 ergs s- 1 for d = 8.5 kpc. For the
LECS, the 0.1 -- 1 keV count rate in a 2′ radius circle centered
on XTE J2123 -- 058 is 1.230 ks- 1, which is actually somewhat
lower than the expected background rate of 1.466 ks- 1. Using
the expected background rate, the 3-σ upper limit on the un-
absorbed 0.1 -- 1 keV flux is 1.75 × 10- 12 ergs cm- 2 s- 1, corre-
4
sponding to a luminosity < 1.5 × 1034 ergs s- 1 for d = 8.5 kpc.
J.A. Tomsick et al.
4. OPTICAL RESULTS
Figure 2 shows the R magnitudes for XTE J2123 -- 058 from
three exposures taken in 2000 July at MDM, two exposures
taken in 2002 September at Lick, and eight exposures taken
in 2002 November at MDM. Previously, Shahbaz et al. (2003)
reported on extensive R-band photometry of XTE J2123 -- 058
taken between 1999 June and 2000 August. The R-band
light curve was relatively stable at that time, showing or-
bital ellipsoidal modulations with a peak-to-peak amplitude
of about 0.25 magnitudes. The dashed lines in Figure 2 in-
dicate the range of the modulation in 1999 -- 2000, which was
21.70 < R < 21.95. While our 2000 May measurements of
21.63 ± 0.07, 21.75 ± 0.08, and 21.57 ± 0.07 are consistent
with the brighter end of this range, our later observations indi-
cate that XTE J2123 -- 058 brightened between 2000 and 2002.
The range of R between 21.41 ± 0.07 and 21.01 ± 0.03 over
25% of the 6 hr binary orbital period during the 2002 Novem-
ber run indicates that the source continues to vary, but with
a peak level that is at least 0.6 -- 0.7 magnitudes brighter than
the peak level of R = 21.7 measured by Shahbaz et al. (2003).
As our observations do not cover the full binary orbit, it is
not clear whether this change is due to the addition of a con-
stant component to the light curve or if the shape of the mod-
ulations has changed. Although it would be useful to know
the orbital phases corresponding to the 2002 data, the binary
ephemeris is too uncertain to extrapolate to 2002 as discussed
in Tomsick et al. (2002). In the future, it may be worthwhile
to obtain an X-ray observation while measuring the optical or
IR light curve for a full 6 hr binary orbit.
As we are reporting the first sensitive X-ray observations
of XTE J2123 -- 058 in quiescence, comparison can be made
with a prior prediction for the quiescent X-ray luminosity of
this source that was made by modeling quiescent R-band light
curves in 1999 and 2000. Shahbaz et al. (2003) predicted an
X-ray luminosity of ∼1033 ergs s- 1, which is only slightly
higher than our upper limit from the BeppoSAX observation
made in 2000, but an order of magnitude higher than we ob-
served with Chandra in 2002. While the X-ray observations
would allow for the possibility that the quiescent X-ray lu-
minosity decreased significantly between 2000 and 2002, our
optical observations made in 2000 and 2002 indicate that the
R-band flux actually increased over this time. If the 2000 R-
band light curve showed a significant contribution from X-
ray heating, it is difficult to see how a drop in X-ray flux
could lead to an increase in the optical, thus, it is possible that
Shahbaz et al. (2003) over-estimated the contribution from X-
ray heating.
5. INTERPRETATION
X-ray observations of neutron star SXTs in quiescence
provide tests of theoretical models for the thermal com-
ponent. According to the theory of deep crustal heating
by Brown, Bildsten & Rutledge (1998), the temperature of
the neutron star core is maintained by nuclear reactions in
the deep crust that occur when the mass accretion rate is
high during outburst. As the thermal time scale for the
core is ∼10,000 years (Colpi et al. 2001), the level of qui-
escent thermal emission is set by the average mass accre-
tion rate over this time span according to Lq = 9 × 1032
< M>- 11 ergs s- 1 (see Equation 1 of Rutledge et al. 2002b),
assuming 1.45 MeV of heat deposited in the crust per ac-
creted nucleon (Haensel & Zdunik 1990). Here < M>- 11 is
a
749.880
749.865
749.870
749.875
HJD-2451000 (days)
b
c
1529.78
1529.74
1529.76
HJD-2451000 (days)
e
d
u
t
i
n
g
a
m
R
-
e
d
u
t
i
n
g
a
m
R
-
21.0
21.2
21.4
21.6
21.8
22.0
749.860
21.0
21.2
21.4
21.6
21.8
22.0
1529.72
e
d
u
t
i
n
g
a
m
R
-
21.0
21.2
21.4
21.6
21.8
22.0
1603.56
1603.58
1603.60
1603.62
HJD-2451000 (days)
1603.64
(b) 2002 September at Lick Observatory.
FIG. 2. -- R-band measurements of XTE J2123 -- 058. (a) 2000 July at MDM
Observatory.
(c) 2002 Novem-
ber at MDM Observatory. The dashed lines delimit the levels measured by
Shahbaz et al. (2003) during 1999 -- 2000. An increase in the R-band flux be-
tween 2000 and 2002 is apparent.
the mass accretion rate averaged over the thermal time scale
of the core in units of 10- 11 M⊙ yr- 1, and Lq is the quiescent
bolometric luminosity. While the time-averaged mass accre-
tion rate is only a predictor of Lq if the neutron star core has
reached thermal equilibrium, the lifetimes of LMXB systems
are much longer than the thermal time scale of the core, and it
is expected that thermal equilibrium has been established for
all or nearly all of the LMXBs. While the long thermal time
scale for the core precludes appreciable changes in the core
temperature during a single SXT outburst, the neutron star
crust can be heated significantly for the systems with longer
outbursts (years to decades) so that the quiescent emission is
determined by the evolution of the physical conditions in the
crust (Rutledge et al. 2002b). However, for the systems with
shorter outbursts (weeks to months), the properties of the qui-
escent thermal emission are primarily set by the conditions in
the neutron star core (Brown, Bildsten & Rutledge 1998).
For a system like XTE J2123 -- 058 that has undergone one
≈ 40 day outburst, the deep crustal heating theory implies
that quiescent thermal emission provides information about
the neutron star core. Here, we compare the quiescent prop-
erties of XTE J2123 -- 058 to similar systems. The compar-
ison group includes field neutron star SXTs with outbursts
lasting less than 1 year for which quiescent X-ray observa-
tions have been reported. We only consider LMXB systems
XTE J2123 -- 058 in Quiescence
5
from which X-ray bursts have been detected, proving that the
accreting object is a neutron star. We restrict the compari-
son group to field systems because the outburst histories for
most transients in globular clusters are uncertain due to source
confusion in instruments with angular resolution worse than
Chandra. Thus, our comparison group consists of the neu-
tron star SXTs Aql X-1, 4U 1608 -- 522, Cen X-4, and SAX
J1810.8 -- 2609. We also compare the quiescent properties of
XTE J2123 -- 058 to those of the millisecond X-ray pulsar SAX
J1808.8 -- 3658 as the quiescent X-ray luminosities of the two
sources are comparable.
Table 4 compares the parameters of the thermal compo-
nent for the five systems, including kTeff, R, L∞, and the
gravitational redshift parameter, g = p1 - 2GM/Rc2, assum-
ing that all the neutron star masses are 1.4 M⊙. For the
spectral fits to XTE J2123 -- 058, we set R = 10 km, and
Jonker, Wijnands & van der Klis (2004) make the same as-
sumption for SAX J1810.8 -- 2609. However, for the other
sources, we use the best fitted radii, and the parameters are
taken from Rutledge et al. (1999, 2001, 2002a). The thermal
component is detected for Cen X-4, Aql X-1, and 4U 1608 --
522, and a power-law component is also required for Cen X-
4 and for some of the observations of Aql X-1. For XTE
J2123 -- 058 and SAX J1810.8 -- 2609, the thermal component
is not statistically required, and we report upper limits on the
thermal component parameters for these two systems. Also,
for all the sources we assume the best current distance mea-
surement. The temperature upper limits for XTE J2123 -- 058
(66 eV) and SAX J1810.8 -- 2609 (72 eV) are about a factor
of 2.5 lower than the 4U 1608 -- 522 measurement of 170 eV,
and the luminosity upper limits for XTE J2123 -- 058 and SAX
J1810.8 -- 2609 are about a factor of 50 lower than the maxi-
mum quiescent luminosity measured for Aql X-1. The tem-
perature and luminosity of Cen X-4 are intermediate between
the hottest and most luminous sources (Aql X-1 and 4U 1608 --
522) and the coolest and least luminous sources (XTE J2123 --
058 and SAX J1810.8 -- 2609).
Here, we characterize the outburst histories of these
they are related to the
five sources to determine if
expected in the
as
quiescent
thermal properties
Brown, Bildsten & Rutledge (1998) theory.
From a re-
view of SXT light curves covering the time period between
1969 and 1996 (Chen, Shrader & Livio 1997) and the RXTE
All-Sky Monitor (ASM) covering 1996 to 2003, 4U 1608 --
522 had 16 outbursts with a mean peak X-ray luminosity,
¯Lpeak, of 2.5 × 1037 (d/3.6 kpc)2 ergs s- 1. Aql X-1 had 21
outbursts with ¯Lpeak = 3.6 × 1037 (d/5 kpc)2 ergs s- 1. Cen X-4
had two outbursts (in 1969 and 1979) with ¯Lpeak = 5.6 × 1037
(d/1.2 kpc)2 ergs s- 1, and XTE J2123 -- 058 had one outburst
(in 1998) with ¯Lpeak = 1.7 × 1037 (d/8.5 kpc)2 ergs s- 1. Like
XTE J2123 -- 058, SAX J1810.8 -- 2609 had only one outburst,
also in 1998.
is
SAX J1810.8 -- 2609 was discovered by Natalucci et al.
(2000); and the only reported X-ray detections of this source
in outburst were from BeppoSAX and ROSAT in 1998 March
(Natalucci et al. 2000; Greiner et al. 1999). However, the
ASM light curve shown in Figure 3c indicates that SAX
J1810.8 -- 2609 was in outburst by 1998 January 4 (MJD
50,817) and probably as early as 1997 December 18 (MJD
50,800). The dashed lines in Figure 3c mark the times
of the BeppoSAX discovery at MJD 50,882 -- 50,885 and the
ROSAT observation at MJD 50,896, while it is clear that
the source was considerably brighter prior to this. From
)
V
e
k
2
1
-
5
.
1
(
1
-
s
M
S
A
)
V
e
k
2
1
-
5
.
1
(
1
-
s
M
S
A
)
V
e
k
.
2
1
-
5
1
(
1
-
s
M
S
A
)
V
e
k
.
2
1
-
5
1
(
1
-
s
M
S
A
a. Aql X-1
80
60
40
20
0
2600
2650
2700
2750
MJD-50000 (days)
2800
2850
b. 4U 1608-522
80
60
40
20
0
2600
900
950
2400
2450
2500
2550
MJD-50000 (days)
10
c. SAX J1810.8-2609
5
0
-5
700
750
800
850
MJD-50000 (days)
10
d. XTE J2123-058
5
0
-5
900
950
1000
1050
MJD-50000 (days)
1100
FIG. 3. -- All-Sky Monitor (ASM) light curves for the four neutron star
SXTs in our comparison group that had outbursts during the RXTE lifetime.
Each point represents the average ASM count rate over 1 day. One Crab
flux equals 75 ASM counts s- 1. For SAX J1810.8 -- 2609 (c), the vertical
dashed lines mark the times of previously reported BeppoSAX and ROSAT
observations. Aql X-1 and 4U 1608 -- 522 have had many outbursts during the
lifetime of RXTE, and the light curves in (a) and (b) are representative.
the 1.5 -- 12 keV ASM light curve, ¯Lpeak = 1.0 × 1037 (d/4.9
kpc)2 ergs s- 1. Although it has been suggested that SAX
J1810.8 -- 2609 had an unusually low outburst
luminosity
(Jonker, Wijnands & van der Klis 2004), the value we derive
from the ASM measurements is comparable to the other neu-
tron star SXTs. Figure 3 also shows the ASM light curve
for the XTE J2123 -- 058 outburst as well as sample outburst
light curves for Aql X-1 and 4U 1608 -- 522 (Cen X-4 has not
6
J.A. Tomsick et al.
had an outburst during the RXTE lifetime). These light curves
demonstrate that individual outbursts from different sources
are similar in duration (typically 40 -- 80 days) and overall
shape, although it should be noted that exceptional outbursts
do occur (Chen, Shrader & Livio 1997).
For each of the neutron star SXTs, the time-averaged mass
accretion rate over the past 33 years can be expressed as
< M >=
¯LpeakNtoutburst f
ǫc2(33yr)
= s¯LpeakN
,
(1)
where N is the number of outbursts, toutburst is the typical du-
ration of an outburst from the source,
f is a factor with a
value less than 1.0 that accounts for the shape of the outburst
light curve, ǫ is the fraction of the rest mass energy released
from accreted matter, and s = toutburst f /ǫc2(33yr). As the du-
rations and light curve shapes are similar for the sources in
our comparison group, s is approximately the same for these
sources, and the time-averaged mass accretion rate over the
past 33 years is proportional to ¯LpeakN. We estimate s us-
ing toutburst = 60 days and ǫ = 0.2, which is the value of ǫ
that is typically assumed for accretion onto a neutron star
(Rutledge et al. 2002b). Although the precise value of f de-
pends on the exact shape of the outburst light curve, it can be
approximated as the mean outburst flux divided by the peak
flux. For the XTE J2123 -- 058 ASM light curve, the mean flux
over a 60 day period that includes the 1998 outburst is 41%
of the peak level, and we use f = 0.4, giving s = 1.1 × 10- 23 s2
cm- 2. Using this value of s for all five sources and the values
for ¯Lpeak and the number of outbursts (N) given above, we cal-
culate estimates of the time-averaged mass accretion rate over
the past 33 years. The values for < M> are given in Table 4,
and they show that this quantity is a relatively good predictor
of L∞. Aql X-1 and 4U 1608 -- 522 have the the highest values
of both < M> and L∞, XTE J2123 -- 058 and SAX J1810.8 --
2609 have the lowest, and Cen X-4 is intermediate. The ratio
of the value of < M> for Aql X-1 to that for XTE J2123 --
058 is 45, which is consistent with their ratio of L∞, which
is > 38. Similarly, the Aql X-1 to SAX J1810.8 -- 2609 ratio
of < M> is 76, and this is consistent with their ratio of L∞,
which is > 27. While these measurements are in line with the
Brown, Bildsten & Rutledge (1998) theory, a caveat is that we
cannot be certain that X-ray flux history over the past 33 years
reflects the behavior over the last 10,000 years.
We can use the Brown, Bildsten & Rutledge (1998) theory
to predict the recurrence time for outbursts of XTE J2123 --
058, assuming that the outbursts are similar. Using L =
ǫ Mc2 and the expression relating Lq to the time-averaged
mass accretion rate, Lq = 9 × 1032 < M>- 11 ergs s- 1, one ob-
tains (see also Wijnands et al. 2001) Lq = [toutburst/(toutburst +
tq)](<Loutburst>/130). Here, tq is the average time the source
spends in quiescence between outbursts, <Loutburst> is the
average luminosity during an outburst, and toutburst is de-
fined above. For XTE J2123 -- 058, we used the RXTE/ASM
light curve shown in Figure 3d to determine that the mean
1.5 -- 12 keV count rate during the 40 days (toutburst) between
MJD 50,990 and MJD 51,030 was 3.2 s- 1, corresponding to
<Loutburst>= 1.1 × 1037 ergs s- 1 for d = 8.5 kpc. From the lu-
minosity upper limit of Lq < 1.4 × 1032 ergs s- 1, we obtain
a lower limit on the recurrence time, toutburst + tq, of 67 years.
The known outburst history of XTE J2123 -- 058, one outburst
in 33 years, assuming that no outbursts were missed, is con-
sistent with these long predicted recurrence times. However,
if the recurrence time is in fact shorter than 67 years, then a
mechanism of enhanced cooling of the core would be neces-
sary. The most often mentioned mechanism is neutrino cool-
ing of the core due to the direct Urca process, and this mech-
anism requires a neutron star with a mass higher than 1.7 --
1.8 M⊙(Colpi et al. 2001). The mass of the neutron star in
XTE J2123 -- 058 has been measured at 1.5 ± 0.3 M⊙ or 1.04 --
1.56 M⊙, so that the mass would have to be at the upper end
of the error range for the direct Urca process to operate.
While we have been focusing primarily on the implica-
tions of the upper limit on the temperature and luminosity
of the thermal component, it is notable that the power-law
fit to the spectrum of XTE J2123 -- 058 also indicates that its
X-ray luminosity, at (9+4- 3) × 1031 ergs s- 1 in the unabsorbed
0.3 -- 8 keV band, is among the lowest for neutron star SXTs.
The quiescent luminosity of SAX J1810.8 -- 2609, ≈ 1 × 1032
ergs s- 1 (Jonker, Wijnands & van der Klis 2004),
is essen-
tially the same as for XTE J2123 -- 058, and the only other sys-
tem in this class known to have a lower quiescent X-ray lumi-
nosity, SAX J1808.4 -- 3658, is also the only millisecond X-ray
pulsar system for which there is a measurement of its quies-
cent X-ray spectrum. The XMM-Newton spectrum of SAX
J1808.4 -- 3658 is well described by a power-law with Γ = 1.5
and a 0.5 -- 10 keV unabsorbed luminosity of 5 × 1031 ergs s- 1,
and the upper limit on the contribution from a thermal com-
ponent is 10% of the total X-ray luminosity (Campana et al.
2002). If the spectrum of XTE J2123 -- 058 consists of only
a power-law component, then its Γ = 3.1+0.7
- 0.6 is quite a bit
softer than that of SAX J1808.4 -- 3658, and the difference
would have to be explained. On the other hand, if the spec-
trum for XTE J2123 -- 058 consists of a thermal component and
a power-law, then its power-law could have the same slope
as SAX J1808.4 -- 3658. Concerning the SAX J1808.4 -- 3658
thermal component, Campana et al. (2002) conclude that en-
hanced neutron star cooling is required to obtain a quiescent
luminosity as low as the XMM-Newton upper limit implies.
However, as mentioned above, such conclusions depend on
the assumption that the recent outburst behavior is indicative
of the average mass accretion rate over the past 10,000 years.
In summary, our X-ray measurements of XTE J2123 -- 058
in quiescence show that this system is one of the faintest and
least luminous of the neutron star SXTs. A comparison of the
XTE J2123 -- 058 X-ray properties in outburst and quiescence
to that of four other neutron star SXTs that exhibit short out-
bursts shows that the systems with a higher degree of outburst
activity tend to have more luminous thermal components in
quiescence, and this is in-line with predictions of the theory of
deep crustal heating. For XTE J2123 -- 058, the upper limit on
the thermal luminosity is consistent with this theory without
requiring enhanced neutron star cooling if the outburst recur-
rence time is >67 years, which is consistent with the known
outburst history of this source.
JAT would especially like to thank Keith Arnaud for his
help with statistical techniques and XSPEC. JAT acknowl-
edges useful conversations with Tariq Shahbaz and Rudy
Wijnands.
JAT and PK acknowledge partial support from
Chandra awards GO2-3051X and GO3-4043X issued by the
Chandra X-ray Observatory Center, which is operated by
the Smithsonian Astrophysical Observatory for and on behalf
of NASA under contract NAS8-39073. DMG acknowledges
support from a CASS postdoctoral fellowship.
XTE J2123 -- 058 in Quiescence
7
REFERENCES
Brown, E. F., Bildsten, L., & Rutledge, R. E., 1998, ApJ, 504, L95
Campana, S., Colpi, M., Mereghetti, S., Stella, L., & Tavani, M., 1998,
1999, ApJ, 514, 945
Rutledge, R. E., Bildsten, L., Brown, E. F., Pavlov, G. G., & Zavlin, V. E.,
Campana, S., Israel, G. L., Stella, L., Gastaldello, F., & Mereghetti, S., 2004,
2001, ApJ, 551, 921
A&A Rev., 8, 279
ApJ, 601, 474
MNRAS, 329, 29
Campana, S., & Stella, L., 2003, ApJ, 597, 474
Campana, S., et al., 2002, ApJ, 575, L15
Casares, J., Dubus, G., Shahbaz, T., Zurita, C., & Charles, P. A., 2002,
Cash, W., 1979, ApJ, 228, 939
Chen, W., Shrader, C. R., & Livio, M., 1997, ApJ, 491, 312
Colpi, M., Geppert, U., Page, D., & Possenti, A., 2001, ApJ, 548, L175
Dickey, J. M., & Lockman, F. J., 1990, ARA&A, 28, 215
Freeman, P. E., Kashyap, V., Rosner, R., & Lamb, D. Q., 2002, ApJS, 138,
185
Greiner, J., Castro-Tirado, A. J., Boller, T., Duerbeck, H. W., Covino, S.,
Israel, G. L., Linden-Vørnle, M. J. D., & Otazu-Porter, X., 1999, MNRAS,
308, L17
Haensel, P., & Zdunik, J. L., 1990, A&A, 227, 431
Heinke, C. O., Grindlay, J. E., Lloyd, D. A., & Edmonds, P. D., 2003, ApJ,
Rutledge, R. E., Bildsten, L., Brown, E. F., Pavlov, G. G., & Zavlin, V. E.,
Rutledge, R. E., Bildsten, L., Brown, E. F., Pavlov, G. G., & Zavlin, V. E.,
2002a, ApJ, 577, 346
Rutledge, R. E., Bildsten, L., Brown, E. F., Pavlov, G. G., Zavlin, V. E., &
Ushomirsky, G., 2002b, ApJ, 580, 413
Shahbaz, T., Zurita, C., Casares, J., Dubus, G., Charles, P. A., Wagner, R. M.,
& Ryan, E., 2003, ApJ, 585, 443
Soria, R., Wu, K., & Galloway, D. K., 1999, MNRAS, 309, 528
Tavani, M., 1991, ApJ, 379, L69
Tomsick, J. A., Halpern, J. P., Kemp, J., & Kaaret, P., 1999, ApJ, 521, 341
Tomsick, J. A., Heindl, W. A., Chakrabarty, D., Halpern, J. P., & Kaaret, P.,
2001, ApJ, 559, L123
570
Tomsick, J. A., Heindl, W. A., Chakrabarty, D., & Kaaret, P., 2002, ApJ, 581,
Wijnands, R., Heinke, C. O., Pooley, D., Edmonds, P. D., Lewin, W. H. G.,
Grindlay, J. E., Jonker, P. G., & Miller, J. M., 2003a, astro-ph/0310144
Wijnands, R., Homan, J., Miller, J. M., & Lewin, W. H. G., 2003b,
588, 452
1999, ApJ, 513, L119
Homan, J., Méndez, M., Wijnands, R., van der Klis, M., & van Paradijs, J.,
Wijnands, R., Miller, J. M., Markwardt, C., Lewin, W. H. G., & van der Klis,
Hynes, R. I., Charles, P. A., Haswell, C. A., Casares, J., Zurita, C., & Serra-
Wilson, C. A., Patel, S. K., Kouveliotou, C., Jonker, P. G., van der Klis, M.,
astro-ph/0310612
M., 2001, ApJ, 560, L159
Ricart, M., 2001, MNRAS, 324, 180
Jonker, P. G., Wijnands, R., & van der Klis, M., 2004, MNRAS, 349, 94
Landolt, A. U., 1992, AJ, 104, 340
Levine, A., Swank, J., & Smith, E., 1998, IAU Circular, 6955
Natalucci, L., Bazzano, A., Cocchi, M., Ubertini, P., Heise, J., Kuulkers, E.,
in 't Zand, J. J. M., & Smith, M. J. S., 2000, ApJ, 536, 891
Lewin, W. H. G., Belloni, T., & Méndez, M., 2003, ApJ, 596, 1220
Zavlin, V. E., Pavlov, G. G., & Shibanov, Y. A., 1996, A&A, 315, 141
Zurita, C., et al., 2000, MNRAS, 316, 137
8
J.A. Tomsick et al.
XTE J2123 -- 058 X-RAY AND OPTICAL OBSERVATIONS
TABLE 1
UT Date
2000 July 24
2000 May 12 -- 13
Energy Band Exposure Time (s)
0.1 - 10 keV
Observatory
BeppoSAX
MDM
Lick
Chandra
MDM
aThis is the MECS exposure time. The LECS exposure time is 17,080 s.
2002 September 11 Kron-Cousins R
2002 November 13
2002 November 25
46,340a
3 × 600
2 × 700
17,706
8 × 600
0.3 - 8 keV
Harris R
Harris R
TABLE 2
SPECTRAL FITS
Model
a
NH
Γ
b
Fpl
L/d2
8.5
c
Fit
kTbb
(eV)
--
260+50
- 110
100+130
- 60
--
(1020 cm- 2)
(10- 14 ergs cm- 2 s- 1)
3.1+2.8
- 0.8
--
2.5+2.9
- 2.3
3.1+0.7
- 0.6
--
2.5+2.1
- 2.5
7.0+40.0
- 7.0
0.0+40.0
- 0.0
8.0+38.0
- 8.0
6.0
6.0
6.0
1.2+31.9
- 0.5
--
0.70+3.67
- 0.48
1.1+0.5
- 0.4
--
0.7+0.5
- 0.5
pl
bb
bb+pl
pl
bb
bb+pl
aErrors are 90% confidence for all parameters.
b0.3 -- 8 keV unabsorbed flux. At d = 8.5 kpc, Fpl = 10- 14 ergs cm- 2 s- 1 corresponds to 8.6 × 1031 ergs s- 1.
cBolometric luminosity at d = 8.5 kpc.
dFraction of 10,000 simulated spectra for which the C-statistic was better than the C-statistic obtained for the actual spectrum.
--
0.49+2.76
- 0.12
0.9+160
- 0.7
--
0.63+0.22
- 0.20
0.6+36.9
- 0.5
240+60
- 50
110+180
- 90
(1032 ergs s- 1) Qualityd
0.69
0.97
0.66
0.69
0.987
0.64
TABLE 3
NSA PLUS POWER-LAW FITS
d
kTeff
FNSA
a
(kpc) (eV) (10- 14 ergs cm- 2 s- 1) (1032 ergs s- 1)
b
L∞
Γ
c
Fpl
(10- 14 ergs cm- 2 s- 1)
< 1.1
< 1.1
< 1.0
6 < 57
8.5 < 66
11 < 73
a0.3 -- 8 keV unabsorbed flux.
bBolometric luminosity as seen by a distant observer using 4πR2 σT 4
c0.3 -- 8 keV unabsorbed flux. At d = 8.5 kpc, Fpl = 10- 14 ergs cm- 2 s- 1 corresponds to 8.6 × 1031 ergs s- 1.
0.56+0.82
- 0.42
0.63+0.82
- 0.60
0.77+0.65
- 0.77
eff(1 - 2GM/Rc2), where R = 10 km and M = 1.4 M⊙.
2.3+1.4
- 2.5
2.5+1.4
- 3.1
2.8+1.9
- 3.4
< 0.80
< 1.4
< 2.1
NSA PARAMETERS FOR SHORT OUTBURST FIELD NS SXTS AND X-RAY ACTIVITY
TABLE 4
b
R
gc
L∞
(km)
kTeff
(eV)
(ergs s- 1)
Sourcea
d
(kpc)
< M>d
( M⊙ yr- 1)
94 - 108e 13.2 (5.3 - 9.4) × 1033 0.828 1.3 × 10- 10
0.748 6.9 × 10- 11
0.824 1.9 × 10- 11
0.765 1.7 × 10- 12
0.765 2.9 × 10- 12
5.3 × 1033
4.8 × 1032
< 2.0 × 1032
< 1.4 × 1032
5
Aql X-1
3.6 170 ± 30 9.4
4U 1608 -- 522
1.2
12.9
Cen X-4
10
SAX J1810.8 -- 2609 4.9
XTE J2123 -- 058
8.5
10
aParameters for sources other than XTE J2123 -- 058 are from Rutledge et al. (1999, 2001, 2002a) and Jonker, Wijnands & van der Klis (2004).
bBolometric luminosity as seen by a distant observer calculated using 4πR2σT 4
cg = p1 - 2GM/Rc2 assuming M = 1.4 M⊙.
dThe time-averaged mass accretion rate over the past 33 years, estimated according to < M>= s¯LpeakN, where s = 1.1 × 10- 23 s2 cm- 2 (see Equation 1).
eThe range of values found in three separate Chandra observations.
76 ± 7
< 72
< 66
effg2.
|
astro-ph/0102410 | 2 | 0102 | 2001-06-29T15:27:11 | Broad Absorption Line Quasars in the Sloan Digital Sky Survey with VLA-FIRST Radio Detections | [
"astro-ph"
] | We present 13 Broad Absorption Line (BAL) quasars, including 12 new objects, which were identified in the Sloan Digital Sky Survey (SDSS) and matched within 2'' to sources in the FIRST radio survey catalog. The surface density of this sample of radio-detected BAL quasars is 4.5 +- 1.2 per 100 deg^2, i.e. approximately 4 times larger than previously found by the shallower FIRST Bright Quasar Survey (FBQS). A majority of these radio-detected BAL quasars are moderately radio-loud objects. The fraction of BAL quasars in the entire radio quasar sample, 4.8 +- 1.3 %, is comparable to the fraction of BAL quasars among the SDSS optical quasar sample (ignoring selection effects). We estimate that the true fraction of BAL quasars (mostly HiBALs) in the radio sample is 9.2 +- 2.6 % once selection effects are accounted for. We caution that the absorption troughs of 4 of the 13 radio-detected quasars considered do not strictly satisfy the standard BALnicity criterion. One or possibly two of the new radio-detected BAL quasars are of the rare ``iron LoBAL'' type. BAL quasars are generally redder than the median SDSS quasar at the same redshift. | astro-ph | astro-ph | Draft version October 29, 2018
Preprint typeset using LATEX style emulateapj v. 04/03/99
BROAD ABSORPTION LINE QUASARS IN THE SLOAN DIGITAL SKY SURVEY WITH
VLA-FIRST RADIO DETECTIONS
Kristen Menou1, Daniel E. Vanden Berk2, Zeljko Ivezi´c1, Rita S.J. Kim,1 Gillian R.
Knapp1, Gordon T. Richards3, Iskra Strateva1, Xiaohui Fan4, James E. Gunn1, Patrick
5, Tim Heckman6, Julian Krolik6, Robert H. Lupton1, Donald P. Schneider3,
B. Hall1
Donald G. York7
8, S.F. Anderson9, N.A. Bahcall1, J. Brinkmann10, R. Brunner11, I.
Csabai6, M. Fukugita12, G.S. Hennessy13, P.Z. Kunszt6, D.Q. Lamb7, J.A. Munn14, R.C.
,
,
1
0
0
2
n
u
J
9
2
2
v
0
1
4
2
0
1
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
Nichol15, G.P. Szokoly16
Draft version October 29, 2018
ABSTRACT
We present 13 Broad Absorption Line (BAL) quasars, including 12 new objects, which were identified
in the Sloan Digital Sky Survey (SDSS) and matched within 2′′ to sources in the FIRST radio survey
catalog. The surface density of this sample of radio-detected BAL quasars is 4.5 ± 1.2 per 100 deg2,
i.e. approximately 4 times larger than previously found by the shallower FIRST Bright Quasar Survey
(FBQS). A majority of these radio-detected BAL quasars are moderately radio-loud objects. The fraction
of BAL quasars in the entire radio quasar sample, 4.8±1.3%, is comparable to the fraction of BAL quasars
among the SDSS optical quasar sample (ignoring selection effects). We estimate that the true fraction of
BAL quasars (mostly HiBALs) in the radio sample is 9.2 ± 2.6% once selection effects are accounted for.
We caution that the absorption troughs of 4 of the 13 radio-detected quasars considered do not strictly
satisfy the standard BALnicity criterion. One or possibly two of the new radio-detected BAL quasars
are of the rare "iron LoBAL" type. BAL quasars are generally redder than the median SDSS quasar at
the same redshift.
Subject headings: quasars: general, absorption lines -- galaxies: active -- radio continuum: general --
catalogs -- surveys
1.
INTRODUCTION
The spectra of Broad Absorption Line (BAL) quasars
show broad, blueshifted absorption troughs corresponding
to highly-ionized restframe-UV transitions such as C iv,
Si iv, N v, O vi, more rarely Mg ii and Al iii and even
more rarely Fe ii (e.g. Hazard et al. 1987; Weymann et al.
1991; Becker et al. 1997). The similarity of the continuum
and line emission of BAL and non-BAL quasars motivates
the hypothesis that BAL quasars are not intrinsically dif-
ferent from other quasars. The presence of BAL features
in the spectra of only ∼ 10% of optically-selected quasars
could naturally be explained by a difference in viewing an-
gle if the subrelativistic outflow at the origin of the BAL
features is not isotropic (Weymann et al. 1991). The pop-
ular notion that the outflow is preferentially located in the
plane of the disk surrounding the supermassive black hole
has found support in spectro-polarimetric measurements
for BAL quasars (e.g. Goodrich & Miller 1995; Cohen et
al. 1995; see, e.g., Murray et al. 1995 for a theoretical
wind model).
This geometrical interpretation of BAL features sug-
gests the possible existence of links between the BAL char-
acteristics of a quasar and its radio properties (such as jet
orientation, radio spectral index or even radio-loudness).
Along those lines, Stocke et al. (1992) found that their
sample of optically-selected BAL quasars revealed only
radio-quiet sources, while BAL quasars were strikingly ab-
sent from a sample of radio-loud objects. The authors pro-
posed that the BAL phenomenon is simply anticorrelated
with the mechanism giving rise to strong radio emission
in quasars; they also noted the tendency for radio-loud
1Princeton University Observatory, Princeton, NJ 08544
2Fermi National Accelerator Laboratory, P.O. Box 500, Batavia, IL 60510
3Department of Astronomy & Astrophysics, Pennsylvania State University, University Park, PA 16802
4Institute for Advanced Study, Olden Lane, Princeton, NJ 08540
5 Pontificia Universidad Cat´olica de Chile, Departamento de Astronom´ıa y Astrof´ısica, Facultad de F´ısica, Casilla 306, Santiago 22, Chile
6 Department of Physics & Astronomy, The Johns Hopkins University, 3701 San Martin Drive, Baltimore, MD 21218, USA
7The University of Chicago, Astronomy & Astrophysics Center, 5640 S. Ellis Ave., Chicago, IL 60637
8Enrico Fermi Institute, 5640 So. Ellis Ave., Chicago, IL 60615
9University of Washington, Department of Astronomy, Box 351580, Seattle, WA 98195
10Apache Point Observatory, P.O. Box 59, Sunspot, NM 88349-0059
11Department of Astronomy, California Institute of Technology, Pasadena, CA 91125
12Institute for Cosmic Ray Research, University of Tokyo, Midori, Tanashi, Tokyo 188-8502, Japan
13U.S. Naval Observatory, 3450 Massachusetts Avenue NW, Washington, DC 20392-5420
14U.S. Naval Observatory, Flagstaff Station, P.O. Box 1149, Flagstaff, AZ 86002-1149
15Department of Physics, Carnegie Mellon University, 5000 Forbes Ave., Pittsburgh, PA-15232
16Astrophysikalisches Institut Potsdam, An der Sternwarte 16, D-14482 Potsdam, Germany
1
2
RADIO BAL QUASARS IN THE SDSS
quasars at the high-end of the radio-loudness distribution
to lack high-velocity BALs (see also Weymann et al. 1991).
Other interpretations of the BAL phenomenon have been
presented, such as the evolutionary scenario of Briggs,
Turnshek & Wolfe (1984; see also Gregg et al. 2000) in
which BAL quasars later become radio-loud objects.
The mJy flux-limited VLA-FIRST survey (Becker,
White & Helfand 1995) has challenged many of the stan-
dard views concerning radio quasars and the BAL phe-
nomenon.
In the FIRST Bright Quasar Survey (FBQS;
Gregg et al. 1996; White et al. 2000), FIRST sources with
point-source optical counterparts brighter than 17.8 in the
APM catalog POSS-I plates were systematically targeted
for spectroscopic identification. The FBQS, besides chal-
lenging the existence of a clear radio-dichotomy for quasars
(White et al. 2000), has established beyond doubt the ex-
istence of a population of radio-loud BAL quasars (Becker
et al. 2000; see also Hazard et al. 1987; Becker et al. 1997,
Brotherton et al. 1998; Wills, Brandt & Laor 1999). Fur-
thermore, the population of radio-detected BAL quasars
identified by the FBQS was found to exhibit a diversity of
radio spectral indices (both steep and flat spectra) which
may not support an interpretation of BAL quasars as pref-
erentially edge-on oriented systems (Becker et al. 2000; see
also Gregg et al. 2000).
In this paper, we describe preliminary results for BAL
quasars spectroscopically identified by the Sloan Digital
Sky Survey (SDSS) and matched within 2′′ to radio coun-
terparts in the FIRST survey catalog. These results are
part of a more ambitious project aimed at studying all
SDSS sources with FIRST radio detections (see Ivezi´c et
al. 2001; Knapp et al. 2001). Our main result is that,
at the greater depth probed by SDSS in the optical, the
surface density of radio-detected BAL quasars is ∼ 4 times
larger than previously reported by the FBQS, with a ma-
jority of moderately radio-loud objects.
In §2, we describe how SDSS and FIRST sources were
selected to be part of the sample of BAL quasars discussed.
The sample and some of its characteristics are described
in §3, while results concerning the optical (SDSS photo-
metric) properties are emphasized in §4. We conclude in
§5.
2. TARGET SELECTION
2.1. SDSS Photometric and Spectroscopic Systems
The Sloan Digital Sky Survey (SDSS; York et al. 2000)
uses a camera with an array of CCDs (Gunn et al. 1998)
on a dedicated 2.5-m telescope (Siegmund et al. 2001) at
Apache Point Observatory, New Mexico, to obtain images
in five broad optical bands over 10,000 deg2 of the high
Galactic latitude sky (b ∼> 30o) centered approximately
on the North Galactic Pole. The five filters (designated
u′, g ′, r′, i′, and z ′) cover the entire wavelength range of
the CCD response, longward of the atmospheric cutoff at
short wavelengths (Fukugita et al. 1996). Photometric
calibration is provided by simultaneous observations with
a 20-inch telescope at the same site. The survey data pro-
cessing software measures the properties of each detected
object in the imaging data, and determines and applies as-
trometric and photometric calibrations (Pier et al. 2001;
Lupton et al. 2001).17
Based on their optical morphology and colors, a subset
of the sources discovered in the imaging survey is selected
for spectroscopic follow-up. The spectroscopic survey uses
two fiber -- fed double spectrographs designed to cover the
wavelength range 3800 -- 9200 A. The SDSS spectrographs
achieve a spectral resolution of ∼ 1800 across the entire
spectral range. Each spectrograph accepts 320 fibers, each
of which subtends a diameter of 3
on the sky. Exposures
are typically ∼ 45 − 60 minutes long. Details concerning
the SDSS spectroscopic system are given by Uomoto et al.
(2001) and Frieman et al. (2001).
′′
Quasar candidates are selected from their point-source
optical morphology18 and optical colors such that they lie
outside the stellar locus in color-color space (blue colors for
z ∼< 2 quasars -- see box in Fig. 2a -- and red colors for z ∼> 3
quasars; Fan 1999; Fan et al. 1999). In addition, point
sources with radio counterparts within 2′′ in the FIRST
catalog are favored candidates for spectroscopy (because
there is very little stellar contamination in radio matches,
independent of the optical colors; Helfand et al. 1999;
Ivezi´c et al. 2001; Knapp et al. 2001). The quasar selec-
tion criteria used for this sample was not constant through-
out the period of observations; however, all otherwise good
quasar candidates with reddening corrected i∗ magnitudes
brighter than 19 were flagged as spectroscopic targets. In
addition, other quasar candidates (mostly high-redshift
candidates) were selected as faint as i∗ = 20.5. Radio-
detected quasar candidates were selected to the brighter
of these limits, unless they were also high-redshift candi-
dates (see Richards et al. 2001b for a detailed account on
the selection for spectroscopic follow-up).
2.2. Matching to the VLA-FIRST Survey Catalog
The identification in the FIRST catalog of several hun-
dreds of radio counterparts to SDSS optical sources shows
that there is very little contamination by spurious associa-
tions at angular separations ∼< 2′′ (Ivezi´c et al. 2001). The
FIRST catalog provides both peak and integrated 1.4 GHz
flux densities down to 1 mJy (corresponding to a ∼ 5σ
detection). The source radio morphology can be studied
from the FIRST survey images.19 For comparison, the
flux limit for the radio sample of Stocke et al. (1992) is
typically 0.2-0.3 mJy.
3. RESULTS
3.1. The Sample of Quasars
We report on results concerning 60 SDSS spectroscopic
plates, covering a total sky area of ≈ 290 deg2. The total
area was calculated by summing the areas corresponding
to individual spectroscopic plates and by taking into ac-
17SDSS nomenclature: the names for sources have the format SDSSp Jhhmmss.ss+ddmmss.s, where the coordinate equinox is J2000, and
the "p" refers to preliminary. The reported magnitudes are asinh magnitudes (Lupton, Gunn & Szalay 1999) and are based on a preliminary
photometric calibration; to indicate this the filters have an asterisk instead of a prime superscript. The estimated current astrometric accuracies
in each coordinate are 0.1′′ and the calibration of the photometric measurements are currently accurate to 0.02 -- 0.05 magnitudes.
18At low redshift, this morphology cut is not enforced so that extended AGN are included in the spectroscopic sample as well (Richards et
al. 2001b).
19FIRST website: http://sundog.stsci.edu
MENOU ET AL.
3
count the overlaps between the plates. The completeness
of the radio sample is expected to be nearly 100% because
only spectroscopic plates corresponding to regions covered
by the VLA-FIRST survey were considered. Overall, a
total of 2326 targets were spectroscopically identified as
quasars (see Vanden Berk et al. 2001 for definition; see
also Richards et al. 2001a), 96 of which show BAL fea-
tures as determined by visual inspection. A subset of 272
quasars have reliable FIRST detections (i.e. radio coun-
terparts within 2′′), while 13 of the 96 BAL quasars do.20
The focus of this paper is on these 13 radio-detected BAL
objects, whose properties are listed in Table 1: J2000
coordinates, SDSS optical magnitudes u∗, g ∗, r∗, i∗, z ∗
(reddening-corrected), peak radio flux in the FIRST cat-
alog, BALnicity index (see definition below), maximum
outflow velocity Vmax, a representative absolute magni-
tude, Mg∗ , radio-loudness R∗ (k-corrected ratio of radio
to optical fluxes), 1.4 GHz specific radio luminosity LR
(in erg−1 s−1 Hz−1), redshift z and BAL classification.
The absolute magnitudes and specific radio luminosi-
ties were calculated assuming the following cosmology:
Ωm = 1, Ωλ = 0 (q0 = 0.5) and H0 = 50 km s−1 Mpc−1.
The "BALnicity" index is a measure of the strength of
an absorption trough, similar to an equivalent width, but
requiring continuous absorption of at least 10% in depth
and spanning at least 2000 km s−1, and excluding the re-
gion within 3000 km s−1 of the quasar emission redshift
(Weymann et al. 1991). C iv is used in all cases for which
it is accessible and Mg ii is used otherwise. The quantity
"Vmax" is the maximum outflow velocity of the absorption
lines, relative to the quasar emission redshift. We only
list peak flux densities because of their close agreement
with the integrated values for all the sources, as expected
for radio point sources. The compact radio morphology
of the 13 BAL quasars listed was confirmed by inspecting
the FIRST survey images for each object.
of
one
can be
A NASA/IPAC Extragalactic Database
objects
(NED)
(SDSSp
the
that
search reveals
identified with PKS
J115944.81+011207.1)
J1159+0112,
that another one (SDSSp J003923.20-
001452.7) was detected as a radio source in the NVSS
survey (with a compact radio morphology and a peak flux
density consistent with the FIRST values) and that none
of the objects listed has a counterpart in the IRAS survey
catalog. No counterparts were found in the ROSAT All
Sky Survey or 2MASS catalogs either (only 8 of the 13
listed objects are located in the sky region covered by the
2MASS Second Incremental Data Release21).
Following Becker et al. (2000), we do not limit our BAL
sample to the conservative definition of Weymann et al.
(1991), which explains why four of the objects listed have
a zero BALnicity index (i.e. if an absorption feature sat-
isfied all the criteria of Weymann et al. except being be-
yond 3000 km s−1 of the quasar emission redshift, the ob-
ject was still considered as a potential BAL quasar). The
justifications given by Becker et al. (2000) for including
objects which do not strictly satisfy the BALnicity crite-
rion are as follows: (i) nearly black absorption spanning
∼ 4000 km s−1, which is unlikely to break up into narrow
lines, (ii) several absorption systems with a large Vmax
value; none in excess of 2000 km s−1, but taken together
they are suggestive of an intrinsic BAL outflow, (iii) ev-
idence for partial covering, which is a property of BAL
quasars and (iv) low-ionization lines in BAL quasars tend
to be narrower than high-ionization lines (Voit, Weymann
& Korista 1993), so that any BALnicity index which had
to be calculated from low-ionization lines (because high-
ionization lines could not be observed) are likely to be a
lower limit. These same criteria motivate us to include
four objects with a zero BALnicity index in our sample of
radio-detected BAL quasars. SDSSp J003923.20-001452.7
has a measured BALnicity index > 0, but rounded down to
0; it satisfies criterion (i) and maybe criterion (iii). SDSSp
J235702.55-004824.0 satisfies criterion (i) and maybe cri-
terion (ii). SDSSp J115944.81+011207.1 satisfies criterion
(i) except for Vmax ∼ 3000 km s−1; it satisfies criterion
(ii): there is an additional broad absorption system at
Vmax ∼ 8000 km s−1; it also appears to satisfy criterion
(iii): the C iv absorption trough is nearly flat-bottomed
with a residual flux indicating partial coverage. SDSSp
J133903.41-004241.2 is the least convincing of its category;
Mg ii absorption is black with Vmax = 1650 km s−1 , but
the noisy C iv part of the spectrum may allow it to satisfy
criterion (iv).
The radio-loudness R∗ is calculated as the k-corrected
ratio of the 5 GHz radio flux to the 2500 A optical flux
(which is calculated from the B-band magnitude), follow-
ing the definition of Stocke et al. (1992). We adopt the
photometric transformation B ≃ g ∗ + 0.13 which is ap-
propriate for the power-law spectra of quasars (see, e.g.,
Schmidt, Schneider & Gunn 1995) and we assume a power-
law index αr = −0.5 (with fν ∝ να) in the radio (the
same value of the radio power-law index was assumed for
the k-correction when calculating LR). A power-law index
αo = −0.5 was assumed when calculating the absolute
magnitudes (see Vanden Berk et al. 2001 for a measure
of αo from an SDSS quasar composite spectrum; see also
Richards et al. 2001a).
The redshifts were determined automatically by the
SDSS spectroscopic pipeline and modified as necessary
upon visual inspection (BAL quasar redshifts prove dif-
ficult to measure automatically with precision but the
modifications were generally not more than a few hun-
dredths of a unit redshift).
The BAL classification
refers to the ionization states seen in absorption: Hi-
BALs are defined as those exhibiting absorption only by
high-ionization ions, LoBALs show both high- and low-
ionization lines, and FeLoBALs are LoBALs with lines
from meta-stable Fe ii and Fe iii. Examples of specific
lines are Mg ii λ2800 for LoBAL (as well as, e.g., lines of
Al ii, Al iii, Si ii, and C ii), C iv λ1549 for HiBAL, and
Fe ii λ1063, 2261, 2380, 2600, 2750 and Fe iii λ1122, 1914
for FeLoBAL (Hazard et al. 1987). Of course, all the lines
cannot be seen for all the quasars because of the finite
spectral coverage in the observer frame.
We note that one or possibly two of the 13 radio-
detected BAL quasars reported here are of the rare
FeLoBAL type, of which the FBQS has found only four ex-
20In what follows, we refer to objects without a FIRST radio detection as "radio-undetected" and objects with a FIRST radio detection as
"radio-detected".
21Available at http://www.ipac.caltech.edu/2mass/releases/second/index.html
4
RADIO BAL QUASARS IN THE SDSS
amples (Becker et al. 2000) and only one other has been re-
ported (Hazard et al. 1987). Five of the six known FeLoB-
ALs have radio detections, which supports the speculation
by Becker et al. (1997) that these quasars are associated
with rather strong radio emission. We note, however, that
the number of FeLoBALs in the radio-undetected sample
is presently unknown and that these results still concern
a small number of sources. The spectra of the 13 radio-
detected BAL quasars listed in Table 1 are shown in Fig-
ure 1.
3.2. General Properties
Despite the small size of this sample of radio-detected
BAL quasars, some interesting conclusions can be drawn.
The surface density of radio-detected BAL quasars at the
depth probed by SDSS is 4.5 ± 1.2 objects per 100 deg2,
which is approximately 4 times larger than reported by
the FBQS (with the same definition for the BAL sam-
ple; Becker et al. 2000). This is expected given that the
SDSS goes deeper than the limiting magnitude of 17.9 for
the FBQS. Adopting a dividing line between radio-quiet
and radio-loud objects at R∗ = 10 (following Stocke et
al. 1992), we see that most of the objects identified by
SDSS and found in the FIRST catalog are of a moder-
ately radio-loud (10 < R∗ < 100) to strongly radio-loud
(R∗ > 100) nature. This is the result of the greater depth
probed by SDSS in the optical (as compared to the APM
digitized POSS-I catalog used for the FBQS), which, given
the mJy flux limit of the FIRST survey, tends to uncover
objects which are more radio-loud than previously known.
The specific radio luminosity LR alone is also sometimes
used as a direct measure of the radio-loudness of quasars.
Adopting a dividing line between radio-quiet and radio-
loud objects at LR = 1032.5h−2
50 erg−1 s−1 Hz−1 (follow-
ing Gregg et al. 1996 and Stern et al. 2000), we see
again that a majority of the objects listed in Table 1
are on the radio-loud side.22 A more conservative cut at
1033 erg−1 s−1 Hz−1 still results in 7 out of the 13 objects
being on the radio-loud side. Thus, we find no evidence of
the anticorrelation of the radio-loud emission mechanism
with the BAL phenomenon which was found by Stocke et
al. (1992). The trend of moderate radio-loudness in our
sample of radio-detected BAL quasars may, however, be
consistent with the tendency found by Stocke et al. (1992;
see also Weymann et al. 1991) for strongly radio-loud
quasars to lack high-velocity BALs.
One must temper the conclusion of this analysis, how-
ever, by a few words of caution.
In particular, we note
that the BALnicity distribution of the radio-detected BAL
quasars listed in Table 1 appears different (at a ≈ 2σ level)
from that of the sample of radio-quiet BAL quasars of
Weymann et al. (1991). While ∼ 25% of our BAL quasars
have a BALnicity index > 3000 km s−1, ∼ 70% of the
BAL quasars discussed by Weymann et al.
(1991) do.
These authors warn that quasars with a BALnicity index
< 1500 km s−1 risk contamination by unusually strong
"associated narrow absorbers" (as opposed to bona fide
BAL features), which may be preferentially found in radio-
louder objects.
While nine of our radio-detected BAL quasars have a
BALnicity index < 1500 km s−1, we note that those with
a non-zero BALnicity index have large enough Vmax values
to avoid contamination by associated narrow absorbers.
The motivation for including the remaining four objects
with zero BALnicity indices was given in §3.1. The overall
low BALnicity of the sample is evident23 and it is inter-
esting that the objects with zero BALnicity indices are
among the radio-loudest ones. For this reason, we also
quote radio-detected BAL fractions excluding the 4 sys-
tems with a zero BALnicity index and refer to them in
what follows as "conservative values". Interestingly, only
∼ 30% of the radio-detected BAL quasars found by the
FBQS have a BALnicity index > 3000 km s−1, in agree-
ment with our results. The discrepancy with the Wey-
mann et al. sample may therefore indicate a real differ-
ence between the populations of radio-quiet and radio-loud
BAL quasars. We also note that the redshift distribution
of our radio-detected BAL quasars appears different from
that of the FBQS radio BAL quasars: none of our objects
is found below z ∼ 1.5, while 11 of the 29 FBQS objects
are in this low redshift range.
Our results concerning various interesting fractions for
the quasar sample are summarized in Table 2. In the total
SDSS quasar sample, 11.7 ± 0.7% of all spectroscopically
identified quasars have radio detections in the FIRST sur-
vey, in agreement with a fraction 13.5 ± 3.8% for the BAL
subsample only. The fraction of BAL quasars in the entire
SDSS quasar sample is 4.1±0.4%, which is consistent with
the fraction 4.8 ± 1.3% (conservative value: 3.3 ± 1.1%) of
BAL quasars in the subsample of sources with radio de-
tection. These values suggest that BAL quasars are not
preferentially found among radio-detected quasars (in con-
tradiction to Becker et al. 2000 for the FBQS), but are
found in equal number among sources with and without a
FIRST radio detection. This apparent discrepancy may
be explained by the necessity for Becker et al.
(2000)
to compare their BAL quasar fraction (after correction
for the selection effects) to those obtained in independent
optical surveys. The combination of SDSS optical data
and FIRST radio data should provide a more robust es-
timate for the radio-detected and radio-undetected BAL
fractions; a larger number of sources is required to settle
this issue with confidence.
The fractions of BAL quasars quoted above do not take
into account the strong selection effects which affect the
identification of BAL quasars (the restframe-UV absorp-
tion features must be conveniently located in the optical
spectral range observed). Accounting for these selection
effects is usually done by assuming that there is no sig-
nificant redshift dependence of the BAL quasar fraction.
Under this assumption, we calculate the corrected frac-
tion as the ratio of the number of BAL quasars to the
22Note that a FIRST source with a 1 mJy flux (the survey limit) must be at a redshift larger than 3 or so to automatically satisfy the above
luminosity-based radio-loudness criterion. Although two of our objects have z > 3, both of them have flux densities four times brighter than
the FIRST survey limit.
23We note that the absorption troughs in SDSSp J115944.81+011207.1 and SDSSp J235702.55-004824.0 could be characterized as associated
absorption lines in the sense that they are contained within the corresponding broad emission line. On the other hand, FIRST sources being
preferentially flat-spectrum sources, these two objects may not necessarily belong to the class of systems with associated absorption lines and
usually steep radio spectra (see, e.g., Hamann et al. 2001).
MENOU ET AL.
5
total number of quasars in the redshift range such that
the detection of a specific restframe-UV absorption line
is guaranteed given the SDSS spectral coverage (see, e.g.,
Becker et al. 2000 for a similar procedure for the FBQS).
The redshift range relevant to LoBAL quasars is
0.4 ∼< z ∼< 1.7 based on Mg ii absorption. However, we
find that Al iii absorption is present in nearly all the spec-
tra showing Mg ii absorption. This effectively pushes the
maximum detection redshift for LoBALs to z ≈ 3.9. This
situation illustrates how the uniform signal to noise and
extended spectral coverage of the SDSS will allow to de-
fine more sophisticated BAL fractions (based on several
specific lines) which should be less subject to selection ef-
fects than usual. A total of 249 radio-detected quasars, in-
cluding 7 LoBAL quasars, are found in the redshift range
0.4 ∼< z ∼< 3.9 (based on both Mg ii and Al iii absorption),
so that the corrected LoBAL fraction is 2.8±1.1% (conser-
vative value: 2.0 ± 0.9%) in the sample of radio-detected
objects.
The redshift range relevant to HiBAL quasars is z ∼> 1.4,
so that C iv absorption would be seen if present. A total
of 141 radio-detected quasars, including 13 HiBAL quasars
(all the quasars listed in Table 1 show high-ionization BAL
features), are found in this redshift range, so that the cor-
rected HiBAL fraction is 9.2 ± 2.6% (conservative value:
∼ 6.4 ± 2.1%) in the sample of radio-detected objects. For
the FBQS, Becker et al. (2000) find a corrected LoBAL
fraction of ∼ 3%, which is consistent with the ∼ 2.8%
above, but a corrected HiBAL fraction of 18 ± 3.8%, which
is significantly higher than the 9.2 ± 2.6% HiBAL fraction
of the present radio sample. We note that the HiBAL
fraction in optical samples such as the LBQS is consistent
with the ∼ 9% quoted here (e.g. Foltz et al. 1990). The
same exercise for the radio-undetected BAL quasars re-
quires a careful BAL classification for these objects, which
is beyond the scope of this preliminary communication.
3.3. Optical Properties
The optical properties of the sample of radio-detected
BAL quasars are further described in Figures 2 and 3,
where similar properties for the sample of radio-undetected
BAL quasars are also shown. Fig 2a and 2b illustrate how
most of the 96 BAL quasars were selected for spectroscopy,
based on their location in the (u∗−g ∗, g ∗−r∗, r∗−i∗) SDSS
color space. In one color plane or the other, most BAL
quasars lie outside the stellar locus, indicated by contours
and small dots (for individual objects) for an SDSS con-
trol sample of 40, 000 point sources, almost all of which are
stars. The photometric quality of the control sample was
guaranteed by applying the following cuts: u∗, g ∗, r∗ < 21
and corresponding photometric errors < 0.1. The BAL
quasars only fill the red side of the low-z quasar box (in-
dicated by dashed lines in Fig. 2a) because of their rather
large redshifts (z ∼> 1) and comparatively red colors (see
below). The low-z quasar box approximates the region
of color-space where the majority of z < 2.2 quasars are
found. The circled objects (radio-detected BAL quasars)
are evenly distributed among the BAL quasar sample in
Fig. 2, showing that there is no detectable color difference
between the radio-detected and radio-undetected objects.
Figure 3a shows a redshift-magnitude (i∗) diagram
for the BAL quasar sample, showing again no obvious
trend separating the radio-detected sources from the radio-
undetected ones. Since the standard photometric cut ap-
plied in SDSS for spectroscopic follow-up is at i∗ < 19,
Fig. 3a suggests that the sample of radio-detected BAL
quasars is not strongly biased toward faint magnitudes
(relative to the rest of the BAL sample) as could have
resulted from the favored spectroscopic selection for radio-
detected sources.
Figure 3b shows a redshift-color (g ∗ − i∗) diagram for
the BAL quasar samples. The g ∗ − i∗ color was chosen
so that systematics appearing when strong emission lines
enter or leave photometric bands (especially the r∗ band;
Richards et al. 2001a) are minimized. The relation for the
median color as a function of redshift found by Richards et
al. (2001a) for a nearly identical sample of 2625 spectro-
scopically identified quasars with SDSS colors is shown as
a solid line, while the dashed lines indicate the 95% confi-
dence limits. The location of many BAL quasars above the
median value (solid line) clearly shows that BAL quasars
are redder than the typical quasars spectroscopically iden-
tified by SDSS, in agreement with previous results on the
colors of BAL quasars (Sprayberry & Foltz 1992; Broth-
erton et al. 2000; see also Weymann et al. 1991). This
red nature is interpreted as continuum extinction rather
than attributed to the presence of the absorption troughs
in the BAL quasar spectra (Sprayberry & Foltz 1992; Ya-
mamoto & Vansevicius 1999). This was confirmed here
for the 13 radio-detected BAL quasars by inspection of
the spectra and comparison to the SDSS composite spec-
trum of Vanden Berk et al. (2001). We note that the red
colors of BAL quasars could imply that they are strongly
under-represented in magnitude-limited samples. Further
work is required to address this question properly.
Since ∼ 15% of the radio-undetected objects are be-
low the median g ∗ − i∗ color, ∼ 2 of the radio-detected
BAL quasars are expected below the median value if the
two populations have similar color properties. One cannot
conclude, however, from the absence of any radio-detected
BAL quasar below the solid line that they are redder than
radio-undetected BAL quasars because the samples are too
small at this point to draw statistically significant conclu-
sions.
Interestingly, Richards et al. (2001a) found that
many of the reddest quasars in their sample have FIRST
radio detections.
4. CONCLUSION
We described the properties of 13 BAL quasars spectro-
scopically identified by the SDSS which possess radio coun-
terparts in the VLA-FIRST survey catalog. This sample,
from an area of ≈ 290 deg2, is the second largest sample
of radio-detected BAL quasars after that reported by the
FIRST Bright Quasar Survey (Becker et al. 2000).
Despite the small statistical size of this sample, we were
able to isolate some of its important characteristics. Be-
cause of the SDSS limiting magnitude of ∼ 19 − 20 for our
sample, compared to 17.9 for the FIRST Bright Quasar
Survey, we find a density of radio-detected BAL quasars
of ∼ 4.5 objects per 100 deg2,
four times larger
than in the FIRST Bright Quasar Survey. A majority
of the newly identified radio-detected BAL quasars are
moderately radio-loud objects, in contradiction with early
claims of an anti-correlation between the radio-loud emis-
sion mechanism and the BAL phenomenon. The prefer-
ence for moderate radio-loudness in our sample may in-
i.e.
6
RADIO BAL QUASARS IN THE SDSS
dicate that strongly radio-loud quasars tend to lack high
velocity BALs.
Upon completion, the combination of the full set of
SDSS data with the FIRST catalog should provide us with
several hundreds of radio-detected BAL quasars, allowing
a robust characterization of the properties of this popula-
tion of rare objects.
ACKNOWLEDGMENTS
The Sloan Digital Sky Survey (SDSS) is a joint project of
The University of Chicago, Fermilab, the Institute for Ad-
vanced Study, the Japan Participation Group, The Johns
Hopkins University, the Max-Planck-Institute for Astron-
omy, New Mexico State University, Princeton University,
the United States Naval Observatory, and the University of
Washington. Apache Point Observatory, site of the SDSS
telescopes, is operated by the Astrophysical Research Con-
sortium (ARC). Funding for the project has been provided
by the Alfred P. Sloan Foundation, the SDSS member in-
stitutions, the National Aeronautics and Space Adminis-
tration, the National Science Foundation, the U.S. Depart-
ment of Energy, Monbusho, and the Max Planck Society.
The SDSS Web site is http://www.sdss.org/.
This research has made use of the NASA/IPAC Extra-
galactic Database (NED) which is operated by the Jet
Propulsion Laboratory, California Institute of Technology,
under contract with the National Aeronautics and Space
Administration.
We are grateful to M. Strauss for useful discussions,
R. Becker for comments on the manuscript and the ref-
eree for a very useful report. Support for this work was
provided by NASA through Chandra Postdoctoral Fellow-
ship grant number PF9-10006 awarded (to KM) by the
Chandra X-ray Center, which is operated by the Smith-
sonian Astrophysical Observatory for NASA under con-
tract NAS8-39073. GRK and IS acknowledge support from
NASA grant NAG-3364. DPS and GTR acknowledge sup-
port from NSF grant AST99-00703.
REFERENCES
Becker, R.H., Gregg, M.D., Hook, I.M., McMahon, R.G., White, R.L.
& Helfand, D.J., 1997, ApJ, 479, L93
Becker, R.H., White, R.L. & Helfand, D.J., 1995, ApJ, 450, 559
Becker, R.H. et al., 2000, ApJ 538, 72
Briggs, F.H., Turnshek, D.A. & Wolfe, A.M., 1984, ApJ, 287, 549
Brotherton, M.S., et al., 1998, ApJ, 505, L7
Brotherton, M.S., Tran, H.D., Becker, R.H., Gregg, M.D.,
Laurent-Muehleisen, S.A. & White, R.L., 2000, ApJ, in press,
astroph/0008396
Cohen, M.H. et al., 1995, ApJ, 448, L77
Fan, X., 1999, AJ, 117, 2528
Fan, X. et al. 1999, AJ, 118, 1
Foltz, C., Chaffee, F., Hewett, P., Weymann, R. & Morris, S. 1990,
BAAS, 21, 806
Frieman, J.A. et al., 2001, in preparation
Fukugita, M., Ichikawa, T., Gunn, J.E., Doi, M., Shimasaku, K., &
Schneider, D.P. 1996, AJ, 111, 1748
Goodrich, R.W. & Miller, J.S., 1995, ApJ, 448, L73
Gregg, M.D., Becker, R.H., White, R.L., Helfand, D.J., McMahon,
R.G. & Hook, I.M., 1996, AJ, 112, 407
Gregg, M.D., Becker, R.H., Brotherton, M.S., Laurent-Muehleisen,
S.A., Lacy, M. & White, R.L., 2000, ApJ, 544, 142
Gunn, J.E., et al. 1998, AJ, 116, 3040
Hamann, F.W., Barlow, T.A., Chaffee, F.C., Foltz, C.B. &
Weymann, R.J., 2001, ApJ in press, astroph/0011030
Hazard, C., McMahon, R.G., Webb, J.K. & Morton, D.C., 1987,
ApJ, 323, 263
Helfand, D.J., Schnee, S., Becker, R.H., White, R.L. & McMahon,
R.G., 1999, ApJ, 117, 1568
Ivezi´c, Z, et al., 2001, in preparation
Knapp, G., et al., 2001, in preparation
Lupton, R.H., et al. 2001, in preparation
Lupton, R.H., Gunn, J.E. & Szalay, A.S. 1999, AJ, 118, 1406
Murray, N., Chiang, J., Grossman, S.A. & Voit, G.M., 1995, ApJ,
451, 498
Pier, J.R., et al. 2001, in preparation
Richards, G.T. et al., 2001a, AJ, in press, astroph/0012449
Richards, G.T. et al., 2001b, in preparation
Schmidt, M., Schneider, D.P. & Gunn, J.E., 1995, AJ, 110, 68
Siegmund, W., et al. 2001, in preparation
Sprayberry, D. & Foltz, C.B., 1992, ApJ, 390, 39
Stern, D., Djorgovski, S.G., Perley, R.A., de Carvalho, R.R. & Wall,
J.V., 2000, AJ, 119, 1526
Stocke, J.T., Morris, S.L., Weymann, R.J. & Foltz, C.B., 1992, ApJ,
396, 487
Uomoto, A. et al., 2001, in preparation
Vanden Berk, D.E. et al., 2001, AJ, in press, astroph/0105231
Voit, G.M., Weymann, R.J. & Korista, K.T., 1993, ApJ, 413, 95
Weymann, R.J., Morris, S.L., Foltz, C.B. & Hewett, P.C., 1991, ApJ,
373, 23
Wills, B.J., Brandt, W.N. & Laor, A., 1999, ApJ, 520, L91
Yamamoto, T.M. & Vansevicius, V., 1999, PASJ, 51, 405
York, D.G. et al 2000, AJ, 120, 1579
MENOU ET AL.
7
SDSS Broad Absorption Line Quasars with FIRST Detections
Table 1
RA (J2000) Dec (J2000)
u∗
g∗
r∗
i∗
z∗
(1)
(2)
(3)
(4)
(5)
(6)
(7)
00 39 23.20 −00 14 52.7 20.85 20.39 20.07 19.75 19.41
03 05 43.45 −01 06 22.1 21.13 20.05 19.89 19.54 19.55
11 54 04.14 +00 14 19.5 18.85 18.18
17.9 17.76 17.73
11 59 44.81 +01 12 07.1 18.30 17.48 17.23 16.99 16.72
13 02 08.27 −00 37 31.6 18.75 18.46 17.93 17.59 17.57
13 21 39.86 −00 41 52.0 23.37 20.37 19.25 18.65 18.41
13 23 04.58 −00 38 56.7 18.61 18.54 18.24 17.81 17.77
13 31 50.52 +00 45 18.8 19.28 19.26 19.14 18.91
18.9
13 39 03.41 −00 42 41.2 22.14 21.63 21.02 20.27 20.01
14 01 12.01 +01 11 12.3 20.28 19.89 19.59 19.10 19.02
15 16 36.78 +00 29 40.5 19.83 18.51 17.65 17.26 17.12
16 04 12.39 −00 08 07.9 20.74 19.61 19.42 19.12 18.99
23 57 02.55 −00 48 24.0 22.17 19.38 19.03 18.78 18.53
Sp
(mJy)
(8)
21.2
5.2
1.5
266.5
11.2
4.1
8.9
2.9
2.0
3.1
2.2
1.4
4.0
BALnicitya
(km s−1) (km s−1)
Vmax
a Mr
∗
R∗
log(LR)
za
(9)
(10)
(11)
(12)
(13)
(14)
BAL
Type
(15)
0
400
4100
0
1200
3500
300
1300
0
1500
6700
600
0
12150 −25.99
16200 −26.68
10050 −26.48
12150 −25.83
9600 −26.58
12900 −25.94
5100 −25.15 419.3
69.4
4.2
3000 −27.82 377.4
41.3
70.4
34.6
21.9
1650 −23.11 142.6
41.6
7.6
12.6
28.4
10350 −25.17
25000 −27.05
7200 −26.42
4050 −26.77
Hib
Hi
Lo
Hic
Lo
33.62 2.233
33.21 2.850
32.18 1.604
34.63 1.989
33.11 1.672
33.17 3.080 FeLo
33.08 1.821
32.63 1.892
32.28 1.518
32.59 1.771
32.64 2.248 FeLo?
32.64 2.832
33.15 3.005
Hi
Hi
Lo d
Lo
Hi
Lo e
aThe BALnicity values have been rounded to the nearest 100 km s−1 and the Vmax values to the nearest 150 km s−1 to reflect
the estimated precision. The estimated uncertainties in the redshift measurements (based on MgII if available or CIII] otherwise) are
< 0.005 for all the objects, except SDSSp J132139.86-004152.0 for which the redshift measurement is accurate only to 0.02.
bNED: NVSS source J003923-001452; possibly MgII BAL. Measured BALnicity is > 0 (rounded here), but value is consistent with
0 given spectral resolution and flux density errors.
cNED: identified with PKS J1159+0112; possibly MgII BAL; Vmax < 5000 km s−1.
dused MgII BAL; Vmax < 5000 km s−1.
eVmax < 5000 km s−1.
8
RADIO BAL QUASARS IN THE SDSS
Radio and BAL Fractions in the SDSS Quasar Sample
Table 2
Sub-sample
Fraction
Quasars with Radio Detection:
BAL Quasars with Radio Detection:
Quasars with BAL Features (uncorrected):
Radio Quasars with BAL Features (uncorrected):
Radio Quasars with LoBAL Features (corrected):
Radio Quasars with HiBAL Features (corrected):
11.7 ± 0.7%
13.5 ± 3.8%
4.1 ± 0.4%
4.8 ± 1.3% (3.3 ± 1.1%)a
2.8 ± 1.1% (2.0 ± 0.9%)a
9.2 ± 2.6% (6.4 ± 2.1%)a
aThe fraction in parenthesis corresponds to the value obtained when a con-
servative BAL definition is adopted (see text).
MENOU ET AL.
9
Fig. 1. -- Mosaic of rest-frame SDSS spectra for the 13 BAL quasars with FIRST detections (SDSSp J115944.81+011207.1 can be identified
with PKS J1159+0112). The location of Lyβ/O vi, Lyα, N v, O i/Si ii, C ii, Si iv/O iv, C iv, Al iii, C iii] and Mg ii emission lines are
indicated by dotted lines. The spectra have a nearly constant resolution of approximately 1800. They were all smoothed over 3 pixels, except
for the lowest redshift one which was smoothed over 7 pixels.
10
RADIO BAL QUASARS IN THE SDSS
Fig. 1. -- continued
MENOU ET AL.
11
Fig. 2. -- (a) Color-color diagram (u∗ − g∗ vs. g∗ − r∗) for 13 radio-detected BALs (circled symbols) and 96 radio-undetected BALs (bare
symbols). Bright objects (i∗ < 19) are indicated by filled circles and faint ones (i∗ ≥ 19) by filled triangles. The background contours and
dots represent the stellar locus for ∼ 40, 000 point-source objects in an SDSS control sample. The dashed lines show the location of the
low-redshift quasar box (see text for details). (b) The corresponding g∗ − r∗ vs. r∗ − i∗ color-color diagram.
12
RADIO BAL QUASARS IN THE SDSS
Fig. 3. -- Same notation as Fig. 2. (a) Redshift-magnitude (i∗) distribution of the BAL quasars. (b) Color-redshift distribution of the BAL
quasars, compared to the median color-redshift relation (solid line) derived by Richards et al. (2001a; dashed lines indicate 95% confidence
limits). The BAL quasars are redder than the typical quasars spectroscopically identified by SDSS.
|
astro-ph/0006025 | 1 | 0006 | 2000-06-01T19:05:10 | Infrared interferometric observations of young stellar objects | [
"astro-ph"
] | We present infrared observations of four young stellar objects using the Palomar Testbed Interferometer (PTI). For three of the sources, T Tau, MWC 147 and SU Aur, the 2.2 micron emission is resolved at PTI's nominal fringe spacing of 4 milliarcsec (mas), while the emission region of AB Aur is over-resolved on this scale. We fit the observations with simple circumstellar material distributions and compare our data to the predictions of accretion disk models inferred from spectral energy distributions. We find that the infrared emission region is tenths of AU in size for T Tau and SU Aur and ~1 AU for MWC 147. | astro-ph | astro-ph |
Infrared Interferometric Observations of Young Stellar Objects 1
R.L. Akeson2,3, D.R. Ciardi4, G.T. van Belle3, M.J. Creech-Eakman3 and E.A. Lada4
ABSTRACT
We present infrared observations of four young stellar objects using the Palomar
Testbed Interferometer (PTI). For three of the sources, T Tau, MWC 147 and SU
Aur, the 2.2 µm emission is resolved at PTI's nominal fringe spacing of 4 milliarcsec
(mas), while the emission region of AB Aur is over-resolved on this scale. We fit the
observations with simple circumstellar material distributions and compare our data to
the predictions of accretion disk models inferred from spectral energy distributions. We
find that the infrared emission region is tenths of AU in size for T Tau and SU Aur and
∼1 AU for MWC 147.
Subject headings: stars:pre-main sequence, circumstellar matter
1.
Introduction
Observational evidence for circumstellar material around most young stellar objects (YSOs)
includes infrared emission in excess of that expected from the stellar photosphere, broad forbidden
line profiles, and emission at millimeter wavelengths. Although the dust column density is inferred
to be quite high, the sources are often optically visible, implying a geometrically flat distribution
of the material. A disk morphology is also predicted by star formation theories as a consequence
of conservation of angular momentum. Evidence for circumstellar disks has been observed around
sources with a range of masses, from near solar (T Tauri stars) to greater than 10 M⊙ (Herbig
Ae/Be stars) (see e.g. Mundy, Looney and Welch (2000); Natta, Grinin and Mannings (2000)).
Disks not only provide a conduit for material to accrete onto the central star, but are also a reservoir
of material from which a potential planetary system might form.
The structure of YSO circumstellar disks has been studied using spectral energy distributions
(SED), spectral line profiles and imaging at infrared and (sub)-millimeter wavelengths. The dust
continuum emission from disks around several T Tauri sources has been resolved at millimeter
wavelengths (see review by Wilner and Lay (2000)). These observations are sensitive to emission
1to be published in the Astrophysical Journal
2Infrared Processing and Analysis Center, California Institute of Technology MS 100-22, Pasadena, CA, 91125
3Jet Propulsion Laboratory, MS 171-113, 4800 Oak Grove, Pasadena, CA 91109
4University of Florida, 211 Bryant Space Sciences Bldg, Gainesville, FL 32611
– 2 –
from cooler dust and provide spatial information on size scales of several 10's of AU. The disk
physical properties on much smaller scales (< few AU) are generally inferred through examination
of the spectral line shapes and modeling of the SED. Unresolved issues regarding the inner disk
structure include the possible existence of inner disk holes (e.g. Hillenbrand et al. (1992)) and
the validity of simple power law scalings to describe globally the temperature and density profiles
of the disk. Characterizing the physical properties of the inner disk is important for theories of
hydrodynamic disk winds and for understanding the initial conditions of planet formation.
Infrared interferometry provides a method to directly observe the inner disk. To date, only a
few YSOs have been observed using this technique (e.g. FU Ori: Malbet et al. (1998) and AB Aur:
Millan-Gabet et al. (1999)). Here we present K-band long baseline interferometric observations
of four YSOs: T Tau, SU Aur, AB Aur, and MWC 147. PTI has a fringe spacing of ∼4 mas,
corresponding to 0.6 AU at the distance of Taurus-Aurigae (140 pc) and to 3 AU at the distance
of MWC 147 (800 pc).
2. Observations and data reduction
Observations were made in the K band at PTI, which is described by Colavita et al. (1999).
The data were obtained between September and December 1999. For each source, the number of
nights, the total number of records (each of which contains 25 seconds of data) and the calibrators
used are given in Table 1. The data were calibrated using the standard method described in Boden
et al. (1998). A synthetic wideband channel is formed from the five spectrometer channels. The
system visibility is measured with respect to the calibrators. The calibrator sizes were estimated
using a blackbody fit to photometric data from the literature and were confirmed to be internally
consistent when two or more calibrators were observed in a given night, which occurred on most
nights. The calibrators were chosen by their proximity to the sources and for their small angular
size, minimizing systematic errors in deriving the system visibility. All calibrators used in this
reduction have angular diameters < 0.8 mas and were assigned uncertainties of 0.1 mas, except for
HD 46709 (θ = 1.8 ± 0.2 mas). The data are presented in normalized squared visibility, which is an
unbiased quantity. The averaged squared visibility and error are given for each source in Table 1.
The uncertainties for the calibrated visibilities are a combination of the calibrator size uncertainty
and the internal scatter in the data.
3. Models
Of the four objects discussed in this paper, three were resolved and the fourth was over-resolved
(no fringes were detected). In this section we first detail the models and then discuss each source
separately in the following section. The models fit to the data are a uniform brightness profile, a
Gaussian brightness profile, and a binary companion. We also compare the data to accretion disk
– 3 –
models. The uniform and Gaussian profiles are presented as simple geometric distributions which
can be used as size scale estimators. At the distances to these systems (140 to 800 pc) the central
star is unresolved (θ < 0.1 mas). For the uniform and Gaussian distributions and the accretion
disk models, a stellar component has been included as an unresolved source with the appropriate
flux ratio. This ratio was determined by subtracting the photospheric flux from a star of the
appropriate spectral type from the total K band flux. For these objects, the infrared emission on
milliarcsec scales is dominated by thermal emission and scattering can be neglected (Malbet and
Bertout 1995).
3.1. Uniform and Gaussian profiles
Two of the simplest geometric models which can be used to describe the circumstellar material
distribution are a uniform brightness profile and a Gaussian profile. For a face-on Gaussian or
uniform profile, the predicted visibility is simply a function of the projected baseline. For the
uniform profile, the squared visibility is
V 2 =(cid:20) 2J1(πθBp/λ)
πθBp/λ
(cid:21)2
(1)
where J1 is a Bessel function, θ is the diameter, and Bp is the projected baseline. For a Gaussian
profile
(2)
V 2 = exp"−
where D is the FWHM.
π2
ln 2(cid:18) D
2(cid:19)2 B2
λ #!2
p
For an inclined profile, the visibility is also a function of hour angle. As none of the sources show
definitive visibility structure with hour angle, we limit ourselves to the simple face-on case. The
observations cover hour angle ranges of -2 to 2 hours for T Tau, -2 to 0 hours for SU Aur and -3 to 1
hours for MWC 147. Although other inclinations and position angles are not necessarily excluded by
the data, we note that inclination angles near edge-on would produce significant visibility variations
with hour angle, which are not seen. The best fit uniform and Gaussian profile sizes for each source
are given in Table 1.
3.2. Binary companion
A reduction from unity visibility can be produced by a binary companion. If both components
are individually unresolved, the visibility is given by
V 2 =
1 + R2 + 2R cos[(2π/λ)B · s]
(1 + R)2
(3)
– 4 –
where R is the flux ratio, B is the baseline vector and s (in radians) is the binary angular separation
vector. To test if the measured visibilities are consistent with a binary, a grid of binary parameters
was formed and model visibilities were calculated and compared to data binned by projected base-
line. The binary parameter space considered contained primary/secondary flux ratios (R) from 1
to 30 and separations (s) up to 100 mas, which corresponds to the coherence length of the spectral
channels. Binary companions beyond this separation with sufficient magnitude to affect the mea-
sured visibility have been ruled out by speckle or adaptive optics observations (T Tau and SU Aur,
Ghez et al.
(1993); MWC147, Corporon (1998)).
3.3. Accretion disk
Emission from an accretion disk is one of the leading explanations for the infrared excess and
other observed features of T Tauri stars and has also been proposed for Herbig Ae/Be stars. One
common method of describing the physical properties of the accretion disk is to parameterize the
temperature (T ) and surface density (Σ) as power-law functions of the radius (T ∝ r−q, Σ ∝ r−p)
and the dust opacity as a power-law function of wavelength (κ ∝ λα). At 2.2 µm the disk is
optically thick and the emission profile at a given radius depends on the temperature distribution
(Beckwith et al. 1990). As we have only sampled one spatial scale in the disk, we will use accretion
disk models from the literature, where the SED or millimeter imaging has been used to determine
the disk parameters.
4. Results
4.1. T Tau
T Tauri, one of the best-studied YSOs, has an infrared companion, T Tau S, 0.′′7 to the
south, which is optically obscured. Both components have a near-infrared excess, suggestive of
circumstellar material. Recent observations (Koresko 2000) have revealed that T Tau S is also a
binary. The millimeter wave flux is dominated by material surrounding T Tau N and is consistent
with circumstellar disk models with an outer radius of 40 AU (Akeson et al. 1998).
At K band, T Tau N is the component with higher flux and thus, it contributes most to the
measured visibility. The binary separation is sufficiently large such that the fringe envelopes of the
two sources do not overlap and thus T Tau S does not contribute any coherent flux. However, it
is within the field of view of the star and fringe trackers and therefore contributes incoherent flux,
reducing the observed V 2 on T Tau N by R2/(1 + R)2, where R is the flux ratio between T Tau N
and S. For these observations the ratio has two components, the intrinsic K band flux ratio of the
system and an instrumental flux ratio introduced by an optical fiber, both of which are discussed
below.
– 5 –
At PTI, the angle tracking passband is 0.7-1.0 µm. T Tau N is brighter than T Tau S in the
(1998) measured a ratio of >2000 in the I band; thus the field will be
visible, Stapelfeldt et al.
centered on T Tau N. As the optical path for the spectrometer channels includes a fiber with 1′′
FWHM, the flux contribution from T Tau S is reduced. The relative coupling of the N/S flux was
calculated in the following way. An Airy pattern given by the diffraction limit of 1.′′2 was convolved
with a 1′′ Gaussian representing the seeing. We then use a matched filter of a Gaussian with 1′′
FWHM for the fiber, which is centered on T Tau N. The errors were conservatively estimated by
assuming the fiber FWHM to have an error of 50%. The resulting fiber coupling ratio (N/S) is 1.50
±0.2.
The total flux and flux ratio for the T Tau system varies on time scales of weeks to months
(1996)) and so roughly contemporaneous flux ratios are necessary. T. Beck
(e.g. Skrutskie et al.
(private communication) measured a N-S flux ratio of 1.96±0.19 on 31 Oct 1999. The last four
nights of our data were taken on 20 and 26 Oct and 2 and 3 Nov 1999. In the analysis below, we
assume the flux ratio measured by Beck is valid for these 4 nights and use only these data.
We note that we have no direct way of knowing that the detected fringes arise from T Tau
N instead of T Tau S. An alternate explanation for the detection is that the fringes arise from T
Tau S. If this were the case, the expected visibility with T Tau S unresolved is V 2=0.07, much
lower than that observed. The low visibility is due to the flux and fiber ratios still favoring T
Tau N and would be even lower if T Tau S were resolved. The stability of the fringe tracking
and the consistency of the night to night measurements strongly suggest that only 1 of the binary
components is detected, and given the above argument, we deduce that T Tau N is the detected
component.
To correct for T Tau S, the visibilities were divided by R2/(1 + R)2 before modeling was
performed, where R represents the correction for both the intrinsic flux ratio and the coupling
effect described above and is 2.94 ±0.48. For the stellar parameters estimated by Ghez et al.
(1991) using optical photometry, the total to stellar flux ratio at K is 3.6 for T Tau N. The stellar
contribution is included in the models as an unresolved component. The visibilities show a slight
dependence on hour angle, but more data are needed to confirm this effect.
The data were reduced and binned by projected baseline before being fit by Gaussian and
uniform profile models. The best fit uniform profile diameter is 2.62+0.046
−0.044 mas (0.37 AU) and
the best fit Gaussian has a FWHM of 1.61+0.028
−0.031 mas (0.22 AU) (Figure 1a). For the accretion
disk model, we have used the parameters derived by Ghez et al. (1991) with rinner = 0.04 AU,
router = 100 AU, a temperature profile T ∝ r−0.42 and T (1 AU) = 260 K. This accretion disk
model overestimates the measured visibility and, thus, underestimates the size scale. Using the
disk parameters derived by Akeson et al. (1998) from millimeter wave emission (rinner = 0.01 AU,
router = 40 AU, T ∝ r−0.6, T (1 AU) = 100 K), the predicted size is even smaller than the Ghez
model.
The binary parameters which fit the measured visibilities are shown in Figure 1b, where the
– 6 –
contours are for χ2
r of 1, 2 and 4. We note that if a binary companion at the separation shown in
Figure 1b were in a roughly circular orbit around T Tau N, we should have seen visibility changes
in the data given the time span covered by the observations. However, no time dependence was
seen in the data.
1
0.8
0.6
0.4
0.2
2
V
1.61 mas Gaussian
2.62 mas uniform disk
Accretion disk (Ghez)
Accretion disk (Akeson)
Binary companion
0
0
20
40
80
Projected baseline (m)
60
100
120
o
i
t
a
r
x
u
F
l
8
7
6
5
4
3
2
1
1
2
3
4
5
6
7
8
Binary separation (mas)
Fig. 1.- Binned data and models for T Tau. The models are Gaussian profile (solid line), uniform
profile (dotted line), accretion disk from Ghez et al. (1991) (dash-dot line), accretion disk from
Akeson et al. (1998) (thick solid line at top) and binary companion (dashed line). The plotted
visibilities have been corrected for the incoherent flux of T Tau S, as described in §4.1. The binary
parameters represented are separation of 1.5 mas and a flux ratio of 4.3. b) Contour plot of possible
binary companion parameters. The contour levels represent models with a χ2
r of 1, 2, and 4.
4.2. SU Aur
SU Aur is a T Tauri star with an SED similar to that of T Tau. Herbig and Bell
(1988)
designated SU Aur as the prototype of a separate classification from weak-line T Tauris due to
its broad absorption lines and high luminosity (∼ 12 L⊙). The stellar/total flux ratio at K is 0.3
(Marsh and Mahoney 1992). The best fit diameters are 1.92+0.063
−0.039 mas for a
uniform profile and a Gaussian (FWHM) respectively (Figure 2a). This corresponds to a physical
size of 0.27 and 0.16 AU for a distance of 140 pc. The accretion disk model is taken from Beckwith
et al. (1990) with rinner = 0.01 AU, router = 100 AU, a temperature profile T ∝ r−0.51 and T (1 AU)
= 260 K, where the parameters were determined by fitting 10 µm through millimeter wave fluxes.
This model underestimates the observed visibility.
−0.059 mas and 1.16+0.038
The binary parameter space for models with reduced chi-squared, χ2
r of 1, 2 and 4 is shown
– 7 –
in Figure 2b. The pattern shown in the figure repeats in separation space to roughly 20 mas.
The binary companion hypothesis for SU Aur is not well constrained due to the limited time and
baseline coverage of the data. The time coverage we do have favors orbiting companions with
separations at roughly 2 or 6 mas.
1
0.8
0.6
0.4
0.2
2
V
1.16 mas Gaussian
1.92 mas uniform disk
Accretion disk
Binary companion
o
i
t
a
r
x
u
F
l
14
12
10
8
6
4
2
0
0
20
40
80
Projected baseline (m)
60
100
120
1
2
3
4
5
6
7
8
Binary separation (mas)
Fig. 2.- a) Binned data and models for SU Aur. The models are Gaussian profile (solid line),
uniform profile (dotted line), accretion disk (dashed line) and binary companion (dash-dot line).
The binary parameters represented are separation of 2 mas and a flux ratio of 10. b) Contour plot
of possible binary companion parameters. Note that this pattern repeats in separation space to 20
mas. The contour levels represent models with a χ2
r of 1, 2, and 4.
4.3. MWC 147
MWC 147 (HD 259431) is a Herbig Ae/Be star with spectral classifications in the literature
ranging from B2 to B6. Hillenbrand et al. (1992) modeled the SED from this source as arising from
a flat, optically thick disk with an inner hole. Previous studies have used a distance to this source
of 800 pc, which we will use here for consistency, although we note a recent distance determination
from Hipparcos data of 290+200
−84 pc (Bertout et al. 1999). If the distance to MWC 147 is ∼290 pc,
the physical sizes given below decrease by a factor of 2.8. Depending upon which spectral type is
used, the stellar contribution to the flux at K is 0.05 to 0.1 of the total. We use an unresolved
component with 0.1 of the total flux to represent the central star in these models. Using a stellar
contribution of 0.05 would increase the squared visibility due to the disk by 4%.
The data were reduced as described above and binned by projected baseline. The data and
models are shown in Figure 3a. Given the errors on the individual data points, there is no significant
– 8 –
dependence on hour angle in the visibility data, consistent with Millan-Gabet (1999b). The best
fit uniform profile diameter is 2.28+0.017
−0.034 mas (1.8 AU) and the best fit Gaussian has a FWHM of
1.38+0.013
−0.014 mas (1.1 AU). The accretion disk and stellar parameters were taken from Hillenbrand
et al. (1992) with rinner = 0.36 AU, router = 1.8 AU and a temperature profile T ∝ r−3/4. The
M = 10−5
reference temperature is set by the stellar temperature, 2×104 K, and the accretion rate,
M⊙/year. As seen in Figure 3a, this accretion disk model underestimates the measured visibility
and so overestimates the physical size.
The observed visibilities for MWC 147 can also be explained by a binary companion. Figure
3b shows the parameter space for binary models with reduced χ2
r of 1, 2, and 4. An adaptive optics
survey by Corporon (1998) found a binary companion to MWC 147 with a separation of 3.′′1 and
magnitude difference ∆K = 5.7. This source is too widely separated and too faint to have affected
our observations.
1
0.8
0.6
0.4
0.2
2
V
1.38 mas Gaussian
2.28 mas uniform disk
Accretion disk
Binary companion
9
8
7
6
5
4
3
2
o
i
t
a
r
x
u
F
l
0
0
20
40
80
Projected baseline (m)
60
100
120
1
0.5
1
1.5
2
3
Binary separation (mas)
2.5
3.5
4
Fig. 3.- a) Binned data and models for MWC 147. The models are Gaussian disk (solid line),
uniform disk (dotted line), accretion disk (dashed line) and binary companion (dash-dot line). The
binary parameters represented are a separation of 2.1 mas and a flux ratio of 6. b) Contour plot of
possible binary companion parameters. The contour levels represent models with a χ2
r of 1, 2, and
4.
4.4. AB Aur
AB Aur is a Herbig Ae/Be star with a spectral type of A0, at a distance of 140 pc. Millan-
Gabet et al. (1999) resolved the infrared emission from this source with the IOTA interferometer
using baselines of 21 and 38 m. At PTI, fringes were not detected on AB Aur, despite a photon
– 9 –
flux higher than that for T Tau, indicating that the AB Aur is too large to be detected on baselines
of ∼100 m with current sensitivities. Upper limits were found for the visibility using the sensitivity
of the detection algorithm and measuring the system visibility with a calibrator. At K band, the
estimated upper limit was hV 2i < 0.08 ± 0.02, which corresponds to a size >4.1(±0.2) mas (0.57
AU) diameter for a uniform profile and > 2.7(±0.1) mas (0.38 AU) for a Gaussian. These results
are consistent with the size and derived by Millan-Gabet et al. (1999) and a previous upper limit
from PTI of hV 2i < 0.3 from Berger (1998).
5. Discussion
The fundamental result of these observations is that for all four sources observed here, the
infrared emission arising from circumstellar material is resolved by PTI with a nominal fringe
spacing of 4 mas. The measured sizes correspond to physical scales of tenths of AU for the T Tauri
sources T Tau and SU Aur and ∼1 AU for the Herbig Ae/Be star MWC 147. We note that if the
correct distance for MWC 147 is 290 pc, rather than 800 pc, then the measured visibility corresponds
to a size scale of roughly 0.5 AU, similar to that measured for the T Tauri sources, despite the
large difference in stellar mass. Our measured visibilities do not agree with those predicted from
accretion disk models derived from near-infrared SEDs or millimeter interferometric observations.
This may suggest that the single power-law relations used to describe the temperature and density
are inadequate to reproduce both the spectral and spatial characteristics of the emission.
Our data on T Tau and SU Aur require the K band emission to come from a larger region than
that predicted by the accretion disk models, while the opposite is true for MWC 147. Millan-Gabet
(1999b) observed 15 Herbig Ae/Bes using infrared interferometry with a shorter baseline and found
that roughly half of the sources could not be well modeled as emission from an accretion disk and
that the predicted visibilities were higher than the observed data. On the other hand, Malbet et al.
(1998) used PTI for observations of FU Ori and found that an accretion disk model could explain
the measured visibilities.
Further characterization of the circumstellar material on size scales less than one AU can be
achieved by extending the infrared interferometry observations presented here. PTI has a second
baseline, which provides data on shorter spacings, and is equipped to observe in the H band, which
probes higher temperatures and has higher spatial resolution than K band. We plan to extend our
study of young stellar objects to include H band and more spatial scales.
This work was performed at the Infrared Processing and Analysis Center, Caltech and the
Jet Propulsion Laboratory. Data were obtained at the Palomar Observatory using the NASA
Palomar Testbed Interferometer, which is supported by NASA contracts to the Jet Propulsion
Laboratory. Science operations with PTI are possible through the efforts of the PTI Collaboration
(http://huey.jpl.nasa.gov/palomar/ptimembers.html). We particularly thank A. Boden for
– 10 –
his efforts in data reduction software and B. Thompson for useful discussions. We are also grateful
to T. Beck for providing the T Tau flux ratio. DRC acknowledges support from NASA WIRE ADP
NAG5-6751. EAL acknowledges support from a Research Corporation Innovation Award and a
Presidential Early Career Award for Scientists and Engineers to the University of Florida.
REFERENCES
Akeson, R. L., Koerner, D. W., and Jensen, E. L. N. 1998, ApJ, 505, 358
Beckwith, S. V. W., Sargent, A. I., Chini, R. S., and Gusten, R. 1990, AJ, 99, 924
Berger, J.P. 1998, Ph.D. thesis, Univ. Joseph Fourier de Grenbole, France
Bertout, C., Robichon, N., and Arenou, F. 1999, A&A, 352, 574
Boden, A. F., Colavita, M. M., van Belle, G. T., and Shao, M. 1998, SPIE proceedings, 3350, 872
Colavita M. M. et al 1999, ApJ, 510, 505
Corporon, P. 1998, Ph.D. thesis, Univ. Joseph Fourier de Grenbole, France
Ghez, A. M., Neugebauer, G., Gorham, P. W., Haniff, C. A., Kulkarni, S. R., Matthews, K.,
Koresko, C., and Beckwith, S. 1991, AJ, 102, 2066
Ghez, A. M., Neugebauer, G., and Matthews, K. 1993, AJ, 106, 2005
Herbig, G.H. and Bell, K.R. 1988, Lick Obs. Bull. 111, Third Catalog of Emission-line Stars for
the Orion Population (Santa Cruz: Univ. California)
Hillenbrand, L. A., Strom, S. E., Vrba. F. J., and Keene, J. 1992, ApJ, 397, 613
Koresko, C.D., 2000, ApJ, 531, 147
Malbet, F., et al 1998, ApJ, 507, 149
Malbet, F. and Bertout, C. 1995, A&AS, 113, 369
Marsh, K. A. and Mahoney, M. J. 1992 ApJ, 395, 115
Millan-Gabet, R., Schloerb, F. P., Traub, W. A., Malbet, F., Berger, J. P., and Bregman, J. D.
1999, ApJ, 513, 131
Millan-Gabet, R. 1999, Ph.D. thesis, University of Massachusetts
Mundy, L.G., Looney, L.W. and Welch, W.J. 2000, in Prototstars and Planet IV, ed. V. Mannings,
A.P. Boss and S.S. Russell (Tucson: University of Arizona), in press
– 11 –
Natta, A., Grinin, V.P., and Mannings, V. 2000, in Prototstars and Planet IV, ed. V. Mannings,
A.P. Boss and S.S. Russell (Tucson: University of Arizona), in press
Skrutskie, M. F., Meyer, M. R., Whalen, D. & Hamilton, C. 1996, AJ, 112, 2168
Stapelfeldt, K. R. et al. 1998, ApJ, 508, 736
Wilner,D.J. and Lay,O.P 2000, in Protostars and Planets IV, ed. V. Mannings, A.P. Boss and S.S.
Russell (Tucson: University of Arizona Press), in press
This preprint was prepared with the AAS LATEX macros v5.0.
Table 1. PTI observations
Source
# of nights
Calibrators
(records)
< V 2 >
Uniform diska
(diameter)
Gaussiana
(FWHM)
T Tau
SU Aur
MWC 147
AB Aur
4(119)
HD 28024, HD 27946
3(38) HD 28024, HD 27946, HD 25867
HD 43042, HD 46709
HD 32301
9(121)
4c
0.29 ± 0.01b 2.62+0.046
1.92+0.063
0.68 ± 0.02
2.28+0.017
0.53 ± 0.01
< 0.15
−0.044 mas
−0.059 mas
−0.034 mas
>3.7 mas
1.61+0.028
1.16+0.038
1.38+0.013
−0.031 mas
−0.039 mas
−0.014 mas
>2.4 mas
a1 σ uncertainties are given for the best fit Gaussian and uniform disk models.
bThe visibility given for T Tau is the calibrated value uncorrected for the effects of T Tau S.
cOnly nights with good upper limits are listed for AB Aur
|
astro-ph/9610103 | 1 | 9610 | 1996-10-15T01:25:50 | Synthesis Imaging of Dense Gas in Nearby Galaxies | [
"astro-ph"
] | We present BIMA observations of the HCN emission from five nearby spiral galaxies. The HCN observations comprise the first high-resolution (5--10\arcsec) survey of dense molecular gas from a sample of normal galaxies, rather than galaxies with prolific starburst or nuclear activities. The images show compact structure, demonstrating that the dense gas emission is largely confined to the central kiloparsec of the sources. In one of the galaxies, NGC 6946, the ratio of HCN to CO integrated intensities ranges from 0.05--0.2 within the extent of the HCN emission (r = 150 pc), with an average value of 0.11 \pm 0.01 over the whole region; the range and average values of the ratios in NGC 6946 are very similar to what is observed in the central r = 250 pc of the Milky Way. In NGC 6946, NGC 1068 and the Milky Way, the ratio I{HCN}/I{CO} is 5 to 10 times higher in the bulge regions than in their disks; this suggests that the physical conditions in their bulges and disks are very different. In NGC 4826 and M51, as in the Milky Way and NGC 1068, there is a linear offset of \sim 100 pc between the dense gas distribution and the peak of the radio continuum emission. | astro-ph | astro-ph |
To appear in The Astrophysical Journal
Synthesis Imaging of Dense Gas in Nearby Galaxies
Department of Astronomy, University of Maryland, College Park, MD 20742
Tamara T. Helfer1 and Leo Blitz1
ABSTRACT
We present images of the HCN J = 1-0 emission from five nearby spiral
galaxies made with the Berkeley-Illinois-Maryland Association interferometer.
The HCN observations comprise the first high-resolution (θ ∼ 5′′ -- 10′′) survey
of dense molecular gas from a sample of normal galaxies, rather than galaxies
with prolific starburst or nuclear activities. The images show compact structure,
demonstrating that the dense gas emission is largely confined to the central
kiloparsec of the sources. To within the uncertainties, 70 - 100% of the
single-dish flux is recovered for each source; this implies that there is not a
significant contribution to the HCN flux from low-level emission in the disks
of the galaxies. In one of the galaxies, NGC 6946, the ratio of HCN to CO
integrated intensities ranges from 0.05 -- 0.2 within the extent of the HCN
emission (r = 150 pc), with an average value of 0.11 ± 0.01 over the whole
region; the range and average values of the ratios in NGC 6946 are very similar
to what is observed in the central r = 250 pc of the Milky Way. A comparison
with single-dish observations allows us to place an upper limit of 0.01 on the
ratio of integrated intensities in the region 150 < r < 800 pc in NGC 6946.
In NGC 6946, NGC 1068 and the Milky Way, the ratio IHCN/ICO is 5 to 10
times higher in the bulge regions than in their disks; this suggests that the
physical conditions in their bulges and disks are very different. Furthermore, the
presence of dense gas on size scales of ∼ 500 pc in the centers of these nearby
galaxies and the Milky Way suggests that the internal pressure is at least 107
cm−3 K in their centers; this is some three orders of magnitude greater than
the pressure in the local interstellar medium in the Milky Way, and it is two
orders of magnitude greater than the pressure from the self-gravity of a solar
neighborhood giant molecular cloud. In NGC 4826 and M51, as in the Milky
1thelfer,[email protected], current address Radio Astronomy Lab, 601 Campbell Hall, UC Berkeley,
Berkeley CA 94720
-- 2 --
Way and NGC 1068, there is a linear offset of ∼ 100 pc between the dense gas
distribution and the peak of the radio continuum emission. We did not detect
HCN towards three additional spiral galaxies.
Subject headings: galaxies:individual (NGC 3628, NGC 4826, NGC 5194
(M51), NGC 5236 (M83), NGC 6946) -- galaxies:ISM -- galaxies:nuclei --
interstellar:molecules
1.
Introduction
Single dish millimeter observations of HCN and CS emission from nearby galaxies
show that most spiral galaxies, not just starburst galaxies, have an appreciable amount of
dense (∼ 105 cm−3) gas in their bulges (Helfer & Blitz 1993 and references therein). These
observations are in good agreement with what is seen in the inner r ∼ 250 parsecs of the
Milky Way, where strong, diffuse emission from CS (Bally et al. 1987) and HCN (Jackson
et al. 1996) is observed despite the moderate star formation rates there. (HCN and CS
trace gas densities of ∼> 105 cm−3 in galaxies; this is two orders of magnitude higher than
the density required to excite CO.) The ubiquity of large-scale emission from HCN and
CS over hundreds of parsecs in the centers of galaxies is surprising when compared with
the known properties of giant molecular clouds (GMCs) in the solar neighborhood, which
contain few and relatively small (∼< pc-scale) clumps of dense gas -- and these only where
stars are actively forming or where stars have formed very recently (i.e. where there is a
local source of pressure, not common to the GMC as a whole).
In galaxies, single-dish millimeter wavelength beams typically cover tens to hundreds
of GMC-sized diameters. As part of our program to measure the degree to which different
galactic environments affect the properties of molecular clouds, we began a program with
the Berkeley-Illinois-Maryland Association (BIMA) interferometer to survey eight nearby
galaxies in HCN emission in order to measure the distribution and amount of dense gas on
size scales of individual GMCs or small associations of GMCs. We consider these galaxies
to be "normal" since they do not have prolific circumnuclear starbursts or active galactic
nuclei; however, like most spiral galaxies, these galaxies have low-level nuclear line emission
(Ho, Filippenko & Sargent 1993). In this paper, we present the results of the BIMA survey.
We also present a comparison of the HCN emission from one of the sources, NGC 6946,
with its CO J = 1-0 emission (Regan & Vogel 1995). A comparison with the CO emission
from the other sources will be the topic of a future paper.
-- 3 --
2. Observations
2.1. BIMA Observations
The sources and their coordinates are listed in Table 1. The sources were selected from
the single-dish survey of Helfer & Blitz (1993) as galaxies with strong CO emission that
were also detected in HCN with a single pointing using the NRAO 12 m telescope. The data
were collected with the BIMA interferometer (Welch et al. 1996), which then comprised 6
antennas, between 1994 February 15 and 1995 May 02; the observations included data from
up to three array configurations. For each source, the receivers were tuned to the redshifted
frequency of the HCN J = 1-0 transition (ν o = 88.61 GHz). The data were processed
using the MIRIAD package (Sault, Teuben, & Wright 1995). The time variations of the
amplitude and phase gains were calibrated using short observations of nearby quasars every
∼ 30 minutes; a planet or strong quasar was observed to set the absolute flux scale as well
as to calibrate the spectral dependence of the gains across the IF passband. The digital
correlator was configured to achieve a maximum spectral resolution of 1.56 MHz (5.3 km
s−1). For some of the observations, there was an intermittent phase lock on one of the
antenna receivers that was not discovered until after the observations; in these cases all
baselines involving that antenna were eliminated from further data reduction.
For each source, the calibrated data were smoothed to 21.1 km s−1 resolution, then
gridded and Fourier transformed using natural weighting. The visibilities were also weighted
by the inverse of the noise variance, which was determined from the system temperatures
and from the gains of the individual antennas. We cleaned the maps using the standard
Hogbom algorithm. Maps of integrated intensity were made by summing those channels
with emission after clipping the channel maps at a 1 σ level. The beam sizes and noise
levels of the final maps are listed in Table 2. The noise levels are somewhat underestimated
relative to the formal uncertainties, since they were determined from the clipped moment
maps. The absolute flux calibration in all maps is probably accurate to ± 30%.
2.2.
Single-dish Observations of CO in NGC 6946
We have also used BIMA to image CO in the centers of three of the galaxies (NGC
3628, NGC 4826, and M83) with detected HCN emission. A fourth galaxy, NGC 6946, has
been imaged in CO using BIMA by Regan & Vogel (1995). The CO emission, unlike the
HCN emission (see below), is significantly "resolved out" in the interferometric images of
these galaxies; it is therefore necessary to fill in the zero-spacing flux with a single-dish
telescope. We are currently still making these measurements for the first three sources; we
-- 4 --
therefore defer further discussion of the CO results for these galaxies to a future paper.
In the case of NGC 6946, we measured the CO short-spacing flux with the NRAO 12
m telescope2 and combined these data with Regan & Vogel's BIMA map. We observed on
95 June 20 -- 21 in the newly implemented "on-the-fly" (OTF) scheme at the NRAO 12 m
(Emerson et al., in preparation) and covered a region roughly 5′ on a side. We observed
orthogonal polarizations using two 256 channel filterbanks, each with a spectral resolution
of 2 MHz per channel. The data were gridded and a linear baseline removed from the
resulting data cubes in AIPS; the data were then transferred to the MIRIAD package for
further processing. Details of the single dish and interferometric data combination were
very similar to those described in Helfer & Blitz (1995) for the case of NGC 1068.
3. Results
We detected HCN in five of the eight galaxies: NGC 3628, NGC 4826, NGC 5194
(M51), NGC 5236 (M83), and NGC 6946; their images are presented in Figure 1 along
with optical images from the Digitized Sky Survey.3 The spectra from the positions of peak
HCN emission in each of the detected sources are shown in Figure 2.
In each case, the HCN images show compact structure, with the detected emission
confined to the central ∼ 500 pc diameter; the 5′′ -- 10′′ FWHM synthesized beam sizes
correspond to linear resolutions of 125 -- 200 pc at the distances of the galaxies. While the
HCN emission is compact, it appears resolved by the interferometer in each map. In order
to investigate the effect of the lack of uv sampling at small spatial visibilities, we compared
the integrated intensities measured with the interferometer to the single-dish fluxes of
Helfer & Blitz (1993). Table 3 lists the integrated intensities measured at BIMA along with
those measured at the NRAO 12 m and the fraction of the single-dish flux recovered by the
interferometer. With the possible exception of M51 (see § 3.1), the interferometer appears
to recover all the single-dish flux measured by Helfer & Blitz. This means that the maps
shown in Figure 1 are a reliable representation of the distribution of the HCN emission,
2The National Radio Astronomy Observatory is operated by Associated Universities, Inc., under
cooperative agreement with the National Science Foundation.
3Based on photographic data of the National Geographic Society -- Palomar Observatory Sky Survey
(NGS-POSS) obtained using the Oschin Telescope on Palomar Mountain. The NGS-POSS was funded by
a grant from the National Geographic Society to the California Institute of Technology. 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.
-- 5 --
constrained of course by the usual limitation of the signal to noise ratios. Allowing for
low-level, more diffuse emission detected at < 2 σ, we conclude that the HCN emission is
confined to the central kiloparsec of the galaxies (it is perhaps slightly more extended in
M51). The spatial extent of the HCN emission is in good agreement with what is seen in
the Milky Way (Jackson et al. 1996) and in NGC 1068 (Helfer & Blitz 1995).
3.1.
Individual Sources
The galaxies in this survey were originally selected from a list of the brightest
extragalactic CO emitters (see Helfer & Blitz 1993). The following are brief descriptions of
the individual sources.
NGC 3628 -- This galaxy is a member of the Leo triplet (along with NGC 3623 and
NGC 3627). It is a nearly edge-on Sbc galaxy with a prominent and irregular dust lane.
Its nuclear region contains a modest starburst (e.g. Condon et al. 1982, Braine & Combes
1992). The CO from NGC 3628 (Young, Tacconi, & Scoville 1983; Boiss´e, Casoli, & Combes
1987; and Israel, Baas, & Maloney 1990) is strongly peaked in the inner kiloparsec and has
a similar extent and mass as that in the center of the Milky Way (Boiss´e et al. 1987). The
HCN in NGC 3628 is resolved and appears elongated in the east-west direction. There is
emission at the 2 σ level to the northwest of the central source that may be associated with
gas further along the major axis. The HCN, CO and radio continuum centers of NGC 3628
all peak some 21′′ to the southeast of the optical nucleus; however, the optical nucleus is
heavily obscured and its position is highly uncertain (Boiss´e et al. 1987).
NGC 4826 -- This Sab galaxy has been called variously the "Black Eye," the "Evil
Eye," or somewhat more optimistically (Rubin 1994) the "Sleeping Beauty" galaxy for its
conspicuous dust lane (see Sandage 1961; see also the cover of The Astronomical Journal,
1994, 107, 1). NGC 4826 has gained notoriety recently for the discovery that the inner
kiloparsec-scale disk is counterrotating with respect to the larger scale rotation of the
galaxy (Braun, Walterbos, & Kennicutt 1992; Rubin 1994; Braun et al. 1994). The CO
emission (Casoli & Gerin 1993) is confined to the inner r ∼ 1′. The HCN emission appears
symmetric and centrally peaked in NGC 4826.
NGC 5194 (M51) -- M51 is the prototypical grand design spiral and is one of the best
studied galaxies in CO emission. Interferometric CO images of the nuclear region of M51
(e.g. Lo et al. 1987; Rand & Kulkarni 1990; and Adler et al. 1992; the last includes
zero-spacing flux) show a notable lack of a single central concentration of CO, despite the
strong molecular emission associated with the spiral arms in the nuclear region. M51 is
-- 6 --
the only galaxy of the five detected in HCN at BIMA that does not have a strong central
concentration of CO. In contrast to the CO emission, the HCN emission does appear
to be centrally concentrated, though there is a significant contribution to the total flux
from low-level (< 2 σ) emission. We note that Kohno et al. (1996) mapped the central
concentration as well as more extended structure in their HCN map of M51 observed with
the Nobeyama Millimeter Array. The linear extent of the low-level (< 2 σ) HCN emission
in the BIMA map is somewhat larger than those of the other four galaxies studied here,
and the interferometer may have "resolved out" a nonnegligible contribution to the total
flux in this source (Table 3). For M51, the flux of structures larger than ∼ 18′′ is attenuated
by ∼> 50%.
NGC 5236 (M83) -- This well-studied source is a grand design, Sc/SBb spiral galaxy
with strong circumnuclear star formation within the central few hundred pc (e.g. Gallais et
al. 1991). The HCN emission is strongest at the position of the peak of the radio continuum
emission (Condon et al. 1982; Turner & Ho 1994); the condensations to the east and to
the south of the strongest HCN emission are also apparent in the radio continuum. The
synthesized beam of the interferometer is quite elongated because of the foreshortening of
the north-south baselines toward this low declination source. The north-south elongation of
the HCN emission is therefore almost certainly an artifact of the observations.
NGC 6946 -- This late-type, grand design spiral galaxy contains a moderate starburst
(Turner & Ho 1983). Within the inner 1.5 kpc diameter, the CO in this galaxy has a
non-axisymmetric, north-south distribution that has been interpreted as a bar (Regan &
Vogel 1995 and references therein). However, Regan & Vogel (1995) combined new CO and
K-band observations and showed that the CO traces gas on the trailing side of spiral arms;
their observations are consistent with what is expected for the gas and stellar response
to a spiral density wave rather than a bar. The CO peaks up strongly in the inner 300
pc diameter of NGC 6946. The HCN emission is detected in this region and is resolved
and slightly extended in an east-west direction, with an additional elongation towards the
northwest (there is a similar northwest extension in the CO map of Regan & Vogel 1995).
We discuss this source more fully in the following section.
3.2.
IHCN/ICO in NGC 6946
The ratio of the 3 mm integrated intensities, IHCN/ICO, may be used as a qualitative
measure of the molecular gas density (e.g. Helfer & Blitz 1996). In order to determine any
line ratio from interferometric measurements, one must first take into account the possibility
that the flux measured with the interferometer is missing a significant contribution from
-- 7 --
large-scale structures in the maps. In NGC 6946, the interferometer recovers all the
single-dish HCN flux to within the errors of the measurement (Table 3). For the CO, the
Regan & Vogel (1995) BIMA map recovered about half the single-dish flux; we therefore
modeled the short spatial frequency visibilities from the NRAO 12 m data (§ 2.2) and
combined these with the BIMA CO map. The CO distribution and flux in the resulting
map did not change appreciably interior to r = 15′′; at larger radii, the most dramatic
flux increases were distributed over radii from 20′′ -- 50′′, though the shape of the structures
remained about the same. With the fully-sampled CO map, we can now make a legitimate
comparison of the HCN and CO intensities for this source.
To determine the ratio IHCN/ICO in NGC 6946, we convolved the CO map to match the
resolution of the HCN image, and we converted both intensities to a main beam brightness
scale (R TMB ∆v); the ratio was computed only for regions with detected HCN emission
(IHCN > 2.5 σmom, or r ∼< 15′′). The resulting ratio map of IHCN/ICO is presented in Figure
3. Although the emissions from CO and HCN both peak at the center of NGC 6946, the
distribution of the HCN/CO ratio is saddle-shaped, with the highest values to the east
and west of the nucleus by about 7′′ (175 pc) and the lowest values to the northwest and
southeast of the nucleus by about 5′′ (125 pc). At first, it seems surprising that the ratio
does not rise monotonically to the central position. However, at the small size scales
resolved by the interferometer (r ≈ 70 pc), the characteristics of individual molecular clouds
start to dominate the distribution, rather than the integrated effects of dozens of GMCs.
It may be that the small-scale ratio is dominated by local effects from individual clouds;
this effect is seen within the central few hundred pc of the Milky Way, where the ICS/ICO
ratio (Helfer & Blitz 1993) and the IHCN/ICO ratio (Jackson et al. 1996) look very clumpy
and irregularly distributed. In the Milky Way, features like the Sgr A and B GMCs are
characterized by relatively high values in the ratio maps.
The average ratio over the extent of the HCN emission (∼ 12′′, or 300 pc diameter) in
NGC 6946 is 0.11 ± 0.01, with a peak value of 0.19 and a minimum of 0.049. The range
of IHCN/ICO in NGC 6946 is very similar to what is seen in the Milky Way by Jackson et
al. (1996); on small scales within the inner few degrees of the Milky Way, IHCN/ICO ranges
from 0.04 to 0.12, and the average over the extent of the HCN emission, or r ≈ 300 pc, is
0.08 (see below).
We can compare the ratio we measure in the central 300 pc of NGC 6946 with those
measured with larger apertures: the single dish ratios of IHCN/ICO in NGC 6946 are 0.063
± 0.007 at a resolution of 24′′ (600 pc) (Nguyen-Q-Rieu et al. 1989; Weliachew, Casoli, &
Combes 1988) and 0.025 ± 0.003 at a resolution of ∼ 1′ (1500 pc) (Helfer & Blitz 1993).
Figure 4a shows this radial distribution of the average integrated IHCN/ICO ratio. It is
-- 8 --
important to note that the points shown in Figure 4a represent the average ratios over the
area enclosed at the radius r, i.e. that the plot represents the integrated ratios IHCN/ICO
as a function of r. The monotonic falloff in the ratio with radius is simply a result of the
confinement of the HCN emission to the inner r = 150 pc, while the CO is distributed over
a much larger radius (there is detectable emission at least to r = 3.5 kpc, Tacconi & Young
1989). What is perhaps a more interesting quantity physically is the annular ratio measured
as a function of radius; that is, if IHCN/ICO is 0.10 ± 0.01 measured as an average from
0 < r < 150 pc, what is the value of IHCN/ICO from 150 < r < 800 pc (where 800 pc is the
radius of the NRAO beam in Helfer & Blitz 1993)? We can set an upper limit to IHCN/ICO
in this annulus by comparing the BIMA data with the NRAO 12 m HCN flux. Since the
BIMA measurement recovered 0.81 ± 0.17 of the single-dish HCN flux measured at the
NRAO 12 m (Table 3), let us assume that 20% of the single-dish HCN flux is distributed at
radii larger than the interferometer was able to measure, yet within the half power beam
area of the the NRAO 12 m -- that is, radii within the annulus 150 < r < 800 pc. (This
is a conservative estimate, since any "missing" large-scale flux could also contribute to the
flux at the central position.) If we then measure the flux in the Regan & Vogel CO map
from 150 < r < 800 pc, we find that the ratio IHCN/ICO in this annulus is at most 0.01.
This ratio is an order of magnitude lower than that measured over the central r = 150 pc,
as shown in Figure 4b.
3.3. Nondetections
We did not detect HCN emission from NGC 4321 (M100), NGC 4527, or NGC 4569.
While it is possible that the dense structure in these galaxies is so extended that the
interferometer resolves out the single-dish HCN emission (for these sources, the flux of
structures larger than ∼ 20′′ is attenuated by ∼> 50%), it is also likely that the observations
simply were not sensitive enough to detect the HCN from these sources. For those
observations which suffered from an intermittent phase lock (see § 2.1), the antenna that
was flagged was one of the two that made up the shortest baseline pair; thus the calibration
solution may not have been reliable and also the zero spacing problem may have been
exacerbated for these observations.
4. Discussion
-- 9 --
4.1. The Radial Dependence of Dense Gas Ratios
Spectroscopic studies of CS and HCN emission in normal external galaxies suggest
that most spiral galaxies contain an appreciable amount of gas at densities of ∼> 105 cm−3
in their centers (Mauersberger et al. 1989; Sage, Solomon, & Shore 1990; Nguyen-Q-Rieu et
al. 1992; Israel 1992; Helfer & Blitz 1993). The maps in Figure 1 show the distribution of
that dense gas, namely, that it is confined to the central kiloparsec of the sources imaged.
This situation appears to be very similar to that seen in the Milky Way, where widespread
emission from the dense gas tracers CS (Bally et al. 1987) and HCN (Jackson et al. 1996)
is found only within the central ∼ 500 pc diameter.
What does the distribution of dense gas tell us about the physical conditions in the
molecular gas as a function of its location in a galaxy? To investigate the physical conditions
rather than the total gas content, we normalize the HCN to that of the CO and consider
the ratio IHCN/ICO. If the kinetic temperature of the gas responsible for the cospatial HCN
and CO emissions is about the same, then IHCN/ICO may be considered as a qualitative
indicator of the density or thermal pressure in the gas. (Indeed, since the J = 1 state lies
only 5.5 K above the ground state for CO and 4.3 K above ground for HCN, even rather
cold gas has the energy to populate the J = 1 state for both molecules. It is the molecular
density that is more important in determining the excitation. See Helfer & Blitz 1996.) In
NGC 6946 (§ 3.2), IHCN/ICO is 0.11 ± 0.01 averaged over the central r < 150 pc, whereas
we deduce an upper limit of IHCN/ICO ≤ 0.01 averaged in the annulus 150 < r < 800 pc, a
region that includes the inner part of the disk. A similar radial dependence of IHCN/ICO is
seen in the unusual Seyfert/starburst hybrid galaxy NGC 1068, where the ratio approaches
0.6 in the central r = 175 pc (Helfer & Blitz 1995), and the ratio falls off monotonically to
about 0.1 at the large reservoir of molecular gas at about 1 kiloparsec from the nucleus.
In the Milky Way, the ratio IHCN/ICO is about 0.0814 ± 0.004 averaged over the central
r = 315 pc (Jackson et al. 1996); between 3.5 < r < 7 kpc in the plane of the Milky
Way, the average ratio is ∼ 0.026 ± 0.008, and in solar neighborhood GMCs, we measure
IHCN/ICO ratios of 0.014 ± 0.020 when averaged over ∼ 50 pc GMCs (Helfer 1995 -- and
these are upper limits to the ratio averaged over several hundred parsecs). These numbers
are summarized in Table 4.
What seems apparent from these comparisons is that IHCN/ICO is a strong function of
galactocentric radius -- or more precisely, that there is at least a bimodal distribution in the
ratio in normal galaxies: the ratio at the center is substantially higher than elsewhere in
the galaxy. Even though local effects can dominate on scales of individual GMCs (§ 3.2),
4See note c to Table 4.
-- 10 --
the average ratio of dense gas emission is highest at a galaxy's center and drops at larger
distances from the center. Furthermore, the general agreement between the ratios in the
central ∼ 500 pc of NGC 6946 and the Milky Way suggest that the physical conditions in
the centers of the two galaxies are similar.
4.2. GMCs in the High Pressure Environments of Galactic Bulges
The measurement that the ratio IHCN/ICO is 5 to 10 times higher in the bulge regions
of the Milky Way and NGC 6946 than in their disks suggests that the physical conditions of
the molecular gas in the bulge and disk regions are very different. Furthermore, the presence
of dense gas on size scales of ∼ 500 pc suggests that the internal pressure is very high in
the molecular gas. Let us assume that the intrinsic line widths of the clouds are dominated
by nonthermal, bulk motions as in local clouds, and that their intrinsic linewidths are ≥ 1
km s−1. Although we cannot model the density accurately with the observation of a single
transition of HCN or CO, a simple LVG analysis of the line ratios suggests that a line ratio
of IHCN/ICO = 0.11 implies densities of 104.2−5.2 cm−3 for gas at a kinetic temperature in the
range TK = 15 -- 70 K. These densities are consistent with what is measured in the Milky
Way bulge molecular clouds, where n(H2) ∼> 104 cm−3 (Gusten 1989). If we take the gas
densities in NGC 6946 to be 104.6 cm−3 , then the typical kinetic pressure throughout the
molecular gas is ρv2/k ∼ 1 × 107 cm−3 K.
Spergel & Blitz (1992) considered the effects of the extended, hot coronal gas in the
bulge of the Milky Way on the thin molecular layer that is embedded within it, and they
argued that the pressure in the center of the Galaxy is two to three orders of magnitude
greater than that in the solar neighborhood. How does such an extraordinary difference in
the environmental pressure affect the properties of molecular clouds? In disk GMCs, the
external pressure of the ambient ISM (∼ 104 cm−3 K) is small compared with the pressure
from the self-gravity of a cloud (∼ 105 cm−3 K), and a source of "local" pressure (i.e.,
ongoing or recent star formation) is required to support any localized high-density clumps
within the cloud. In the high-pressure (∼ 5 × 106 cm−3 K, estimated for the Milky Way
from the X-ray measurements of Yamauchi et al. 1990) environments of bulge molecular
gas, on the other hand, the GMCs need not be self-gravitating -- in fact, only the most
massive clouds could have the Jeans masses required to be self-gravitating. In bulge clouds,
then, it is the external pressure of the environment that dominates, and these high pressures
can support the dense gas throughout the molecular component (regardless of whether
there is any active star formation in the GMCs).
These arguments are easily extended to observations of external galaxies, which
-- 11 --
typically also have X-ray emission associated with their bulges (Fabbiano, Kim, & Trinchieri
1992). From our observations of HCN in the bulges of external galaxies, it appears that
the physical conditions in molecular gas in the centers of galaxies are much more similar to
each other than they are to local GMCs in the Milky Way.
4.3. Positional Offsets Between the Dense Gas and Radio Continuum
Distributions
In the Milky Way, there is a pronounced offset of ∼ 80 pc in the position of the peak
radio continuum emission (Sgr A*, l = 0◦, b = 0◦) and the centroid of the dense gas (traced
by CS and HCN) distribution (l = 0.6◦, b = 0◦). There is also evidence for an offset of ∼
100 pc in NGC 1068 between the peaks of the HCN emission and the radio continuum
emission (Helfer & Blitz 1995), and a kinematic analysis of the molecular gas suggests the
existence of an m = 1 mode (i.e. a dipole asymmetry in the mass distribution) in NGC
1068. Whether it is the radio continuum emission or the molecular gas that traces the
center of the mass distribution, the dynamical timescale of the offsets is brief (< 106 years).
If the offsets are indeed ephemeral, and not a steady-state condition of the galaxies, then it
is appropriate to look for a source of the instability that causes them.
How common are such offsets between the peaks of the dense molecular gas distribution
and the radio continuum emission in galaxies? We compared the positions of peak HCN
emission with those of the peak radio continuum in our sample (Figure 1, Table 5). While
three of the galaxies show a reasonable coincidence between the two positions, two of the
five sources, NGC 4826 and M51, show significant offsets: in NGC 4826, the offset is 3′′.7
or 74 pc; in M51, it is 2′′.6 or 130 pc. (In M83, there is a 1′′.6 or 40 pc offset, but because
of the elongated beams and extended distribution in both HCN and the radio continuum,
the peak positions are less certain.)
It appears that these offsets are a common feature in spiral galaxies. In this sample,
two out of five detected galaxies have offsets; we have already mentioned the offsets in
the Milky Way and in NGC 1068. In a recent study of the K-band morphology in 18
face-on spiral galaxies, Rix & Zaritsky (1995) found that about one third of the disks have
significant m = 1 modes at 2.5 disk exponential scale lengths.
It may be significant that the two galaxies in our sample that show offsets, NGC 4826
and M51, also share the characteristic that they have low-level nonstellar nuclear activity
(both are LINERs; these are low-level active galactic nuclei, or AGN); the other three
galaxies detected here (NGC 3628, M83, and NGC 6946) have low-level starburst activity
-- 12 --
instead (L. Ho, private communication). A recent study by Ho et al. (1993) suggests that
up to 80% of the 500 brightest galaxies in the northern sky harbor some kind of activity
in their nuclei; of these, about half are classified as LINERs, and half are starbursts. It is
surprising both that most spiral galaxies appear to show some kind of nuclear activity and
also that the activity seems to be roughly evenly divided between stellar and nonstellar
mechanisms. The offsets of the dense gas from the peak of the radio continuum in NGC
4826 and M51, as well as that in NGC 1068 (a galaxy with a more energetic AGN, but
similar conceptually to the LINERs) may help to distinguish empirically the kind of activity
that dominates in a given galaxy's nucleus.
5. Conclusions
We have presented the results of a survey of HCN emission from eight nearby spiral
galaxies made with the BIMA interferometer. These observations comprise the first
high-resolution survey of dense (∼ 105 cm−3) gas from a sample of relatively normal
galaxies, rather than galaxies with prolific starburst or nuclear activities.
We imaged five of the eight galaxies in HCN: NGC 3628, NGC 4826, NGC 5194 (M51),
NGC 5236 (M83), and NGC 6946. To within the uncertainties, the interferometer recovers
all of the single-dish flux measured for each source in a single pointing at the NRAO 12
m telescope (Helfer & Blitz 1993); this implies that there is not a significant contribution
to the HCN fluxes from extended emission in the disks of the galaxies. In all the galaxies
observed, the HCN emission is confined to the central kiloparsec of the sources.
We added zero-spacing data from the NRAO 12 m telescope to the BIMA CO map
of NGC 6946 by Regan & Vogel (1995) in order to compare the ratio of HCN to CO
intensities in this galaxy. The ratio IHCN/ICO ranges from 0.05 -- 0.2 within the central
r = 150 pc; the average ratio over this region is IHCN/ICO = 0.11 ± 0.01. A comparison with
single-dish observations allows us to place an upper limit of IHCN/ICO ≤ 0.01 in the annulus
150 < r < 800 pc in NGC 6946.
The extent of HCN emission in NGC 6946, r ∼ 150 pc, and the ratio IHCN/ICO = 0.11 in
this region are similar to what is observed in the Milky Way (r ∼ 250 pc, IHCN/ICO = 0.08,
Jackson et al. 1996); this suggests that the physical conditions in the centers of these two
galaxies are similar. Furthermore, in NGC 6946, NGC 1068, and the Milky Way, the ratios
at the centers are 5 to 10 times higher than those in the disks of these galaxies. This result
is consistent with an enhancement of two to three orders of magnitude in the pressure of
the bulges compared with the disks (Spergel & Blitz 1992). It appears that the physical
-- 13 --
regions in the centers of galaxies are much more like each other than the conditions in the
center of a galaxy relative to its disk.
In NGC 4826 and M51, as in the Milky Way and in NGC 1068, there is a linear offset of
∼ 100 pc between the dense gas distribution and the peak of the radio continuum emission.
These offsets appear to be a common feature in galaxies and may indicate that their disks
are non-axisymmetric. It may be significant that of the five galaxies imaged in HCN, NGC
4826 and M51 are LINERs, whereas the other three are starbursts.
We did not detect HCN in three galaxies with positive single-dish HCN emission: NGC
4321 (M100), NGC 4527, and NGC 4569. It could be that the emission is extended enough
in these sources that the interferometer resolved out any detectable emission; however, we
cannot rule out the possibility that there was some intrinsic problem with the observations
of these sources.
We thank the referee, Paul Ho, for his careful reading and suggestions; these helped us
to improve the manuscript. We thank Mike Regan for providing us with the BIMA NGC
6946 CO data, and we thank Darrel Emerson, Phil Jewell, Tom Folkers and the staff of
the NRAO 12 m telescope for assistance with the OTF observations and data reduction.
Luis Ho helped with the early stages of the BIMA observations. We thank Kotaro Kohno
for kindly providing us with the NRO map of HCN in M51 prior to publication. TTH
thanks Jack Welch for hospitality while visiting UC-Berkeley. This research was partially
supported by a grant from the National Science Foundation, with additional support from
the State of Maryland.
-- 14 --
REFERENCES
Adler, D.S., Lo, K.Y., Wright, M.C.H., Rydbeck, G., Plante, R.L., & Allen, R.J. 1992, ApJ,
392, 497
Bally, J., Stark, A.A., Wilson, R.W., & Henkel, C. 1987, ApJS, 65, 13
-- -- -- -- . 1988, ApJ, 324, 223
Boiss´e, P., Casoli, F., & Combes, F. 1987, A&A, 173, 229
Braine, J. & Combes, F. 1992, A&A, 264, 433
Braun, R., Walterbos, R.A.M., & Kennicutt, R.C. 1992, Nature, 360, 442
Braun, R., Walterbos, R.A.M., Kennicutt, R.C., & Tacconi, L.J. 1994, ApJ, 420, 558
Casoli, F. & Gerin, M. 1993, A&A, 279, L41
Condon, J.J., Condon, M.A., Gisler, G., & Puschell, J.J. 1982, ApJ, 252, 102
Fabbiano, G., Kim, D.-W., & Trinchieri, G. 1992, ApJS, 80, 531
Gallais, P., Rouan, D., Lacombe, F., Tiphene, D., & Vauglin, I. 1991, A&A, 243, 309
Helfer, T.T. 1995, Ph.D. Thesis, U. Maryland
Helfer, T.T. & Blitz, L. 1993, ApJ, 419, 86
-- -- -- -- -- . 1995, ApJ, 450, 90
Ho, L.C., Filippenko, A.V., & Sargent, W.L.W. 1993, in IAU Symp. 159, Multi-
Wavelength Continuum Emission of AGN, eds. A. Blecha & T.J.-L. Courvoisier
(Dordrecht:Reidel), 275
Israel, F. P. 1992, A&A, 265, 487
Israel, F.P., Baas, F., Maloney, P.R. 1990, A&A, 237, 17
Jackson, J.M., Heyer, M.H., Paglione, T.A.D., & Bolatto, A.D. 1996, ApJ, 456, 91
Kohno, K., Kawabe, R., Tosaki, T., & Okumura, S.K. 1996, ApJ, in press
Lo, K.Y., Ball, R., Masson, C.R., Phillips, T.G., Scott, S., and Woody, D.P. 1987, ApJ,
317, L63
-- 15 --
Mauersberger, R., Henkel, C., Wilson, T. L., & Harju, J. 1989, A&A, 226, L5
Nguyen-Q-Rieu, Jackson, J. M., Henkel, C., Truong-Bach, & Mauersberger, R. 1992, ApJ,
399, 521
Rand, R.J. & Kulkarni, S.R. 1990, ApJ, 349, L43
Regan, M., & Vogel, S.N. 1995, ApJ, 452, 21
Rix, H.-W. & Zaritsky, D. 1995, ApJ, 447, 82
Rubin, V.C. 1994, AJ, 107, 173
Sage, L. J., Shore, S. N., & Solomon, P. M. 1990, ApJ, 351, 422
Sandage, A. 1961, The Hubble Atlas of Galaxies (Washington: Carnegie Institute of
Washington)
Sault, R.J., Teuben, P.J., & Wright, M.C.H. 1995, in Astronomical Data Analysis Software
and Systems IV, eds. R.A. Shaw, H.E. Payne, & J.J.E. Hayes, A.S.P. Conference
Series 77, 433
Spergel, D.N. & Blitz, L. 1992, Nature, 357, 665
Tacconi, L.J. & Young, J.S., 1989, ApJS, 71, 455
Turner, J.L. & Ho, P.T.P. 1983, ApJ, 268, L79
Turner, J.L. & Ho, P.T.P. 1994, ApJ, 421, 122
Welch, W.J., et al. 1996, PASP, 108, 93
Weliachew, L., Casoli, F., & Combes, F. 1988, A&A, 199, 29
Yamauchi, S., Kawada, M., Koyama, K., Kunieda, H., & Corbet, R.H.D. 1990, ApJ, 365,
532
Young, J.S., Tacconi, L.J., & Scoville, N.Z. 1983, ApJ, 269, 136
This preprint was prepared with the AAS LATEX macros v4.0.
-- 16 --
Table 1. Sources
Source
α(J2000)
δ(J2000)
d
(Mpc)
a
vLSR
(km s−1)
NGC 3628b
NGC 4321 (M100)
NGC 4527
NGC 4569
NGC 4826
NGC 5194 (M51)
NGC 5236 (M83)
NGC 6946
11h20m16.s27
12 22 54.80
12 34 08.80
12 36 50.02
12 56 44.25
13 29 53.30
13 37 00.23
20 34 51.91
◦
′
′′
13
35
39.
0
15 49 23.0
02 39 10.3
13 09 53.1
21 40 52.3
47 11 50.0
-29 52 04.5
60 09 11.9
9
17
20
17
5
10
5
5
847
1550
1734
-235
408
463
516
52
a vLSR is defined by the radio convention, vLSR/c = ∆λ/λo, where λo is the
wavelength in the rest frame of the source.
b Pointing center is the optical center of NGC 3628. The radio continuum center
is some 21′′ to the SE of the optical center and is coincident with the HCN emission
(see § 3.1).
-- 17 --
Table 2. Mapping Details
Source
NGC 3628
NGC 4321 (M100)
NGC 4527
NGC 4569
NGC 4826
NGC 5194 (M51)
NGC 5236 (M83)
NGC 6946
Beam
(′′ × ′′)
4.7 × 4.1
8.2 × 6.5
13.5 × 10.7
5.5 × 4.2
10.5 × 8.0
7.9 × 6.5
12.5 × 4.1
5.9 × 5.0
K Jy−1
σmom
a
(Jy bm−1 km s−1)
8.13
2.91
1.07
6.85
1.87
3.03
3.03
5.31
2.0
2.1
3.6
1.0
3.2
1.5
3.6
1.7
aσmom is lower than the statistical noise in the integrated intensity ("moment")
maps because emission below the level of the statistical noise in the channel maps
was masked out in order to calculate the moment maps (see text).
-- 18 --
Table 3. Comparison with Single Dish Observations
Source
NGC 3628
NGC 4826
NGC 5194 (M51)
NGC 5236 (M83)
NGC 6946
IBIMA
a
(Jy kms−1)
49.0 ± 17.5
39.0 ± 16.7
26.4 ± 10.5
103 ± 23
39.2 ± 5.5
b
IKP
(K kms−1)
1.5 ± 0.2
1.3 ± 0.2
1.2 ± 0.2
2.7 ± 0.2
1.5 ± 0.2
Fractionc
recovered
1.01 ± 0.36
0.93 ± 0.41
0.68 ± 0.27
1.18 ± 0.27
0.81 ± 0.17
a IBIMA from this study, corrected for primary beam effects from the BIMA
interferometer and the NRAO 12 m telescope for comparison with IKP.
b IKP from Helfer & Blitz 1993, corrected to main beam brightness temperature
scale.
c Fraction recovered = (IBIMA × K Jy−1 × θ1θ2/(71′′)2) / IKP, where θ1, θ2, and
K Jy−1 are listed in Table 2.
-- 19 --
Table 4. Differential IHCN/ICO Ratios
Region
Milky Way
NGC 6946
NGC 1068a
Bulgeb
Diskd
Local GMCs
0.081c± 0.004
0.026e± 0.008
0.014 ± 0.020e
0.11 ± 0.01
≤ 0.01
--
0.6
0.1
--
aHelfer & Blitz 1995
bMilky Way: r = 300 pc; NGC 6946: r = 150 pc; NGC 1068: r = 175 pc
cJackson et al. 1995. We use their correction for HCN and CO emission at
nonzero Galactic latitude. In their "aperture photometry", Jackson et al. smooth
their HCN map to a spatial resolution larger than that of their CO map by
a factor of λ(HCN)/λ(CO) in order to emulate the single-dish measurements
of extragalactic ratios. This correction is not appropriate for comparison with
our interferometeric measurements, in which the HCN and CO maps have been
smoothed to the same resolution. We therefore "re-correct" their ratio so that it
is appropriate for direct comparison to these results.
dMilky Way: 3.5 < r < 7 kpc; NGC 6946: 150 < r < 800 pc; NGC 1068:
1.0 < r < 1.4 kpc
eHelfer 1995
-- 20 --
Table 5. Offsets Between HCN and Radio
Continuum Emission
Source
NGC 3628
NGC 4826
NGC 5194 (M51)
NGC 5236 (M83)
NGC 6946
∆α, ∆δ
(′′ × ′′)
-0.5, -0.1
+2.2, -3.0
-2.2, +1.3
-1.3, -0.9
-0.1, -0.5
References for Radio Continuum Data: NGC 3628,
NGC 5236 and NGC 6946: Condon et al. 1982; NGC
4826: Braun et al. 1994; NGC 5194: Turner & Ho
1994
-- 21 --
Fig. 1. -- BIMA images of HCN in (a) NGC 3628, (b) NGC 4826, (c) NGC 5194 (M51), (d)
NGC 5236 (M83), and (e) NGC 6946. The left panel in each plot shows a 10′ × 10′ field from
the Digitized Sky Survey around the center of each galaxy. The circle represents the BIMA
primary beam size of 132′′ FWHM, and the white square shows the field presented in the
right panel. The right panel shows the BIMA HCN image of each source. The FWHM size
of the synthesized beam is shown in the lower left corner of each image, and the horizontal
bar represents a linear size scale of 500 pc. The contour levels are ± 2,3,4... σmom, where
σmom is listed in Table 2 for each source. The white cross marks the position of the peak
radio continuum emission (see text). The images have not been corrected for primary beam
attenuation.
-- 22 --
Fig. 1. -- continued
-- 23 --
Fig. 1. -- continued
-- 24 --
Fig. 2. -- Spectra from the positions of peak HCN emission in (a) NGC 3628, (b) NGC
4826, (c) NGC 5194 (M51), (d) NGC 5236 (M83), and (e) NGC 6946. The abscissa is the
radio-defined LSR velocity in km s−1, and the ordinate is main beam brightness temperature
in K.
-- 25 --
Fig. 3. -- Ratio of HCN/CO integrated intensities in the central r = 150 pc of NGC 6946.
The halftone limits are (0.04,0.2). The lowest contour level is 0.06, and the contour interval
is 0.02. The CO map used to construct the ratio map includes zero-spacing data from the
NRAO 12 m telescope.
-- 26 --
Fig. 4. -- (a) The integral ratio of HCN/CO integrated intensities as a function of radius in
NGC 6946. The innermost point is the BIMA measurement from this study, the point at r
= 300 pc uses data from the IRAM 30 m telescope (Nguyen-Q-Rieu et al. 1992; Weliachew
et al. 1988), and the outermost point was measured using the NRAO 12 m telescope (Helfer
& Blitz 1993). Each point represents the average ratio over the area enclosed at the shown
radius. (b) The more physical differential or annular ratio as a function of radius in NGC
6946. The figure emphasizes the sharp boundary between the physical conditions in the r =
300 pc HCN-emitting region and those at larger radii.
|
astro-ph/0409587 | 1 | 0409 | 2004-09-24T14:11:30 | Detection of hard X-ray pulsations and a strong iron K_beta emission line during an extended low state of GX 1+4 | [
"astro-ph"
] | We present here results obtained from a detailed timing and spectral analysis of three BeppoSAX observations of the binary X-ray pulsar GX 1+4 carried out in August 1996, March 1997, and August 2000. In the middle of the August 2000 observation, the source was in a rare low intensity state that lasted for about 30 hours. Though the source does not show pulsations in the soft X-ray band (1.0-5.5 keV) during the extended low state, pulsations are detected in 5.5-10.0 keV energy band of the MECS detector and in hard X-ray energy bands (15-150 keV) of the PDS instrument. Comparing the 2-10 keV flux during this low state with the previously reported low states in GX 1+4, we suggest that the propeller regime in GX 1+4 occurs at a lower mass accretion rate than reported earlier. Broad-band (1.0-150 keV) pulse averaged spectroscopy reveals that the best-fit model comprises of a Comptonized continuum along with an iron K_alpha emission line. A strong iron K_beta emission line is detected for the first time in GX 1+4 during the extended low state of 2000 observation with equivalent width of ~550 eV. The optical depth and temperature of the Comptonizing plasma are found to be identical during the high and low intensity states whereas the hydrogen column density and the temperature of the seed photons are higher during the low state. We also present results from pulse phase resolved spectroscopy during the high and low flux episodes. | astro-ph | astro-ph | Accepted for publication in The Astrophysical Journal
Detection of hard X-ray pulsations and a strong iron Kβ emission
line during an extended low state of GX 1+4
4
0
0
2
p
e
S
4
2
1
v
7
8
5
9
0
4
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
S. Naik1, B. Paul2 and P. J. Callanan1
ABSTRACT
We present here results obtained from a detailed timing and spectral analysis
of three BeppoSAX observations of the binary X-ray pulsar GX 1+4 carried out
in August 1996, March 1997, and August 2000. In the middle of the August
2000 observation, the source was in a rare low intensity state that lasted for
about 30 hours. Though the source does not show pulsations in the soft X-ray
band (1.0 -- 5.5 keV) during the extended low state, pulsations are detected in
5.5 -- 10.0 keV energy band of the MECS detector and in hard X-ray energy bands
(15 -- 150 keV) of the PDS instrument. Comparing the 2 -- 10 keV flux during
this low state with the previously reported low states in GX 1+4, we suggest
that the propeller regime in GX 1+4 occurs at a lower mass accretion rate
than reported earlier. Broad-band (1.0 -- 150 keV) pulse averaged spectroscopy
reveals that the best-fit model comprises of a Comptonized continuum along
with an iron Kα emission line. A strong iron Kβ emission line is detected for
the first time in GX 1+4 during the extended low state of 2000 observation
with equivalent width of ∼ 550 eV. The optical depth and temperature of the
Comptonizing plasma are found to be identical during the high and low intensity
states whereas the hydrogen column density and the temperature of the seed
photons are higher during the low state. We also present results from pulse
phase resolved spectroscopy during the high and low flux episodes.
Subject headings: stars : neutron -- Pulsars : individual (GX 1+4) -- X-rays :
stars
1Department of Physics, University College Cork, Cork, Ireland, [email protected], [email protected]
2Tata Institute of Fundamental Research, Homi Bhabha Road, Mumbai 400 005, India, [email protected]
-- 2 --
1.
Introduction
The luminous accretion-powered X-ray pulsar GX 1+4 has several unique characteristics
which make it an ideal source for a detailed study in a wide X-ray energy band. The optical
counterpart is a M5 III giant star in a rare type of symbiotic system (Chakrabarty & Roche
1997). The neutron star in the binary system is a slow pulsar with a spin period of about 2
minutes. It is one of the brightest and hardest X-ray sources in the sky with a large rate
of change of pulse period. In 1970s, the pulsar exhibited spin-up behavior and after an
extended low intensity state in early 1980s, it showed spin-down activity (Makishima et al.
1988). BAT SE monitoring of the source, since 1991, confirmed the spin down trend with
occasional torque reversal events (Chakrabarty et al. 1997). Though the spin-up torque
is expected to be correlated with the X-ray luminosity (Ghosh & Lamb 1979), they have
been found instead to be anticorrelated for GX 1+4 (Chakrabarty et al. 1997, Paul et al.
1997a). The pulsed X-ray luminosity of GX 1+4 monitored with BAT SE, and the total
X-ray luminosity monitored with RXTE-ASM, shows strong variability over days to years
time scale, but without any periodic nature.
The pulse profile of GX 1+4 has a characteristic dip, which has been attributed due
to accretion column eclipses (Galloway et al. 2001). The dip is broad when the source is
bright and it becomes narrow with decreasing intensity. In the hard X-rays, a change in
pulse profile, from a single peak to a double peak with associated flux change was noted
in GX 1+4 with balloon borne observations (Rao et al. 1994, Paul et al. 1997b). It
occasionally shows a low state of a few hours duration (∼ 40 ks in July 1996) during which
the pulsations are absent (Cui 1997, Cui and Smith 2004). This has been interpreted as
due to centrifugal prohibition of accretion, also known as the 'propeller effect'. As GX 1+4
is spinning down in spite of a large accretion rate (the persistent X-ray luminosity is a few
times 1037 erg s−1), the magnetic field strength of the neutron star is expected to be high,
1013−14 G (Makishima et al. 1989) although this is yet to be confirmed from spectroscopic
measurements. Based on the fluctuation of the spin-up rate, Pereira et al. (1999) suggested
a binary period of 304 days.
The continuum energy spectrum of accreting pulsars is generally described by a model
consisting of a power-law, which is known to be a signature of unsaturated Comptonization
and a Gaussian function for the Iron Kα emission line feature. Iron K shell emission
lines in X-ray pulsars are believed to be produced by illumination of neutral or partially
ionized material in the accretion disk, stellar wind of the high mass companion, material
in the line of sight, or in the accretion column. The spectral fitting to the phase averaged
RXT E data, however, shows that the GX 1+4 energy spectrum is best fitted by a model
consisting of an analytic approximation to a Comptonization continuum component and
-- 3 --
a Gaussian component, attenuated by the neutral absorption column density (Galloway
2000). Pulse-phase resolved spectroscopy shows a significant increase in the value of optical
depth τ during the dips. This reveals that the dip features in the pulse profiles are due to
the eclipses of the emitting region by the accretion column (Galloway et al. 2000).
In this paper, we present the results obtained from a detailed timing and spectral
analysis of three observations of GX 1+4, made in August 1996, March 1997, and
August 2000, with the Low Energy Concentrator Spectrometer (LECS), Medium Energy
Concentrator Spectrometers (MECS), and the hard X-ray Phoswich Detector System
(PDS) instruments of BeppoSAX in 1.0 -- 150.0 keV energy band. During a major part of
the August 2000 observation, the X-ray flux was at a very low level, similar to a state
previously seen with RXT E in 1996 July, but for a longer duration of about 30 hr. The
BeppoSAX observation provides us with an opportunity to investigate the temporal and
spectral properties of GX 1+4 during this extended low state over a much wider energy
band and better energy resolution than before. In subsequent sections we give details of the
three BeppoSAX observations, the results obtained from the timing and spectral analysis,
followed by a discussion of the results.
2. Observations
The observations of GX 1+4, used for the present study, were carried out with LECS,
MECS, and PDS instruments of the BeppoSAX satellite in 1996 August, 1997 March,
and 2000 August. The details of the observations with useful exposure times are given
in Table 1. The MECS consists of three grazing incidence telescopes with imaging gas
scintillation proportional counters in their focal planes. The LECS uses an identical
concentrator system as the MECS, but utilizes an ultra-thin entrance window and a driftless
configuration to extend the low-energy response to 0.1 keV. The PDS detector system
is composed of 4 actively shielded NaI(Tl)/CsI(Na) phoswich scintillators with a total
geometric area of 795 cm2 and a field of view of 1.3o (FWHM). The LECS, MECS and
PDS are sensitive in the energy bands of 0.1-5.0, 1.0-10.0 and 15 -- 300 keV respectively. The
energy resolutions of LECS, MECS, and PDS are 25% at 0.6 keV, 8% at 6 keV and ≤ 15%
at 60 keV respectively. Time resolution of the instruments during these observations was
15.25 µs. For a detailed description of the BeppoSAX mission, refer to Boella et al. (1997)
and Frontera et al. (1997).
-- 4 --
3. Timing Analysis
GX 1+4 suffers heavily from absorption at soft X-ray energy bands by the intervening
cold material. We have, therefore, used data from MECS and PDS detectors for the timing
analysis. A barycentric correction was applied to the arrival times of the photons. It
is not possible to correct the arrival times for the the binary motion as the key orbital
parameters are not known yet. A circular region of radius 4′ around the source was selected
to extract light curves from MECS event data. Light curves in the energy bands of 1 -- 10
keV and 15 -- 150 keV were extracted from the MECS and PDS data respectively with a
time resolution of 1 s. During the 2000 observation, light curves from which are shown in
Figure 1, high count rates were detected at the beginning and at the end, separated by an
extended low intensity state of duration ∼30 hr. Near the end of this observation, there
was a flare followed by a shallow dip.
The two minute pulses of GX 1+4 were visible in the raw light curves. For accurate
measurement of the pulse period, pulse folding and a χ2 maximization method was applied
to all the light curves and pulse period was determined to be 124.404(3) s, 126.018(8) s and
134.9256(10) s during the 1996 August 18, 1997 March 25, and 2000 August 29 observations
respectively. The quoted uncertainties (1σ confidence level) in the pulse periods represent
the trial periods at which the χ2 decreases from the peak value by one standard deviation
of the chi-sq values over a wide period range far from the peak. The pulse profiles, obtained
from the MECS and PDS data of the three observations are shown in Figure 2. The pulse
phases were adjusted so as to obtain the minimum in the pulse profile at phase zero. The
characteristic narrow dips are clearly seen in the MECS pulse profiles of the 1996 and 1997
observations. The pulse fraction (defined as the ratio of the difference of maximum and
minimum flux to the maximum flux in the pulse profile) in 1 -- 10 keV energy band (MECS
data), during these two observations is in the range of 70 -- 75%. However, the MECS pulse
profile of the 2000 observation is different. The dip seems to be absent and the pulse
fraction is reduced to about 30%. Though the PDS pulse profiles of all three BeppoSAX
observations look alike, the pulse fraction during the 2000 observation is less (∼ 50%)
compared to the 1996 and 1997 observations (70 -- 80%).
To get a detailed picture of the pulsation properties in this extended low state during
the 2000 observation, we divided the light curve into three different regions as shown in
Figure 1. Pulsations with profiles similar to those seen in the 1996 and 1997 observations
were detected in the MECS light curves of regions 1 and 3, whereas in region 2, the 3σ
upper limit of soft X-ray pulse fraction is 10%. However, pulsations are detected in all
three regions in hard X-ray light curve of the PDS. The MECS and PDS pulse profiles of
the three different regions are shown in Figure 3.
-- 5 --
Figures 2 and 3 show a clear difference in the shape of pulse profiles of GX 1+4 in the
soft and hard X-ray energy bands. The dip is narrow in soft X-rays and gradually becomes
broader with increasing energy. To investigate the change in the shape of pulse profiles, we
generated light curves in 16 different energy bands from MECS and PDS event files of the
1996 observation. The corresponding pulse profiles are shown in Figure 4. An increase in
the width of the dips in the pulse profiles with energy range is apparent from the figure. It
is found that the light curve above 100 keV is mainly background dominated and pulsations
were not detected above 100 keV. Energy resolved pulse profiles from region 2 of the 2000
observation are shown in Figure 5. The pulse profiles in different energy bands of region 1
and 3 of 2000 August observation are similar to those seen in 1996 and 1997 observations.
But, as mentioned previously, pulsations are absent or much reduced in 1 -- 5.5 keV energy
band of region 2 of 2000 observation (Figure 5). A change in pulse shape is also noticed
between the pulse profiles in 5.5 -- 10 keV (a double peaked profile with a pulse fraction of a
mere ∼ 13%) and the same in the higher energy bands.
4. Pulse phase averaged spectroscopy
For spectral analysis, we have extracted LECS spectra from regions of radius 6′
centered on the object for the 1997 and 2000 observations. The combined MECS source
photons (MECS 1 was not operational in 2000) were extracted from circular regions with a
4′ radius. Background spectra for both LECS and MECS instruments were extracted from
appropriate source-free regions of the field of view by selecting annular regions around the
source. The software package named SAX Data Analysis System (SAXDAS) was used to
extract background subtracted PDS spectra from the event files of all the three BeppoSAX
observations. For spectral fitting, response matrices released by SDC in 1998 November,
were used. A rebinning scheme (ftp://heasarc.gsfc.nasa.gov/sax/cal/responses/grouping)
suggested in the SAX data analysis guide was used to rebin the LECS spectra to allow
use of the χ2 statistics. Events were selected in the energy ranges 1.0 -- 4.0 keV for LECS,
1.65 -- 10.0 keV for MECS and 15.0 -- 150 keV for PDS where the instrument responses are well
determined. Combined spectra from the LECS, MECS and PDS detectors, after appropriate
background subtraction, were fitted simultaneously. All the spectral parameters, other than
the relative normalization, were tied to be the same for all the detectors.
The spectra were first fitted to a model consisting of a power law continuum with
a Gaussian for the iron line emission and absorption by matter along the line of sight,
which yielded poor values of reduced χ2. Following this, we fitted the spectra with a model
consisting of a Comptonized continuum along with a Gaussian function and interstellar
-- 6 --
absorption and obtained better values for the reduced χ2. Addition of an absorption edge
at ∼ 30 keV further improved the spectral fitting with a decrease in reduced χ2 from 1.1
to 1.0 for region-1 and 1.50 to 1.41 for region-3, whereas the region-2 did not show any
improvement in spectral fitting with a reduced χ2 of 2.63. The edge energies and the
absorption depth at the threshold are found to be consistent (within errors) for all three
BeppoSAX observations. The residual around 35 keV can be fitted with either a cyclotron
absorption feature or an absorption edge. However, as the magnetic field strength of the
neutron star is estimated to be of the order of 1013−14 G, the cyclotron absorption line is
expected at higher energies (> 100 keV). As there is no physical reason for the absorption
feature at ∼ 35 keV, the absorption edge detected here is possibly an instrumental effect.
But we note that with a similar analysis of the PDS spectrum of Crab, we did not detect
any such feature.
The equivalent width of the 6.4 keV iron Kα line was found to be in the range of
180 -- 320 eV during the 1996, 1997 and regions 1 and 3 of the 2000 observation. In the
region 2, the equivalent width of the iron line was very high (∼ 2.9 keV) with significant
line like residuals just above 7 keV. The region 2 spectrum, when fitted with the addition
of another single Gaussian function gives a much smaller reduced χ2 of 1.72. We have
considered if the second line detected most clearly in the region 2 spectrum of 2000 can
be emission from hydrogen like iron as was found in an observation of GX 1+4 with the
Chandra HETG (Paul et al. 2004). The 90% confidence limit for the centre energy of the
second line detected in the MECS spectrum of 2000 region 2 is 7.10±0.05 keV. If we fix the
center energy of the second line at 6.95 keV, (at which the second line is detected with a
3σ confidence level in the Chandra spectrum) for 65 degrees of freedom the χ2 increases
by 10 compared to a line at 7.10 keV. We therefore conclude that the second emission
line detected in the MECS spectrum is indeed Kβ line of lowly ionized or neutral iron.
Following this, we have added the second line feature in the model for all the spectrum. The
emission line at ∼ 7.1 keV, with an equivalent width in the range of 25 -- 80 eV is present in
all BeppoSAX spectra of GX 1+4 except region 2 of the 2000 observation where it is very
strong with an equivalent width of about ∼ 0.55 keV.
A soft X-ray excess was detected in the residuals of the spectra from the 1996 and region
2 of the 2000 data. We tried various model components such as blackbody, disk-blackbody,
and bremsstrahlung to fit this soft excess. However, addition of a bremsstrahlung component
to the spectral model gives a better fit compared to the other two models for the soft excess
with an improved reduced χ2 of 1.11 and 1.27 for 1996 and the region-2 of 2000 observation
respectively. The integrated flux of the soft component in 1.0 -- 5.0 keV energy band for 1996
and region-2 of 2000 observations is estimated to be 1.12 × 10−12 ergs cm−2 s−1 and 4.7 ×
10−13 ergs cm−2 s−1 which is ∼ 4% and 18% of the total model flux in above energy band
-- 7 --
respectively. The 3σ lower limit of above two observations are 2.2 × 10−13 ergs cm−2 s−1
and 4.2 × 10−13 ergs cm−2 s−1 respectively whereas the 3σ upper limit of the soft X-ray
flux during 1997, region-1 and region-3 of 2000 observations are 6.3 × 10−13, 4.1 × 10−13
and 1.6 × 10−13 ergs cm−2 s−1 respectively. We did not try the non-solar abundance for the
soft component as it is seen only in 1996 and region-2 of 2000 observations and not in the
others. Therefore, the detected soft excess in GX 1+4 is unlikely to be due to abundance
anomaly. The spectral parameters with 3σ error estimates and the reduced χ2 obtained
are given in Table 2. The region-2 spectrum along with the best fit model components and
residuals without and with the iron Kβ line are shown in Figure 6. The unfolded spectra
measured during 1996, 1997 and 2000 BeppoSAX observations are shown in Figure 7.
5. Pulse phase resolved spectroscopy
Since the dip in the pulse profile of GX 1+4 shows significant energy dependence,
pulse phase resolved spectroscopy may hold a clue to the origin of the dip. To estimate the
change in spectral parameters during the dips, we have carried out pulse phase resolved
spectroscopy during all the observations. The LECS and MECS spectra were accumulated
into 10 pulse phase bins by applying phase filtering in the FTOOLS task XSELECT and
SAXDAS software package was used to extract PDS spectra at the same pulse phases. As
in the case of phase-averaged spectroscopy, the background spectra were extracted from
source free regions in the event files and appropriate response files were used for the spectral
fitting. The LECS, MECS and PDS spectra of region 1 of 2000 observation were added
with the corresponding spectra of region 3 to improve the statistics.
For the phase resolved spectral analysis, initially all the spectral parameters were
allowed to vary. However, we did not find any systematic variation in the value of NH
and iron emission line parameters with the pulse phase. We tried to fit the phase resolved
spectra of all three BeppoSAX observations by freezing NH , the absorption edge and iron
line parameters, which did not show any variation over pulse phase, to the phase averaged
values. We included the Bremsstrahlung component to the spectral fitting for the 1996 and
2000 region-2 phase resolved spectra as it was present in the phase averaged spectra. It
was found that the change in continuum flux with pulse phase in 2 -- 10 keV and 10 -- 100
keV energy ranges is consistent with that of the MECS and PDS pulse profiles respectively.
We found a marginal change in the values of the input spectrum temperature kT0, plasma
temperature kT and the scattering optical depth τ over pulse phase in three BeppoSAX
observations. The results of 1996, 1997 and combined regions of 1 & 3 of 2000 observations
are found to be similar. It is found that the results obtained from the phase resolved
-- 8 --
spectroscopy of region-2 of 2000 observation are similar to that obtained by Galloway et al.
(2001) e.g. an increase in the optical depth at the pulse minimum.
6. Discussion
6.1. Pulsations during the extended low state
On several occasions, the accretion powered X-ray pulsar GX 1+4 was found to show
a decrease in X-ray luminosity by an order of magnitude compared to the flux before and
after such a state. The low state lasts for at least a few hours and can have rapid ingress
to and egress from the low state within only about 4-6 hours (Giles et al. 2000; present
work). Many short exposures with RXTE-PCA also found the source at a very low flux
level, about an order of magnitude lower than usual, (Cui 1997, Cui & Smith 2003). During
these low flux episodes, the pulsations were either absent, or the pulse fraction was reduced
significantly from the usual 50-80%, often below the detection limit of RXTE-PCA. This
has been interpreted as due to onset of centrifugal barrier or the propeller effect. With
decrease in the mass accretion rate, when the inner radius of the accretion disk recedes
beyond the co-rotation radius of the neutron star, this effect may set in. To compare the
2 -- 10 keV flux during the low-state reported here with the same during earlier reported low
intensity episodes with the RXTE-PCA, when the pulsations became undetectable, we have
reanalyzed the RXTE-PCA data of those observations. The 2 -- 10 keV flux measured from
these observations are shown in Table 2.
During the low state reported here, the absorbed X-ray flux in 2 -- 10 keV band was
2.9× 10−11 ergs cm−2 s−1 and is comparable to the earlier episodes when pulsations were
below the detection level of the RXTE. However, we found that as the X-ray spectrum
of GX 1+4 is very hard, the total X-ray luminosity in 10 -- 100 keV energy band during
this state is more than an order of magnitude higher than that in the 2-10 keV band and
pulsations are clearly detected in the hard X-rays. The spectral analysis of the BeppoSAX
data showed that in the low state, the medium energy (2 -- 10 keV) X-ray flux is reduced by
an order of magnitude, mostly due to an increase in the absorption. In comparison, the
hard X-ray flux is reduced by a factor of two only. It was also pointed out by Cui (1997)
from RXTE-PCA data that the X-ray spectrum is harder in the low state. Therefore, there
can be significant flux in the hard X-ray band, where the effective area of the RXTE-PCA
detectors fall rapidly. It is possible that the low states reported by Cui (1997) and Cui and
Smith (2004) are similar to the one presented here. Only short exposures and decreasing
hard X-ray effective area of the RXTE-PCA did not allow detection of the hard X-ray
pulses in RXTE-PCA observations during the low states. In such a scenario, the propeller
-- 9 --
regime for GX 1+4 may occur at a still lower mass accretion rate than that reported earlier.
6.2. Broad band X-ray spectrum and its changes during the extended
low-state
From the three BeppoSAX observations reported here, the pulse-phase-averaged broad
band X-ray spectrum in both high and low states of GX 1+4 is found to agree well with a
Comptonization model with additional components for absorption, line emissions, and in
some cases, a soft excess. During the extended low state in 2000, the continuum flux in
the 2 -- 10 keV decreased by a factor of ∼10 whereas the 10 -- 100 keV flux decreased only by
about 50%. An intense iron emission line at 6.4 keV is clearly seen in the phase averaged
spectra in both the high and low intensity states. In addition, a strong emission line at
7.1 keV was also detected in the low state spectrum, which was very weak during the high
intensity states. The Kα line flux was about 3 -- 5 times larger than the Kβ line flux in all
the observations and is consistent with the ratio expected for fluorescence emission from
neutral iron. With a moderate energy resolution of MECS it is not possible to determine
the exact ionization state of the reprocessing material.
We have also examined changes of the two hardness ratios (1.0 -- 4.0/7.0 -- 10.0 keV and
4.0-7.0/7.0 -- 10.0) in the MECS energy bands during the BeppoSAX observation in 2000.
The hardness ratios clearly show increased absorption during the low state. There was
no significant spectral change associated with the intensity variations near the end of this
observation.
Although a similar low state of GX 1+4 was observed earlier, the event lasted for only
about 6 hours and detailed spectroscopy was not feasible due to poorer photon statistics
and lower energy resolution of the RXTE-PCA (Galloway et al. 2000). In this paper, we
present the first detailed spectroscopy of GX 1+4 in this state. Though both the soft and
hard X-ray flux decreased during the low state, the optical depth and temperature of the
Comptonizing plasma are comparable (within errors) with those during the high intensity
states just before and after the low intensity episode. The absorption column density is
an order of magnitude higher during the low state. An increase in absorption may not
be sole reason for the low state as the hard X-ray flux also decreased by about 50%, and
the temperature of the seed photons increased during the low state by a factor of about
two. However, the decrease in the hard X-ray flux may arise due to Thompson scattering
from the line of sight if the absorbing material is partially ionized and the actual material
in the line of sight is more than that measured with the simple absorber model. In the
Comptonization model for X-ray pulsar spectra, the seed photons are likely to be produced
-- 10 --
from a heated surface of the neutron star. If there is a spherical component of accretion, it
may cause heating of the surface and these seed photons undergo Compton upscattering in
the accretion column. An increase in the temperature of the seed photons during the low
state also indicates that the low state is not only due to an increased absorption along the
line of sight. There must have been some additional change in the accretion process, for
example, and enhancement in the spherical component of mass accretion.
Fluorescence emission lines with very high equivalent width similar to those found in
GX 1+4 in the low state are seen in some persistent X-ray binary pulsars like Her X-1,
LMC X-4 and SMC X-1 during their low-intensity phases of the super orbital period (Naik
& Paul 2003; Naik & Paul 2004, Vrtilek et al. 2001). In these systems, the low-states arise
due to increased absorption of the X-ray continuum by precessing warped inner accretion
disk, resulting in a very large equivalent width. However, unlike these systems, in GX 1+4
the extended low state is observed intermittently, without any hint of a periodicity. The
RXTE/ASM light curve of GX 1+4 does not show any periodic or quasi-periodic long
term intensity variation as seen in Her X-1, LMC X-4 and SMC X-1. The difference in the
duration and absence of periodicity in the occurrence of the low states of GX 1+4 rules out
a similar mechanism for this source.
6.3. Spectral features of the absorption dip in pulse profiles of GX 1+4
Energy resolved pulse profiles of GX 1+4 obtained from BeppoSAX observations
provide useful information towards understanding the accretion geometry of the binary
system. The characteristic absorption dips are detected in wide energy range (1 -- 100
keV). The pulse profiles in narrow energy bands as presented here in this paper, show a
gradual change with energy. An increase in width of the absorption dip with energy is
clearly seen in all three BeppoSAX observations. These narrow and sharp dips have been
attributed to the interaction between the accretion column and the emission region of the
neutron star (Galloway et al. 2001). Partial eclipses of the emission region by the accretion
column occurring during the spinning of the neutron star are reflected as dips in the pulse
profile. The closest approach of the accretion column to the line of sight corresponds to
the minimum intensity in the pulse profile. Similar sharp and narrow dips are observed
in the pulse profiles of two other X-ray pulsars RX J0812.4 -- 3114 (Reig & Roche 1999)
and A 0535+262 (Cemeljic & Bulik 1998). However, a gradual increase in the width of
the absorption dip with energy is seen only in GX 1+4. A widening of the dip at higher
energies may result from changes in the beam pattern with energy. Outside the dip, the
high energy photons are typically those which have undergone many scatterings. Except
-- 11 --
along the accretion column, where both the low and high energy photons are strongly
suppressed, the high energy photons escape from the column preferentially at large angles,
whereas low-energy photons are more isotropic. The emission geometry may change from a
pencil beam at low energy to a fan beam at high energy.
We have detected marginal evidence of spectral changes during the pulse phase in
the form of a varying optical depth, plasma temperature and temperature of the soft seed
photons. The pulse phase dependence of these parameters is not identical during different
observations and the implication of the pattern of these variations is not clear.
Acknowledgments
We thank an anonymous referee for suggestions which helped us to improved the
content of this paper. The Beppo-SAX satellite is a joint Italian and Dutch program. We
thank the staff members of Beppo-SAX Science Data Center and RXTE/ASM group for
making the data public. SN is supported by IRCSET through EMBARK postdoctoral
fellowship.
REFERENCES
Boella, G., Butler, R. C., Perola, G. C., et al. 1997, A&AS, 122, 299
Cemeljic, M., & Bulik, T. 1998, Acta Astron., 48, 65
Chakrabarty, D., Bildsten, L., Finger, M. H., et al. 1997, ApJ, 481, L101
Chakrabarty, D. & Roche P. 1997, ApJ, 489, 254
Table 1. BeppoSAX observations of GX 1+4
Year of
Observation
Start Time
(Date, UT)
End Time
(Date, UT)
LECS Exp. MECS Exp.
PDS Exp.
(ks)
(ks)
(ks)
1996
1997
2000
August 18, 06:11 August 19, 03:38
March 25, 22:43 March 26, 16:08
August 29, 12:36
September 02, 03:38
-- --
13
56.5
38.6
31.5
132
17.6
13.5
58.5
-- 12 --
Table 2. Spectral parameters for GX 1+4 during different intensity states
Parameter
1996 August
1997 March
2000 August
2000 August
2000 August
Region -- 1
Region -- 2
Region -- 3
1
NH
kTBr (keV)
T0 (keV)
kTCo (keV)
τ
21±2
0.36+0.12
−0.11
1.48+0.07
−0.06
13+0.4
−0.5
3.04+0.19
−0.14
1.1±0.1
-- --
1.28+0.04
−0.05
12.7+1.2
−1.1
3+0.4
−0.3
1.37+0.05
−0.15
-- --
1.62+0.04
−0.02
15.2+0.5
−1.6
2.38+0.38
−0.09
14.8+1.0
−3.4
0.26+0.06
−0.04
3.5+0.6
−0.4
12.6+1.9
−0.9
3.1+0.3
−0.5
2.4±0.1
-- --
1.73±0.03
14.3±0.6
2.55±0.13
Iron Kα emission line
FeE (keV)2
Line width (keV)
Eqw. width (keV)
Line Flux3
6.42±0.02
0.1±0.02
0.22±0.01
1.31+0.07
−0.10
6.45+0.03
−0.02
0.05∗
0.19±0.02
0.85±0.08
6.49±0.03
0.12±0.06
0.24±0.03
0.77±0.09
6.47±0.01
0.05∗
3.0+0.6
−0.3
1.08+0.21
−0.12
6.45±0.02
0.07±0.05
0.33±0.03
1.04±0.09
Iron Kβ emission line
FeE (keV)2
Eqw. width (keV)
Line Flux3
6.99+0.30
−0.05
0.06±0.01
0.38±0.08
7.06+0.10
−0.07
0.06±0.02
0.25+0.10
−0.08
7.1†
0.02±0.02
0.05+0.10
−0.01
7.10+0.02
−0.05
0.55+0.14
−0.09
0.23+0.06
−0.04
7.01+0.08
−0.09
0.08+0.04
−0.02
0.25+0.12
−0.07
Eedge (keV)
Absorption depth
Reduced χ2
33+1
−2
0.13+0.03
−0.05
1.18 (239 dof)
30±3
0.28+0.12
−0.11
1.08 (271 dof)
33+2
−3
0.19+0.11
−0.07
1.0 (168 dof)
32+13
−8
0.11+0.11
−0.06
1.27 (158 dof)
35±2
0.15±0.06
1.23 (168 dof)
Model Flux4
2 -- 10 keV
10 -- 100 keV
2.4+0.3
−0.2
14.9+1.1
−0.9
3.2+0.5
−0.3
7.3+0.8
−1.1
2.07+0.34
−0.08
10.1+1.6
−0.3
0.29+0.08
−0.05
5.1+1.1
−0.9
2.01+0.13
−0.15
11.2+0.7
−0.8
The errors given here are for 90% confidence limit. 1 : 1022 atoms cm−2, 2 : Iron emission line energy, 3 :
10−11 ergs cm−2 s−1, 4 : 10−10 ergs cm−2 s−1
Fitted Model = Wabs (Br + CompTT + Ga1 + Ga2) Edge. Wabs = Photoelectric absorption
parameterized as equivalent hydrogen column density NH , Br = thermal-bremsstrahlung-type component
with plasma temperature kTBr, CompTT = thermal Comptonized component, and Ga1 & Ga2 = Gaussian
function for iron Kα and Kβ emission lines. ∗ : Upper limit to the line width, † : Parameter kept fixed.
2 -- 10 keV flux during the low state observed in GX 1+4 with RXTE on various occasions are
1996 July 20 RXTE Observation (Giles et al. 2000) = 0.35 × 10−10 ergs cm−2 s−1
1996 September 25 RXTE observation (Cui 1997) = 0.40 × 10−10 ergs cm−2 s−1
2002 July 2 RXTE observation (Cui & Smith 2004) = 0.23 × 10−10 ergs cm−2 s−1
-- 13 --
Cui, W. 1997, ApJ, 482, L163
Cui, W. & Smith, B. 2004, ApJ, 602, 320
Dotani, T, Kii, T., Nagase, F., et al. 1989, PASJ, 41, 427
Frontera, F., Costa, E., Dal Fiume, D., et al. 1997, A&AS, 122, 357
Galloway, D. K. 2000, ApJ, 543, L137
Galloway, D. K., Giles, A. B., Greenhill, J. G., & Storey, M. C. 2000, MNRAS, 311, 755
Galloway, D. K., Giles, A. B., Wu, K., & Greenhill, J. G. 2001, MNRAS, 325, 419
Ghosh, P., & Lamb, F. K. 1979, ApJ, 234, 296
Giles, A. B., Galloway, D. K., Greenhill, J. G., et al. 2000, ApJ, 529, 447
Greenhill, J. G., Sharma, D. P., Dieters, S. W. B. et al. 1993, MNRAS, 260, 21
Makishima, K., Ohashi, T., Sakao, T., et al. 1988, Nature, 33, 746
Naik, S., & Paul, B. 2003, A&A, 401, 265
Naik, S., & Paul, B. 2004, ApJ, 600, 351
Paul, B., Rao, A. R., & Singh, K. P. 1997a, A&A, 320, L9
Paul, B., Agrawal, P. C., Rao, A. R., & Manchanda, R. K. 1997b, A&A, 319, 507
Paul, B., Dotani, T., Nagase, F., Mukherjee, U., & Naik, S. 2004, ApJ (submitted)
Pereira, M. G., Braga, J., & Jablonski, F. 1999, ApJ, 526, L105
Rao, A. R., Paul, B., Chitnis, V. R., Agrawal, P. C., & Manchanda, R. K. 1994, A&A, 289,
L43
Reig, P., & Roche, P. 1999, MNRAS, 306, 95
Vrtilek, S. D., Raymond, J. C., Boroson, B., Kallman, T., Quaintrell, H., & McCray, R.
2001, ApJ, 563, L139
This preprint was prepared with the AAS LATEX macros v4.0.
-- 14 --
Fig. 1. -- The MECS and PDS light curves with time bin same as the spin period of the
neutron star obtained from the 2000 BeppoSAX observation of GX 1+4.
Fig. 2. -- The MECS and PDS pulse profiles of the source, of the three observations, are
shown in top and bottom panels with 32 phase bins per pulse respectively. Two pulses are
shown for clarity.
-- 15 --
Fig. 3. -- The MECS and PDS pulse profiles of the three different regions of 2000 BeppoSAX
observation (as shown in Figure 1) of GX 1+4.
Fig. 4. -- The MECS and PDS pulse profiles of GX 1+4 in 16 different energy bands (as
described in text) of 1996 August observation.
-- 16 --
Fig. 5. -- The MECS and PDS pulse profiles of GX 1+4 in 12 different energy bands (as
described in text) of region-2 of 2000 August observation.
-- 17 --
Fig. 6. -- Energy spectra of GX 1+4 measured with BeppoSAX during the region 2 of 2000
August observation along with the fitted model (without iron Kβ emission line) and the
residuals (middle panel). The presence of iron Kβ line is evident from the line like structure
at ∼ 7.1 keV in the residual. Residuals obtained after including a Gaussian function at 7.1
keV in the model are shown in the bottom panel.
-- 18 --
Fig. 7. -- The unfolded energy spectra of GX 1+4 measured with 1996 August, 1997 March,
and three regions of 2000 August BeppoSAX observations.
|
astro-ph/0011206 | 1 | 0011 | 2000-11-09T18:36:18 | Exploding and Non-exploding Stars: Coupling Nuclear Reaction Networks to Multidimensional Hydrodynamics | [
"astro-ph"
] | After decades of one-dimensional nucleosynthesis calculations, the growth of computational resources has meanwhile reached a level, which for the first time allows astrophysicists to consider performing routinely realistic multidimensional nucleosynthesis calculations in explosive and, to some extent, also in non-explosive environments. In the present contribution we attempt to give a short overview of the physical and numerical problems which are encountered in these simulations. In addition, we assess the accuracy that can be currently achieved in the computation of nucleosynthetic yields, using multidimensional simulations of core collapse supernovae as an example. | astro-ph | astro-ph |
Exploding and Non-exploding Stars:
Coupling Nuclear Reaction Networks
to Multidimensional Hydrodynamics
K. Kifonidis∗, T. Plewa†,∗ and E. Muller∗
∗Max-Planck-Institut fur Astrophysik, Karl-Schwarzschild-Strasse 1, D-85741 Garching,
†Nicolaus Copernicus Astronomical Center, Bartycka 18, 00716 Warsaw, Poland
Germany
Abstract. After decades of one-dimensional nucleosynthesis calculations, the growth
of computational resources has meanwhile reached a level, which for the first time allows
astrophysicists to consider performing routinely realistic multidimensional nucleosyn-
thesis calculations in explosive and, to some extent, also in non-explosive environments.
In the present contribution we attempt to give a short overview of the physical and
numerical problems which are encountered in these simulations. In addition, we as-
sess the accuracy that can be currently achieved in the computation of nucleosynthetic
yields, using multidimensional simulations of core collapse supernovae as an example.
INTRODUCTION
Thermonuclear reactive flows are ubiquituous in astrophysics and occur in non-
explosive environments as, e.g., in most (hydrostatic) stars as well as in explosive
events, for which novae and supernovae are examples. Often they provide the en-
ergy which powers stellar outbreaks (as in the case of novae, X-ray flashes, and
thermonuclear, i.e. Type Ia, supernovae) and even for stellar explosions where this
is not the case (as e.g.
in core collapse supernovae, which are driven by neutrino
heating), the strong coupling of hydrodynamic advection and thermonuclear re-
actions is of utmost importance for the nucleosynthesis which accompanies these
events.
It is by a proper numerical modelling of this coupling through which a
more detailed insight into the origin of the nuclear abundances in the solar system
can be gained, which are themselves the result of a superposition of material which
has been processed in explosive and non-explosive thermonuclear environments.
By comparing the results of numerical models with the observed solar abundance
pattern, on the other hand, one might also hope to learn more about the thermo-
dynamic conditions in the otherwise unaccessible nucleosynthetic sites and events
themselves.
The high precision with which nuclear abundances can be measured nowadays
poses great demands on the accuracy of the numerical models, especially since it
was convincingly demonstrated in recent years that due to the importance of hydro-
dynamic instabilities, rotation, and other effects, most of the nucleosynthetic sites
do not possess spherical symmetry. Thus a reliable computation of the highly non-
linear interaction of hydrodynamic advection and nuclear burning requires multi-
dimensional numerical models. In the following sections we give a general overview
of the methods which are currently employed for modelling thermonuclear flows
and discuss some of the problems which are hereby encountered. Further reviews
on reactive flow modelling can be found in [9], [10] and [11].
THE GOVERNING EQUATIONS
A rather wide range of astrophysical reactive flows, in which relativistic effects,
viscosity and magnetic fields can be neglected,
is described by the well-known
(reactive) Euler equations. This system of non-linear partial differential equations
which expresses the conservation of the total mass, momentum, total (i.e. kinetic
+ internal) energy and baryons of the fluid reads
∂ρ
∂t
+ ∇ · (ρv) = 0
+ ∇ · (ρvv) + ∇P = ρg + ρfadd
+ ∇ · ([ρE + P ] v) = ρv · g + ρ Qadd + ρ Qnuc
∂ρv
∂t
∂ρE
∂t
∂ρXi
∂t
+ ∇ · (ρXiv) = ρ Xi
Xi = 1,
X
i
(1)
(2)
(3)
(4)
(5)
where ρ, v, E = v2/2+e, and P have their usual meanings, Xi is the mass fraction of
nucleus i, and ρ Xi as well as ρ Qnuc are source terms due to nuclear transmutations.
If self-gravity is important, the gravitational acceleration
g = −∇Φ
(6)
which appears in the source terms of Eqs. (2) and (3) and which depends on the
gravitational potential, Φ, has to be obtained from a solution of the Poisson equa-
tion
∆Φ = 4πGρ.
(7)
In mathematical terms Eqs. (1 -- 7) describe a mixed initial/boundary value prob-
lem due to the hyperbolic and elliptic nature of the Euler and Poisson equations,
respectively. Given appropriate initial and boundary conditions, an equation of
state relating ρ, P and e, and the additional source terms fadd and Qadd, which in
general will be problem-dependent, Eqs. (1 -- 7) can be solved after an appropriate
flow representation as well as a suitable numerical scheme have been adopted.
FLOW REPRESENTATIONS AND NUMERICAL
SCHEMES
There are two primary approaches to solve the homogeneous part of the system
of equations (1 -- 5). In the Eulerian framework the system of conservation laws is
solved on a grid which is fixed in space and the evolution of the flow is followed
by advecting the fluid through the computational cells. The principal assets of
this method are its straightforward extension from one to two or three spatial
dimensions and the simplicity of its implementation on serial and parallel computer
architectures. If an appropriate shock-capturing, finite-volume numerical scheme
is used, it is equally straightforward to obtain strict numerical conservation of
all physically conserved quantities and a sharp resolution of shocks. The major
drawback is numerical diffusion. Consider the continuity equation (1) in its Eulerian
form, which can be written as
∂ρ
∂t
+ v · ∇ρ + ρ∇ · v = 0,
(8)
where the second term describes advection and the third term compression. Nu-
merical diffusion is introduced into a numerical solution of this equation as a result
of discretization errors of the v · ∇ operator. There appears to be a simple remedy
to this problem: using the comoving derivative d/dt = ∂/∂t + v · ∇ we can rewrite
Eq. (8) in the frame comoving with the matter to obtain its Lagrangian form
dρ
dt
+ ρ∇ · v = 0.
(9)
Note that in this frame the advection term v · ∇ρ has vanished. Therefore the
Lagrangian approach is (in principle) not prone to numerical diffusion of mass (or
composition). In Lagrangian methods each cell of the numerical grid represents a
discretized fluid element which evolves subject to forces which are due to interac-
tions with its neighbors and the time rate of change of the density of such a fluid
element is solely determined by the compression (or expansion) that it experiences.
Density interfaces (contact discontinuities) as well as composition discontinuities
can be easily aligned with the boundaries of grid cells and do not have to be ad-
vected through the grid in the course of the calculation.
While this very desirable property of the Lagrangian approach has made it the
method of choice for one-dimensional nucleosynthesis calculations, considerable dif-
ficulties are experienced when Lagrangian schemes are applied to multidimensional
flows. Shear and vortices can severely distort a Lagrangian grid. The discrete ap-
proximation of differential operators over such a grid results in large errors in the
80
60
40
20
)
s
e
n
o
Z
(
h
t
i
d
W
0
0
Donor Cell
Godunov
Lax-Wendroff
PPM
SADIE
PPM (without detection)
200
400
600
Distance (Zones)
FIGURE 1. Comparison of the diffusivity of different advection schemes for the problem of the
propagation of a contact discontinuity through an Eulerian grid. The curves give the width of
the discontinuity (in grid zones) as a function of the number of zones it has propagated through
the grid (adapted from [4] and [5]).
numerical derivatives, and in the extreme case that the grid lines cross (grid tan-
gling) the calculations have to be stopped. Some remapping procedure to a new,
more regular grid must then be applied which unavoidably introduces numerical
diffusion to the solution. The distortion problem can be overcome if triangular
instead of quadrilateral grids are used [11] or if (as in the Smoothed Particle Hy-
drodynamics, or SPH approach) no grid at all is adopted and instead the flow is
sampled by a finite number of particles. In the former method considerable logic
overhead is added in restructuring the deformed triangular grid, while in the latter
case, due to the Monte-Carlo nature of the sampling, Poisson noise is introduced.
Due to the aforementioned drawbacks and due to significant progress in the de-
velopment of accurate Eulerian schemes in the early 1980's, Lagrangian methods
employing quadrilateral or triangular grids have not been used extensively in mul-
tidimensional calculations of astrophysical flows (see [18], [12] as well as [11] and
the references therein for examples). On the other hand, the simplicity of SPH
has made this method very popular for astrophysical (especially cosmological) sim-
ulations. Without attempting to escalate the very vigorous discussion, whether
SPH or grid-based Eulerian schemes are to be prefered in astrophysical calcula-
tions (see e.g. [10]), we will argue below that, due to its Monte-Carlo nature, the
SPH scheme appears to be rather unsuited for multidimensional nucleosynthesis
calculations, especially in cases where hydrodynamic instabilities are known to be
important.
Among Eulerian schemes, the so-called shock-capturing schemes have proven to
be the most accurate ones for problems which involve discontinuities in the flow
as shock waves (see [19] for details). The latter are very frequently encountered in
explosive events, since in these cases the flows can attain supersonic speeds. Shock-
capturing schemes derive their accuracy from a discretization of the hydrodynamic
equations which closely mimics the physics of compressible flows by making use of
the Riemann problem, i.e. the dissolution of an arbitrary flow discontinuity into a
set of simple waves (shocks, contact discontinuities and rarefaction waves). Suitably
constructed Riemann problems at the interfaces between adjacent computational
cells are solved within each time step, from which the complete solution of the
system of conservation laws is constructed. This allows one to avoid the use of large
amounts of artificial viscosity in order to obtain a well-behaved numerical scheme
in the vicinity of shocks. One of the most accurate shock-capturing schemes, which
has been widely applied in astrophysics, is the (direct Eulerian) PPM scheme of
[3], a second order extension of Godunov's original (and rather diffusive) first-
order shock-capturing scheme [6]. In addition to its accurate treatment of shocks
PPM includes a special detection and steepening algorithm to minimize numerical
diffusion across contact discontinuities.
The superiority of shock-capturing schemes in computing compressible flows has
in [19], and their performance for computing reactive as-
been demonstrated e.g.
trophysical flows was studied in [4] and [13]. Fig. 1 shows a representative result
from [4] in which PPM was compared to a number of older Eulerian schemes which
were in wide-spread use until the mid 1980's. The figure shows the width of a
contact discontinuity as a function of the number of zones that it has travelled
across a numerical grid. Most Eulerian schemes tend to smear such fluid (and also
composition) interfaces without limit, i.e. the width of the "discontinuity" tends to
grow with time. Of all the schemes investigated, only PPM maintained a sharp res-
olution of the interface within two zones. Still however, numerical diffusion cannot
be completely avoided in Eulerian calculations and its minimization necessitates an
adequate spatial resolution in addition to an excellent advection scheme. This has
led to the development of adaptive mesh refinement methods [1], which concentrate
the computational effort in critical regions of the flow and thereby often allow for
substantial savings in computer time.
ADDITIONAL PHYSICS
While the numerical problems encountered in solving the homogeneous part of
the Euler equations are difficult to overcome, they represent only a part of the
computational difficulties for a realistic simulation. The source terms, which are
usually taken into account using the operator splitting technique [9], often require
much more computer time than the solution of the hydrodynamic equations them-
selves. This holds, e.g.
if large nuclear networks need to be evolved with the
hydrodynamics or transport processes need to be taken into account (as e.g. neu-
trino transport in core collapse supernovae, see [15] and the references therein and
A. Burrows, this volume). In some cases even phenomenological (sub-grid) models
might have to be introduced. This is e.g. the case for turbulent combustion in ther-
monuclear supernovae, where a white dwarf is incinerated by a deflagration front
whose propagation speed is impossible to compute in a direct simulation since this
would require a resolution of the turbulent energy cascade down to the dissipation
length scale [16]. Exacerbating the situation is the fact that stellar models, which
serve as initial data for supernova simulations, might be affected by considerable
uncertainties.
In the absence of computational schemes and resources which al-
low for a consistent multidimensional treatment of stellar convection and rotation
over stellar evolutionary time scales, one is forced to describe these phenomena by
one-dimensional appoximations (see the contribution of N. Langer, this volume).
Finally, uncertainties in nuclear reaction rates enter the calculations.
It is apparent that progress in only a single of the involved fields is not going to
improve the accuracy of the desired nucleosynthetic yields considerably. In fact, a
concerted effort in all areas appears to be required, since, as we will show in our
example below, the different effects can conspire in falsifying the nucleosynthetic
yields.
CORE COLLAPSE SUPERNOVAE: A CASE STUDY
Nucleosynthesis in core collapse supernovae is a good example for illustrating the
aforementioned problems and we will start with a discussion of numerical diffusion
using results of simple one-dimensional calculations. We subsequently address the
complications introduced by convection in multidimensional calculations as well as
by "additional physics", i.e. neutronization due to neutrino matter interactions.
Finally we show how a multidimensional numerical failure, the so-called "odd-
even-decoupling" phenomenon, an instability which appears to plague most shock-
capturing schemes and whose effects have not yet been discussed extensively in
the numerical astrophysics literature, can enhance neutronization by strengthening
hydrodynamic convection and affect the nucleosythetic yields in multidimensional
simulations.
Nucleosynthesis in a 15 M⊙ star (1D)
In core collapse supernovae nucleosynthesis is triggered by a shock wave which
forms after the collapse of the iron core of a massive star has proceeded to supranu-
clear densities. The shock, while initially powerful, stalls after a few milliseconds
due to the energy losses from which it suffers while propagating through the outer
iron core, but eventually ejects the outer stellar layers if heating by neutrinos from
the collapsed core is able to overcompensate for the energy losses.
In most nu-
cleosynthesis calculations, however, these processes are not modelled in detail and
instead a shock is initiated by simply depositing the typical observed supernova
energy of ∼ 1051 erg near the center of a presupernova model.
FIGURE 2. Left: Eulerian PPM calculation of explosive nucleosynthesis in the presupernova
model of [20] using an α-nucleus network and the Consistent Multifluid Advection scheme (CMA)
(from [13]). Right: Comparison of Eulerian PPM results using: 1st order advection for nuclear
species (top), the FMA advection scheme for multifluid flows of [4] (middle), and the CMA scheme
of [13] (bottom). Note the decreasing amount of diffusion and the sharp interfaces obtained with
CMA (from [13]).
In the left panels of Fig. 2 we show snapshots of the mass fractions from the first
500 ms of such a calculation, focusing on the silicon-rich layers just outside the iron
core in which explosive nucleosynthesis takes place. Only 100 ms after the start
of the calculations explosive silicon burning has frozen out and has left behind a
significant abundance of 56Ni as well as 40Ca, 36Ar and 32S. Particularly noteworthy
for the following discussion is the nucleus 44Ti. From one-dimensional Lagrangian
nucleosynthesis calculations [17], [21] it is known that this isotope should be primar-
ily synthesized in the innermost stellar layers which experience an α-rich freezeout.
However, in the present Eulerian calculation a significant abundance of 44Ti has
formed in zones with mass coordinates around 1.36 M⊙ (marked with arrows in
the left panels of Fig. 2), i.e. at the interface of the regions enriched in 4He and
40Ca. This 44Ti "bump" results from the reaction 40Ca(α, γ)44Ti and the amount
FIGURE 3. Final 44Ti yield of the one-dimensional calculations shown in Fig. 2 as a function
of radial resolution and different multifluid advection schemes. Top curve: 1st order species
advection. Middle: FMA. Bottom: CMA (from [13]).
of 44Ti thereby produced is very sensitive to numerical diffusion in this region.
Consequently, the strength of the 44Ti "bump" varies with the diffusivity of the
numerical scheme which is used to advect the nuclear species. This is illustrated
in the right panels of Fig. 2. The top right panel depicts results from a calculation
with Godunov's first-order scheme. In this case, all mass fraction profiles are heavily
smeared due to strong diffusion, as can be seen by a comparison to the middle right
and bottom right panels which show results that were obtained with the FMA and
CMA advection schemes of [4] and [13], respectively. Of all these schemes, CMA is
the least diffusive since it is the only method which includes a detection and steep-
ening algorithm for composition interfaces which was derived from PPM's original
detection and steepening algorithm for contact discontinuities. Note the size of the
44Ti bump for the three different runs. The more diffusive schemes produce much
more 44Ti. This is also illustrated in Fig. 3 which summarizes how the 44Ti yield
depends on the adopted advection scheme and the spatial resolution. While the
CMA results (bottom curve) are already converged for a resolution of ∆r = 40 km,
FMA (middle curve) needs a resolution of about 10 km. The first-order scheme (top
curve) would need much finer zoning than ∆r = 5 km to yield results of comparable
quality. Note also that, if a diffusive advection scheme and coarse resolution are
used, the errors might be as large as a factor of four!
It should be pointed out, however, that 44Ti is a somewhat extreme (though very
important) example. The (relative) errors due to numerical diffusion are usually
smaller for the more abundant nuclei. This is illustrated in Fig. 4 which shows
the dependence of the yields of different α-nuclei on resolution in a 1D calculation
from [8], in which no ad hoc energy deposition was adopted, but where the shock
revival phase was followed in detail by including the effects of neutrino heating
from a central light bulb neutrino source [7]. It can be seen from this figure that
FIGURE 4. Accuracy of various elemental yields from one-dimensional shock-revival and explo-
sive nucleosynthesis calculations in the post-bounce model of [2] (see also [20] for the presupernova
model). The logarithm of the deviations of the elemental yields of various α-nuclei as compared
to an essentially converged 6400 zone calculation (∆r ≤ 5 km) is displayed as a function of the
radial resolution (in grid zones)(from [8]).
individual elemental yields are more accurate than the 44Ti yield by more than an
order of magnitude. However, if one is aiming at a numerical accuracy of about 1%
for all yields, about 3000 radial zones are required for this calculation. This amount
of spatial resolution makes accurate multidimensional simulations very expensive.
Convection and 56Ni synthesis in core collapse supernovae
Neutrino matter interactions play a crucial role in the explosion of core collapse
supernovae. They heat the matter behind the stalling shock and thereby trigger the
explosion. On the other hand they determine the electron fraction per baryon, Ye,
(or equivalently the ratio of protons to neutrons) in the ejecta and thus influence the
nucleosynthetic yields. If the Ye value of material that has been photodissociated
by the shock is significantly reduced below 0.5 by neutrino/matter interactions, this
matter will recombine mainly to neutron-rich nuclei after expanding and cooling.
In that case nuclei with Z = N like 56Ni will not form in the ejecta. Fig. 5 which
shows results of a two-dimensional simulation of shock revival illustrates this effect.
In this calculation the luminosities of νe and ¯νe were such that ¯νe absorption on
protons was favored against νe absorption on neutrons and thus the heated gas
was neutronized [7]. The sharp division between this material which is visible in
the bubbles behind the shock in Fig. 5 and the lepton-rich material farther out
which has formed 56Ni (the region enclosed by the white contour line in Fig. 5) is
clearly visible. The negative entropy gradient that the heating has imprinted on the
layers between the radius of maximum neutrino heating (close to the center) and
the shock farther out, has also led to strong convective motions: bubbles of heated
FIGURE 5. Distribution of the entropy (in units of kB/nucl.) 320 ms after core bounce in a
two-dimensional core collapse supernova simulation of a 15 M⊙ star. The white contour line
encloses the region in which the 56Ni mass fraction exceeds 20% (from [8]).
deleptonized gas rise toward the shock while lower entropy flux tubes transport
lepton-rich matter to deeper layers where it interacts with the neutrino fluxes much
more efficiently. This interplay of convection and deleptonization is crucial for the
56Ni yield. The shock is only able to heat a certain amount of lepton-rich material
to temperatures in excess of the 5 × 109 K which are required for 56Ni synthesis. If
convection is strong enough to advect significant amounts of this matter close to
the neutron star, where the gas will experience deleptonization, the 56Ni yield will
be lower than in a model with no or weak convection.
Multidimensional numerical failures: Odd-even decoupling
Quirk [14] has reported on a subtle flaw in a number of shock capturing schemes
which becomes evident when calculating multidimensional flows with strong, grid-
aligned shocks. He has dubbed this failure the "odd-even decoupling" phenomenon.
The problem shows up only if a sufficiently strong shock is either fully or nearly
aligned with one of the coordinate directions of the grid, and if, in addition, the
flow is slightly perturbed. This can be due to either perturbations intentionally
introduced in order to study physical instabilities, as it is done in all studies of
convection in supernovae, or due to perturbations caused by other flow features.
Many Riemann solvers show the tendency to allow these perturbations to grow
without limit along the shock surface, thus triggering a strong rippling of the shock
front as well as the post shock state. In supernova simulations these perturbations,
whose amplitudes can exceed those of the seed perturbations by several orders of
magnitude, enhance the growth of hydrodynamic instabilities. In case of neutrino
driven convection they lead to large-scale overturn and angular wavelengths of con-
vective bubbles which are significantly larger than in a "clean" calculation. This
FIGURE 6. Top: Entropy distribution 208 ms after core bounce (in units of kB/nucl.) in a
two-dimensional supernova model showing odd-even decoupling. Bottom: Entropy distribution
for an equivalent calculation in which odd-even decoupling has been suppressed (from [8]).
artificial enhancement of convection is demonstrated in Fig. 6 where the entropy
distribution of a simulation exhibiting odd-even decoupling (top panel) is compared
to one in which the numerical failure has been cured (bottom panel). A modifi-
cation of PPM's original dissipation algorithms [3] was necessary for this purpose.
Alternatively the hybrid Riemann solver method of [14] might be used. Note that
the calculations have been carried out in spherical coordinates (r, θ) so that the
(initially spherical) shock wave was fully aligned with the grid. However, cylindri-
cal coordinates which have been used for plotting are indicated in the figures. The
difference in the final 56Ni yield for these two simulations was about 40%, the cal-
culation not exhibiting odd-even decoupling showing the larger yield, as expected
from our discussion in the previous section. This demonstrates that due to the
strong coupling of neutrino physics and convection, the 56Ni yield in multidimen-
sional calculations is much more difficult to calculate correctly than in one spatial
dimension. Therefore the error of at most a few percent which can be deduced
for the latter case from Fig. 4 can be deceptive. The results shown in Fig. 6 also
suggest that numerical noise, whatever its origin is, leads to a grossly overestimated
efficiency of convection. Thus, schemes which are known to suffer from this prob-
lem (like SPH) do not appear to be suited for nucleosynthesis calculations in core
collapse supernovae.
CONCLUSIONS
Realistic nucleosynthesis calculations in astrophysical contexts represent a chal-
lenge in many respects. The difficulties involve the numerical treatment of multidi-
mensional hydrodynamic advection, complex physics in addition to hydrodynamics
and burning, disparate length and time scales, realistic initial conditions and uncer-
tainties in reaction rates. It is our conviction, that substantial efforts are required
in each of these fields in order to obtain reliable yields in multidimensional nucle-
osynthesis calculations.
REFERENCES
1. Berger, M., and Colella, P., Jour. of Comp. Phys., 82, 64 (1989).
2. Bruenn S. W., Nuclear Physics in the Universe, Bristol: IOP, 1993, p.31
3. Colella, P., and Woodward, P. R., Jour. of Comp. Phys., 54, 174 (1984).
4. Fryxell B., Muller E., and Arnett W. D., MPA Preprint, 449, 1 (1989).
5. Fryxell B., Muller E., and Arnett W. D., ApJ, 367, 619 (1991).
6. Godunov S. K., Mat. Sb., 47, 271 (1959).
7. Janka H.-T., and Muller E., A&A, 306, 167 (1996).
8. Kifonidis K., PhD thesis, Technische Universitat Munchen, (2000).
9. LeVeque R. J., Computational Methods for Astrophysical Fluid Flow, Berlin:
Springer, 1998, pp. 84-101
10. Muller E., Computational Methods for Astrophysical Fluid Flow, Berlin: Springer,
1998, pp. 463-480
11. Oran, E. S., and Boris, J. P., Numerical Simulation of Reactive Flow, New York:
Elsevier, 1987, ch. 10, pp. 358-394.
12. Pen, U.-L., ApJS, 115, 19 (1998).
13. Plewa T., and Muller E., A&A, 342, 179 (1999).
14. Quirk J. J., Int. J. Num. Meth. Fluids, 18, 555 (1994).
15. Rampp M., and Janka H.-T., ApJ, 539, L33 (2000).
16. Reinecke M., Hillebrandt W., and Niemeyer J. C., A&A, 347, 739 (1999).
17. Thielemann F.-K., Nomoto K. I., and Hashimoto M., ApJ, 460, 408 (1996).
18. Woodward, P. R., ApJ, 207, 484 (1976).
19. Woodward, P. R., and Colella, P., Jour. of Comp. Phys., 54, 115 (1984).
20. Woosley S. E., Pinto P. A., and Ensman L., ApJ, 324, 466 (1988).
21. Woosley S. E., and Weaver T. A., ApJS, 101, 181 (1995).
|
astro-ph/9701146 | 1 | 9701 | 1997-01-20T20:30:45 | Quantitative morphology and color gradients of E+A galaxies in distant galaxy clusters | [
"astro-ph"
] | Narrow band multifilter photometry of the galaxy populations of four galaxy clusters at z=0.4-0.5 has provided us with a sample of 73 E+A galaxies which are secure cluster members. For 34 of them HST images are available (WFPC2). We have determined surface brigthness profiles and by means of an exponential or r^1/4 law fit we have classified the galaxies on the basis of their bulge to disk ratio. We have found that most of these galaxies are disk systems judging from the exponential nature of their profiles. We have also derived color gradients for 10 E+A galaxies in Cl0016+16 to study the spatial distribution of the starburst and to distinguish among the physical mechanisms at its origin. Color gradients indicate that most of the E+A galaxies in Cl0016+16 show little spatial variation in V-I colors (restframe U-V). Thus, our large and homogeneous sample of distant E+A galaxies confirms the claim that their enhanced star formation is a galaxy-wide phenomenon. | astro-ph | astro-ph |
Quantitative morphology and color gradients of
E+A galaxies in distant galaxy clusters
Paola Belloni
Universitatssternwarte, Scheinerstrasse 1, D -- 81679 Munchen, Germany
Three basic scenarios have been invocated to explain the sudden rise and
decline of star formation in distant clusters that lead to the transformation of
"active" galaxies into the elliptical− and S0−types which today dominate rich
clusters, i.e. the so called Butcher-Oemler effect, (Butcher & Oemler,1978). These
are: galaxy interactions or mergers with nearly equal mass neighbors (Lavery
& Henry, 1988), gas-rich field galaxies running for the first time into the hot
intercluster gas which ignites a brief but energetic episode of star formation
(Bothun & Dressler, 1986), and high speed close encounters of gas-rich galaxies
resulting in non-disruptive interactions or "galaxy harassment", (Moore, 1996).
Although neither of the proposed mechanisms for triggering star formation
has a definitive answer in its favour, the high resolution of the HST images of
distant clusters has recently provided a unique tool to face this question. Indeed,
it allows us to study the spatial distribution of the starburst and thus to distin-
guish among the physical mechanisms at its origin. In the case of an infalling
or harassed spiral the enhanced star formation would likely be a galaxy-wide
phenomenon, or confined to the disk. If, on the other hand, the original galaxy
was an elliptical that accreted gas from a dwarf galaxy, the burst signatures
should be detectable as a bluer color of the nuclear region. Finally, in interact-
ing or merging disk galaxies both behaviours have been observed: galaxies with
starburst concentrated to the very center (Scoville et al.,1991) and others that
show enhanced star formation on a galaxy-wide scale (Standford, 1991).
We focuse on one class of "active" galaxies, the post-starburst or E+A galax-
ies, the latter being only a description of the spectra which appear to have an
A-star component added to an old elliptical like component. These spectral fea-
tures are interpreted as evidence of a recent (< 1.5 Gyr) burst of star formation.
Ground based narrow-band photometry of 4 intermediate redshift clusters at
z=0.4-0.5 has provided us with a sample of 73 E+A galaxies being secure clus-
ter members (Belloni et al.,1995; Belloni & Roser, 1996). This is the largest
sample of such galaxies found to date in distant clusters. Cluster membership
and spectral type have been obtained by fitting the observed low-resolution spec-
tral energy distributions with template spectra built up with Bruzual & Charlot
(1997) population synthesis models. They represent the temporal evolution of
a strong star formation episode (involving 20% of the original galactic mass)
in an elliptical galaxy or in a spiral galaxy with star formation truncated after
the burst. Our approach allows us to detect bursts younger than 2 Gyr while
galaxies with an older burst will not be distinguished from a passively evolving
elliptical galaxy.
2
Paola Belloni
Due to the small HST field of view only for 33 E+A galaxies HST images
are available (WFPC2). We have retrieved them and determined: a) surface
brightness profiles for all E+A galaxies but those showing highly irregular
morphology. A fit with an exponential or a r−1/4 law has been performed and
for disk galaxies a Hubble type classification has been attempted on the basis
of the buldge to disk ratio, b) color gradients for the 10 E+A galaxies in
Cl0016+16 (observed in the F555W and F814W filters).
Table 1. Results of the surface brightness analysis of 33 E+A galaxies in intermediate
redshift clusters. The E+A fraction is given with respect to the secure cluster members,
about 120 galaxies brighter than mR=22.5 per cluster. For merging/interacting galaxies
we indicate if possible whether they appear disk or bulge dominated.
Cluster
E+A
E+A Morph. Morph. Morph.
Sb-Sc
(HST) Elliptical
S0-Sa
Merger
Interact.
1
Cl0016+16 20 (21%)
(z=0.54)
Cl0939+47 35 (22%)
(z=0.41)
Cl0303+17 22 (22%)
(z=0.41)
Cl1447+26
(z=0.38)
8 (9%)
11
11
7
4
3
1
1
6
7
5
1
1
Disk-dominated
4
1
Irregular
2
Disk-dominated
We found that most of the galaxies and all those showing signs of
interaction are disk systems judging from the exponential nature of the pro-
files (Tab.1). In our sample only one E+A galaxy has a regular elliptical profile.
We also found that the E+A galaxies in Cl0016+16 show little spatial varia-
tion in V-I colors (restframe U-V) (Fig.1). Only one young E+A galaxy with
estimated post-starburst age 0.8 Gyr has a starburst predominantly located in
the nucleus whereas a second one has a weaker evidence of the same behaviour.
However, the E+A color gradients are flatter compared to those of cluster ellip-
ticals (Belloni et al, 1997) that are consistent with the ∆(U-R)/∆log r=−0.23
mag observed in the nearby ones (Franx et al.,1993). Thus, our large and homo-
geneous sample of distant E+A galaxies confirms the claim that their enhanced
star formation is a galaxy-wide phenomenon. Indeed, similar results have
been obtained for the post-starburst galaxies in the Coma cluster by Caldwell et
al. (1996). We speculate that strong interactions did not create most of the post-
starburst galaxies in our representative sample of galaxy clusters at z=0.4−0.5.
Non-disruptive interactions among galaxies or interactions between galaxies in
Color gradients of E+A galaxies in distant galaxy clusters
3
the cluster or subcluster structures and the ICM are likely to be the dominant
mechanism.
3
2.5
2
1.5
1
3
2.5
2
1.5
1
0.2
0.4
0.6
0.8
1
0.2
0.4
0.6
0.8
1
Fig. 1. a) Typical color gradients of E+A galaxies in CL0016+16 (z=0.54). b) One
of the two E+A galaxies in Cl0016+16 showing strong color gradients. The very blue
central colors (typical of spirals) indicate that the starburst is located in the nucleus.
1′′= 7.3 Kpc (Ho=50 km sec−1 Mpc−1 and q0=0)
References
Belloni, P., Bruzual G., Thimm, G., Roser H.-J., (1995): A&A 297,61
Belloni, P. & Roser (1996): A&AS 118, 65
Belloni, P. et al. (1997): ApJL, submitted
Butcher, H. & Oemler, A. (1978): ApJ 219, 18
Bothun, G. & Dressler A. (1986): AJ 301, 57
Caldwell et al. (1996): AJ.111, 78
Franx, M., Illingworth, G., Heckman,.T. (1993): AJ.98, 2
Lavery, R. & Henry P. (1988): ApJ 330, 596
Moore, et al. (1996): Nature 379, 613.
Scoville, N. et al. (1991): ApJ. 336, L5
Standford, S., (1991): ApJ. 381, 409
|
astro-ph/9811231 | 1 | 9811 | 1998-11-14T02:54:41 | The Survival of The Core Fundamental Plane Against Galactic Mergers | [
"astro-ph"
] | We have developed an effective simulation method and applied it to the problem of the accretion of very dense secondary companions by tenuous primaries. We have studied the accretion of objects of varying luminosity ratios, with sizes and densities drawn from the core Fundamental Plane for elliptical galaxies under the assumption that the mass distribution in the central parts of the galaxies follows the light. The results indicate that in mergers with mass ratios greater than 10, the smaller object is only stripped down to the highest density encountered in the primary during the accretion process. Thus, the form of the core Fundamental Plane suggests that the mass distribution in these galaxy centers is different than the light distribution, or that mergers on the core Fundamental Plane require more than the dynamics of visible stars. | astro-ph | astro-ph | The Survival of The Core Fundamental Plane Against Galactic
Mergers
Kelly Holley-Bockelmann and Douglas Richstone 1
Astronomy Department, University of Michigan
Received
;
accepted
8
9
9
1
v
o
N
4
1
1
v
1
3
2
1
1
8
9
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
1and Institute for Advanced Study, Princeton, NJ 08540
-- 2 --
ABSTRACT
The basic dimensional properties of the centers of elliptical galaxies, such
as length scale, luminosity, and velocity dispersion, lie on a Fundamental
Plane similar to that of elliptical galaxies as a whole. The orientation of this
plane, and the distribution of core parameters within it, project to a strong
correlation of core density with either core or total luminosity, and indicate that
low luminosity ellipticals are much denser than high luminosity galaxies (HST
data suggests that this relationship may be as steep as ρc ∝ L−2). In addition,
low luminosity ellipticals have a much smaller length scale than their high
luminosity counterparts. Since we think small galaxies are occasionally accreted
by big ones, the high density of these galaxies and their likely durability against
the time-varying tidal field of the bigger ones suggests that they will survive
substantially intact in the cores of larger galaxies and would be easily visible.
Their presence would destroy the observed correlation.
Motivated by this apparent inconsistency between an observed fact and a
simple physical argument, we have developed an effective simulation method and
applied it to the problem of the accretion of very dense secondary companions
by tenuous primaries. We have studied the accretion of objects of varying
luminosity ratios, with sizes and densities drawn from the Fundamental Plane
under the assumption that the mass distribution in the central parts of the
galaxies follows the light. The results indicate that in mergers with mass ratios
greater than 10, chosen with an appropriate central density dependence on
luminosity, the smaller object is only stripped down to the highest density
encountered in the primary during the accretion process. Thus, the form of
the core Fundamental Plane suggests that the mass distribution in the galaxy
centers is different than the light distribution, or that an understanding of
-- 3 --
secondary survival requires more than the dynamics of visible stars.
Subject headings: stellar dynamics - galaxies: kinematics and dynamics -galaxies:
evolution - galaxies: clusters - galaxies: nucleii - galaxies: elliptical
-- 4 --
1.
Introduction
In the framework of hierarchical galaxy formation, massive elliptical galaxy centers are
merger archives, in the sense that the history of each component is imprinted upon it.
The stellar Roche limit suggests that the accretion of dense galaxies by larger, more
tenuous ones would leave dense galaxies nearly intact. Hence, the central density of any
massive galaxy should vary widely as a function of the extent of prior merging. Instead,
the global parameters of elliptical galaxies vary simply with the luminosity. Elliptical
galaxies exist in a well-defined slice of global parameter space called the Fundamental Plane
(Dressler et al. 1987, Djorgovski and Davis 1987). Similarly, the central parameters of
galaxies are also arranged on a plane called the core Fundamental Plane (Faber 1997).
On the core Fundamental Plane, small galaxies can be 106 times more dense than large
galaxies (Faber, 1997). This enormous dynamic range in density presents a formidable
challenge for hierarchical merging because in this theory, the dense center that results from
the nearly intact accretion of a small galaxy by a bigger one will cause the remnant to
evolve off the Fundamental Plane. Consequently, the persistence of the Fundamental Plane
is an indication either that mergers are uncommon or that additional forces are at work to
destroy the secondary.
There is considerable evidence that mergers do occur. Observations of brightest cluster
galaxies indicate that massive ellipticals evolve by accretion of small satellites at a rate of
up to 0.2L⋆ per Gyr (Lauer 1988, Blakesee and Tonry 1992). Simulations of galaxy cluster
evolution also demonstrate that mergers occur, although the importance of this process
in the creation of cD galaxies has been debated (Malumuth and Richstone, 1983; Merritt,
1984). So, while it is clear from the persistence of the Fundamental Plane that mergers
must destroy the smaller satellites, it is not clear how such dense systems are destroyed.
-- 5 --
Recently, Weinberg (1997) has argued that internal heating of the secondary can
cause it to evaporate well inside the tidal radius. This heating occurs when some of the
orbits of stars in the secondary resonate with the time-dependent tidal forces imposed
by the primary. To demonstrate this, Weinberg perturbs a solution of the Boltzmann
equation and recomputes the gravitational field of the secondary for King models on the
global Fundamental Plane. He states that a satellite with a mass ratio as high as 100:1
will dissipate entirely by the time it falls to the primary center (Weinberg, 1997). As we
will show later, although we agree with Weinberg's result for the specific problem he did,
his models lack the density contrast observed for Fundamental Plane galaxy cores, so his
simulations have limited application to the problem of core Fundamental Plane persistence.
Mergers of galaxies with very different densities are expensive to simulate with a
straightforward N-body approach due to the disparity in two important timescales. First,
we must consider the orbital decay of the secondary into the primary, which is of order of
the dynamical friction timescale, and varies as M1/M2 ρ−1/2
orbital periods of stars within the secondary, which vary as ρ−1/2
1
. We must also consider the
2
where ρ1 and ρ2 are the
densities of the primary and secondary. So, even for a modest mass ratio, tracking the stars
in each galaxy over the entire merger requires on the the order of 100 M1/M2 (ρ1/ρ2)−1/2
timesteps. There are few published direct simulations of this type of merger for this reason.
In one important case, however, Balcells and Quinn (1990) did simulate an elliptical galaxy
merger with a mass ratio of 10:1 (and density ratio of approximately 1:14). Contrary to
Weinberg's result, the secondary remained intact inside its tidal radius as it sank to the
center of the primary, although the orbital decay was perhaps too fast to produce resonant
heating.
We have designed an approach to determine the conditions under which core
Fundamental Plane secondaries survive mergers. This approach can efficiently merge
-- 6 --
systems with mass ratios as high as 1000, a range unexplored by standard simulations. In
this paper, we describe this method and apply it to the problem of secondary survival. We
review the core Fundamental Plane in § 2, our galaxy models in § 3.1, the approximation
method in § 3.3, and our code in § 3.4. The tests of the method and results of our
simulations can be found in § 4. Briefly, it appears that core Fundamental Plane secondaries
survive mergers in which the mass ratio is 10:1 or larger. Section 5 discusses the implications
of our results on the persistence of the core Fundamental Plane and previews future work.
2. The Core Fundamental Plane
In the parameter space composed of the effective radius, re, the effective surface
brightness, µe, and the central velocity dispersion, σ0, elliptical galaxies are confined to a
plane -- the "Fundamental Plane" or FP (Dressler et al. (1987), Djorgovski & Davis, 1987).
Recently Jorgensen et al. (1996) characterized the FP as log re = 1.24 log σ0 − 0.82Ie + γ,
where γ may depend on environment. A popular interpretation of the Fundamental
Plane is that it is a manifestation of virial equilibrium together with an orderly and slow
dependence of galaxy mass -- to -- light ratio on luminosity. Faber et al. (1997) report the
existence of an analogous plane for elliptical galaxy centers based on core parameters,
though this core Fundamental Plane has more scatter than the global Fundamental Plane.
The existence of such a plane indicates that the processes which act upon galaxies are
primarily scale-independent. More importantly, this plane provides an important constraint
on the merger history of elliptical galaxies, because if a galaxy merges, the remnant must
also lie upon the plane. In particular, this implies that when a small dense galaxy collides
with a massive galaxy, the remnant must have the low density center that is characteristic
of a massive primary.
When simulating the mergers of real galaxies, then, it is imperative that the initial
-- 7 --
models lie on both planes. Although the Fundamental plane is 2-dimensional, we adopt
the convention of Faber (1997) and represent the core galaxies as a 1-dimensional family
dependent on total luminosity. This is a convenient approximation which is made possible
by the tight correlation between global and break luminosity and by the nearly edge-on
projection of the plane onto these axes. The core data can be fit as follows:
log rb = −0.46 Mv − 7.7
log Lb = 1.3 log Le − 5.097
log reff = 0.75 log rb + 2.06,
(1)
(2)
(3)
where rb is the break radius, Mv is the absolute visual magnitude, Lb is the visual
luminosity at the break radius, Le is the visual luminosity at the effective radius, and
ref f is the effective radius. The following additional relations were obtained (Faber et al.
1997) by assuming a slightly varying mass-to-light ratio along the Fundamental Plane of
M/L ∝ L0.25
,
e
e
σ0 ∝ L0.2
0/r2
ρb ∝ σ2
b ∝ L−1.9
e
.
(4)
(5)
This last equation is the crux of the difference between our simulations and Weinberg's.
Weinberg's expression of the Fundamental plane is ρb ∝ M 0.5, where M is the mass of the
galaxy, so for reasonable choices of M/L, his Fundamental Plane has a much shallower
dependence on density with radius. See figure 1 for best-fit projections of the core
Fundamental Plane to the Lauer et al. (1995) core data.
3. Methods
-- 8 --
3.1. Modeling the Core Fundamental Plane
The disruption of the secondary galaxy centers will depend on the density ratio between
the primary and the secondary, and on their density profiles. We therefore construct models
designed to reflect the range in profiles observed on the core Fundamental Plane. We use η
models for this purpose (Tremaine et al. 1994, Dehnen, 1994). These are simple, spherical
systems defined by their density distributions:
ρη(r) =
Mtot
3
rb
∼
ρ (η, r/rb),
where
∼
ρ (η, x) ≡
1
4π
η
x3−η(1 + x)1+η ,
0 < η ≤ 3,
(6)
(7)
where η is the single dimensionless parameter that defines the slope of the inner density
profile, ρ is the stellar density, Mtot is the total mass, rb is the break radius, and x is the
dimensionless radius x = r/rb.
We fit the the deprojected stellar density profiles provided by Gebhardt et al. (1996)
to determine η as a function of the absolute magnitude Mv. The data (see Figure 1) show a
correlation of the structural parameter η with Mv. The ridgeline of the correlation was fit
by eye as follows:
3.0 − 0.375Mv − 9.1,
3.0 − 0.15Mv − 4.7,
if Mv ≤ −19.5;
if Mv − 19.5
.
(8)
η =
While eta models are well-suited to cover the range of luminous matter densities in
elliptical galaxy centers, the model's density profile drops off more rapidly (ρ ∝ r−4) than a
real galaxy's at large radii. Although we anticipate that the envelope of a secondary would
be completely stripped in an encounter with a primary, and is therefore irrelevant, we were
concerned that truncating the envelope might have an unforseen effect on our calculations.
So, we mimicked an r1/4 law outer profile by constructing a basic set of galaxies with eta
-- 9 --
model envelopes. We call these galaxies double η models. Their density is defined as follows:
ρ(r) =
Mc
3
rc
∼
ρ (ηc, r/rc) +
Me
3
re
∼
ρ (ηe, r/re),
(9)
where the subscript c and e corresponds to eta model parameters chosen at the core and
effective radius, respectively. The core radius in equation 9 is chosen to be mimic the break
radius on the core Fundamental Plane. The envelope term parameters are chosen to be
consistent with global Fundamental Plane parameters and with re ∝ rc
(Faber, 1997). For re on the order of 100 rc, ρ(r) provides a reasonable approximation to a
1.33 (see equation 3)
deVaucoleur profile.
3.2. Populating the Galaxies
To populate our galaxies, we require the phase space distribution function. From our
spherically symmetric initial density profile, we can derive the initial distribution function,
f (ε), via Eddington's formula:
f (ε) =
1
√8π2
Z ε
0
d2ρ
dΨ2
+
dΨ
q(ε − Ψ)
1
√ε dρ
dΨ!Ψ=0
,
(10)
where ε = −E + Φ0 is the relative energy, and Ψ = −Φ + Φ0 is the relative potential
(Binney and Tremaine, 1987). Φ0 is chosen so that the distribution function is positive for
ε > 0 and zero for ε ≤ 0. The second term in equation 12 is zero, because Ψ = 0 when
r → ∞, and as r → ∞, ρ ∝ r−4 ∝ Ψ4. Determining f (ε), then, simply requires knowledge
of d2ρ/dΨ2.
Defining u = 1/r,
d2ρ
dΨ2 =
d
du dρ
dΨ!(cid:30) dΨ
du
.
(11)
-- 10 --
For unscaled η models,
d
du" dρ
dΨ# = "12u2 + 4(3 − η)u3
z
4u3 + (3 − η)u4 z ′
z2
#,
−
where z = (1 + u)2, and z ′ = dz/du.
(1 + u)−η,
1/(1 + u),
if η 6= 1.0,
if η = 1.0
.
dΨ
du
=
(12)
(13)
See Tremaine et al, 1994 for an equivalent expression of the second derivative. The second
derivative for the double η model case has many more terms, but follows directly from the
above derivation. Given the distribution function, we know the number of particles within
the phase space d3v d3x. To populate the secondary, then, we choose a random scalar radius
from the mass profile, and determine the speed by choosing a random scalar velocity from:
f (v)dv = 4πv2f (Ψ −
1
2
v2)dv.
(14)
Finally, cartesian position and velocity components are derived from uncorrelated random
orientations of r and v.
These two component galaxies were simulated with 5000 particles. In these simulations,
the outer envelope was stripped away on the first encounter with the center of the primary,
and we took advantage of this result by conducting additional simulations of the inner
portions of the galaxies with an inner eta model of 2000 particles. This markedly improved
our spatial resolution. These simulations contain a reasonable particle number, but in
principle, the simulations can and will be repeated with more particles as computational
speed increases. Model parameters are given in Table 1. Figure 1 illustrates the placement
of the models on the core Fundamental Plane.
-- 11 --
3.3. The Tidal Force on the Secondary
A straightforward technique to simulate mergers is to follow the interaction of the
primary and secondary with an N-body code. However, when the galaxies involved have
very different masses, the interaction takes an unreasonably long time to simulate in
this straightforward manner, as described in the introduction. We have developed an
approximation method which dramatically decreases the computational time required to
simulate high mass contrast mergers. Our approximation hinges on the assumption that
the potential of a high mass galaxy is not significantly altered by the impact of a low mass
object. This approximation allows us to concentrate on the behavior of the secondary only.
The merger, then, can be characterized as the secondary's response to the external force it
experiences as its orbit decays in the field of the primary.
Clearly, the external force on the secondary depends upon its position in the primary
galaxy. We determine the orbital decay trajectory by solving the equations of motion for a
point particle in the field of a galaxy of mass M1:
~F1−2(~S) = −
GM1(~S)m2
~S2
− Ffric(~S),
(15)
where ~S is the vector from the center of the primary and to the secondary center, and
Ffric is a modified Chandrasekhar dynamical friction formula. This also presupposes an
unchanging primary, and neglects the effect of mass loss in the secondary.
Ffric(~S) = −fdrag
4πlnΛG2ρ1m2
2
3
~v2
herf(X) −
where ~v2 is the velocity of the secondary's center of mass, X ≡ v2/(√2σ),
σ = q0.4GM1/r1,eff, Λ is the Coulomb logarithm which was set to M1/M2, and
fdrag is a drag coefficient. This drag coefficient is necessary, because the decay times
2X
√π
e−X 2i~v2,
(16)
-- 12 --
generated by numerical treecode experiments were longer and exhibited more pericenter
passes than the unmodified Chandrasekhar formula. Since the number of pericenter passes
is correlated with the amount of destruction in a particular secondary, we chose to match
the number of pericenter passes and the overall decay time by adjusting the analytical
dynamical friction force by drag coefficient fdrag. Figure 3 shows the treecode-generated
trajectory and several analytically-derived trajectories with different drag coefficients.
Once the orbital decay trajectory is computed, the external force on particles within
the secondary can be written as:
~FS(~r) = −
GM1(~S + ~r)m2
(~S + ~r + ǫ)3
,
(17)
where ~r is the vector which points from the secondary center to a secondary particle, and ǫ
is a softening parameter which we chose to be close to the core radius of the secondary. We
advance the system in the secondary frame, so the appropriate force is the sum of the tidal
force and the secondary's self-gravity:
~Ftot(~r) = ~Fin(~r) + ~FS(~r) − ~FS(0).
(18)
We tested the assumption that the primary is unchanged by the secondary for each
mass ratio by dropping a secondary point mass into a 10000 particle primary using either
Hernquist's treecode (1987) or self-consistent field code (1992). The density profile of the
primary did not change for the 100:1 and 10:1 cases (see Figure 2), so the approximation
is a valid one for those mass ratios. This method is more computationally efficient as the
mass ratio increases, and is over 50 times faster than the Hernquist treecode in the 10:1
mass ratio.
-- 13 --
3.4. Approximating the Self-Gravity of the Secondary
We constructed a self-consistent particle-field code which treats the mass distribution
as a sphere and uses a logarithmic grid. This force is estimated by first convolving the mass
distribution with an adaptive kernel function (Silverman, 1986) to get a density estimate.
An adaptive kernel provides a non-parametric density estimate on a grid. We use an
Epanechnikov kernel (an inverted parabola) for the density estimate:
3
4(1−r2)
if r < 1 ,
0
otherwise,
(19)
Ke =
where r = x − xi/h, x is a particle position, xi is a grid point, and h is the window width.
An initial density estimate is obtained, then the window width of each grid element is
adjusted according to the initial density such that low density regions have a large window
width, and high density regions have greater mass resolution. This gives us a smoothed
M(r) from which force and potential are easily computed (Merritt, 1996). Although it is
true that this code, which forces spherical symmetry, is not an appropriate choice if one
wishes to model the shape of the highly non-spherical process of satellite disruption, we
were concerned about whether a particle was bound to the secondary, not with obtaining an
accurate shape for the stripped debris. Furthermore, since the tidal force is a very strong
function of r, the core of the secondary, where the calculation matters most, is spherically
symmetric.
Particle trajectories are advanced in the secondary frame with a single timestep
leapfrog method.
We tested this code by following the total energy of each secondary without the
influence of an external force. For double η secondaries, energy was conserved to within 1
% over 1700 core crossing times, the time in which the 100:1 mass ratio case would have
decayed had an external force been present. The mean energy and root mean square energy
-- 14 --
of the individual stars were conserved to within 2 % over this interval. Inner η models
exhibited more heating, initially, because the kernel inaccurately determined the density
very near the center of the system. Since inner η models have a better resolution, there
were a small number of particles that were close enough to the center of the secondary for
this inaccuracy to matter. We restored energy conservation by setting the mass distribution
to that of the inner η model inside 0.01rb, and reflecting any particles entering this sphere
back outward with the same energy and angular momentum. These modifications to the
code affected approximately 10 particles, but markedly improved the energy efficiency. See
figure 4 for δEtot vs. crossing time.
3.5.
Initial Conditions
Each secondary was followed as an isolated system for several dynamical times to
ensure its virial equilibrium. The merger simulation begins as the secondary is launched
from 3rhalf of the primary with an initial velocity that depends upon the orbit desired;
for a plunging orbit, we launch the secondary with zero velocity. We define an angular
momentum parameter κ ≡ L/Lcirc, where L is the initial orbital angular momentum,
and Lcirc is the orbital angular momentum that the secondary would have if it were on a
circular orbit about the primary at the launch point. Each simulation is followed until the
secondary's apocenter is less than 0.1 rb of the secondary.
4. Results
Since our goal is to investigate the breakup or survival of dense secondaries orbiting in
primaries, we initially focused on the encounters we thought would be the most destructive.
The destruction of the secondary, however, must occur without leaving a lasting imprint
-- 15 --
on the remnant, and this depends on where in the primary a secondary is stripped. For
example, in a decaying circular orbit, the tidal force from the primary would gain strength
as the secondary's orbit decays. In this case, even if substantial damage occurred, the debris
would continue to orbit near the center of the primary, and the density of the remnant
would be large near the center. The orbits most effective for dispersing the secondary
over a large volume must be the deeply plunging ones, since these encounters would carry
the secondary past the center of the primary on its first orbit with considerable orbital
energy. Our basic set of experiments, therefore, were parabolic encounters with no angular
momentum. For an overview of the different experiments conducted, see table 2.
4.1. Basic Set of Experiments
Anticipating that the most damaging effect would be due to tidal forces experienced
as the secondary passes through the center of the primary, we launched a series of
nearly parabolic, plunging orbits. In this basic set of experiments, we investigated three
mass ratios: 100:1, 10:1, and 2.5:1. These mass ratios correspond to density ratios of
approximately 1:51, 1:20, and 1:1 at a radius of 1 pc, and the density ratio increases at
smaller radii. The secondary was modeled as a double eta model in order to represent
the envelope. We focus first on the 100:1 simulation. Figure 5 shows the time evolution
of the secondary in the force field of the primary. Here, we learn that most of the mass
is stripped at pericenter, and comes off impulsively in a cloud which continues to expand
as the secondary crosses the primary center again. In figure 6, we illustrate the change in
the secondary center from beginning to end. The secondary remained intact inside the
3 ≈ M1/M2 S3. Figure 7 illustrates the change in the secondary for the
10:1 mass ratio. In the 2.5:1 experiment, the secondary and primary have similar central
tidal radius, rtide
densities, and the secondary is destroyed (figure 8).
-- 16 --
We note that the primary and secondary in the 2.5:1 mass ratio are close enough in
mass for the approximation scheme to fail, because the primary is not expected to remain
unaffected by a merger of similar mass. We address this fact by conducting a treecode
experiment, which we discuss in § 4.5.
4.2. Comparison of Drag Coefficients
Recall the drag coefficient, fdrag in § 3.3 which adjusts the strength of Chandrasekhar's
dynamical friction force to be comparable to treecode generated orbital decay times.
Changing the strength of the frictional force, though, also changes the velocity of the
secondary, and therefore both the magnitude of the shock on the secondary as it passes
through pericenter and the number of pericenter passes. We tested the importance of
this effect by simply varying the drag coefficient on the 100:1 fiducial experiment. In
2 simulations with fdrag = 0.1 and fdrag = 0.2, we saw no significant difference in the
secondary's inner density profile, although it is apparent that the secondary suffers more
tidal stripping as it goes through more pericenter passes (see figure 9). Consequently, we set
fdrag to the value determined by a treecode simulation of a point secondary in a responsive
primary for each mass ratio in all further experiments. This provided a more accurate
timescale for the decay, but did not affect the survival of the secondary.
4.3. Comparison between Inner η and Double η Models
To achieve better sampling of the central regions of the secondary, we populated only
the center of the galaxy with an inner η model. While this method excludes the secondary
envelope, it is a reasonable plan because our basic set of experiments determined that the
envelope is stripped away quickly without markedly changing the density profile at the
-- 17 --
center (see § 4.1 and figures 6-8). Note that the orbital decay due to dynamical friction is
the same as the previous experiment. We only followed the stars in the inner part of the
galaxy but used the total mass of the secondary to determine the orbital decay.
As in the fiducial set of experiments, these simulations show that the secondary remains
intact in the 100:1 and 10:1 mass ratio, and that the secondary breaks up in the 2.5:1 mass
ratio (figures 10-12).
4.4. Non-radial Secondary Orbits
To explore the effect that different orbits have upon the disruption of the secondary, we
launched secondaries in the 10:1 mass ratios on κ ≡ L/Lcirc = 0.2 and κ = 0.6 orbits, and
the 100:1 mass ratio on a κ = 0.2 orbit. The ratio of second apocenter to first pericenter is
approximately 66 in the 100:1, κ = 0.2 orbit, and 34 and 16 for the 10:1, κ = 0.2 and κ = 0.6
orbits, respectively. The computation time for these simulations required us to use inner η
models and to follow only the bound particles. Still, the κ = 0.6 run took approximately
2.5 months. In every case, the secondary's innermost density profile remained essentially
intact; the secondary center was steeper and more dense than the primary, so the center of
the remnant is dominated by the secondary (Figures 13-15). For a comparison of the effect
that orbit has upon the disruption of the secondary, we overplot the final secondary profile
for each orbit in the 100:1 and 10:1 inner η experiments (figure 16).
4.5. Treecode Comparison
Our method works best at high mass contrasts, because for a given primary, the
lower the secondary mass, the less the primary is affected by a merger. In addition, the
gravitational force on the secondary is better approximated as a tide when it is small
-- 18 --
compared to the self-gravity of the secondary, as it is for a small, dense, secondary. On the
other hand, the treecode is superior at low mass-contrasts because it can follow changes in
the primary. We compare the results from these two problems at a mass ratio of 2.5:1 We
launched a 2000 particle double η secondary into a 5000 particle double η primary along
the same orbit as described in the fiducial experiments and followed the behavior with the
Hernquist treecode.
In the treecode simulation the primary center is highly disrupted by the secondary.
Our method cannot reproduce this effect. In spite of the marked difference between the
behavior of the primary in each method, the qualitative behavior of the secondary is the
same: the secondary central density decreases to that of the primary (figure 17). The
fact that we obtain the same qualitative results provides some confidence in the method.
In addition, the fact that our method fails to address the change in the potential of the
primary in a clear warning that the method is quantitatively unreliable when the mass
ratios are very close to 1.
4.6. Comparison to Weinberg's Results
A key test of our method is to duplicate Weinberg's experiment using our approximation
method. Weinberg chose to represent the Fundamental Plane with King model galaxies
(King, 1966).
King models are a single parameter set of solutions of the Vlasov equation, which can
be labeled by the dimensionless potential W0 = −V0/(2σ2) (where V0 and σ are the central
potential and dispersion parameter), or equivalently by the ratio of "tidal" to core radii
rt/rc. The mass distribution in a King model is given by M(r) = µ(r/rc)ρ0r3
c , where ρ0 is
the central density and µ is a dimensionless function. Weinberg adopted a single choice of
-- 19 --
W0 in his experiments, so his models all had the same dimensionless mass distribution but
scaled up or down in both density and length scale.
We chose to duplicate one of Weinberg's 10:1 mass ratio simulations. In this experiment,
he populated two W0 = 9.5 King models with densities defined by the global Fundamental
Plane: ρ ∝ M −0.5. He determined the orbital decay with Chandrasekhar's analytical
dynamical friction formula, using the Coulomb logarithm ln Λ = max[ln(Rcirc/rhalf), 0.1],
which caused the orbit to decay in 10 orbits for a choice of κ = 0.1. We prefer our handling
of the orbital dynamical friction because it is calibrated by the treecode. However, we
wish to critically compare our calculation of the effect of the fluctuating tide to Weinberg's
problem, so we selected a choice of fdrag such that our orbit also decayed in 10 passes for
κ = 0.1. In all dimensionless characteristics, our experimental set up is the same. See table
3 for the parameters used in this experiment. 2
Weinberg contends that the secondary is disrupted by the time it reaches the center,
with most of the disruption due to stripping near the primary center. In this particular
case, we obtain the same result (figure 18); that is, the secondary is stripped by nearly a
factor of 10, and ρ2 = ρ1. This shows is that when we do exactly the same experiment
as Weinberg, we get the same result. Therefore, the differences between our results and
Weinberg's are not due to differences in calculation methods. Instead, we believe the
differences are due to different choices of galaxy models and the use of these models to
mimic the Fundamental Plane. In our paper, we carefully selected each galaxy to fall on
2 There is a typographical error in the Weinberg preprint that causes in an inconsistency
between the King model concentration and the tidal radius; this prevented us from taking
Weinberg's precise radius in parsecs and his Coulomb logarithm. We circumvented this
problem by creating two W0 = 9.5 King models that differed in mass by 10 and density by
a factor of 3, consistent with Weinberg's model.
-- 20 --
the global and core Fundamental Planes, assuming light follows mass within the galaxies.
5. Discussion
We have developed an approximation scheme that allows us to investigate the mergers
of systems with high density ratios. We can use this method to simulate merging systems
with central density contrasts as high as 1500 with less than a month of computer time on
an UltraSparc.
Our simulations show that disruption will not occur for secondaries that are much
denser than their primaries. While resonant heating may enhance tidal shocking in the
disruption of secondary galaxies, this only appears to matter when the density contrast
between primary and secondary is not too large. Weinberg's King model secondaries
evaporated because they do not possess the high densities that real secondaries exhibit on
the core Fundamental Plane.
This result presents a quandary. We suspect, from observations of merger rates, that
dense secondaries do merge with more diffuse primaries. In the process, the secondary
must be destroyed, since observations of the centers of massive merger remnants rarely
show dense, small secondaries. Instead, nearly every large remnant has a low-density center
with a density profile that flattens out inside a break radius; this is a key aspect of the
core Fundamental Plane. The inability of purely stellar mergers to destroy the secondaries
suggests that an important component of the problem is missing.
One possible solution to this quandary may be massive black holes. The widespread
presence of massive black holes in galaxy centers has been suggested, for example, by
Kormendy & Richstone (1995), Magorrian (in press), Richstone (in press). Black holes act
as excellent scatterers, and their presence in the centers of massive primaries may provide
-- 21 --
tidal forces sufficient to disrupt a dense secondary. We will present extensive simulations
of the effect of black holes in a second paper. However, alternative possibilities include
radial variations of M/L within individual galaxies, with an amplitude strongly dependent
on galaxy luminosity, in the sense that luminous galaxies possess sharp concentrations
of unseen mass. In order to be effective, the mass concentration would have to be more
dramatic than the M/L variation with luminosity implied by the core Fundamental Plane.
Support for this work was provided by the Space Science Telescope Institute, through
general observer grant GO-06099.05-94A, and by NASA through a theory grant G-NAG5-
2758. DR thanks the Guggenheim Foundation and the Ambrose Monell Foundation for
support at the IAS. We thank the members of the NUKER collaboration for helpful
conversations. We also thank the referee, Martin Weinberg, for his insightful suggestions.
-- 22 --
REFERENCES
Balcells, M., Quinn, P. J. 1990, ApJ, 361, 381-393
Barnes, J., Hernquist, L. 1991, ApJL, 370, L65
Binney, J., & Tremaine, S. 1987, Galactic Dynamics (Princeton University Press)
Blakesee, J., Tonry, J. 1992, AJ, 104, 599
Busarello, G., Capaccioli, M., Capozziello, S., Longo, G., Puddu, E. 1997, A&A, 320, 415B
Dehnen, W. 1993, MNRAS, 265, 250
Djorgovski, S, & Davis, M., 1987, ApJ, 313, 59
Dressler, A., et. al. 1987, ApJ, 313, 42
Faber, S., Jackson, R. 1976, ApJ, 204, 668F
Faber, S. 1996, private communication
Faber, S., Kormendy, J., Byun, Youg-Ik, Dressler, A., Grillmair, C., Lauer, T., Richstone,
D., Gebhardt, K., Tremaine, S. 1997, ApJ, in press
Hernquist, L., Ostriker, J. 1992, ApJ, 386, 375-397
Hernquist, L. 1987, ApJS, 64, 715
Hernquist, L. 1992, ApJ, 400, 460
Jorgensen, I., Franx, M., Kjaergaard, P. 1996, MNRAS, 280, 167J
King, I. 1966, ApJ, 71, 64
Kormendy, J. 1985, ApJL, 292, L9
Kormendy, J. & Richstone D. 1995, Ann Rev Astron and Astroph, 33, 581
Lauer, T. R. 1985a, ApJS, 57, 473
Lauer, T. R. 1985b, MNRAS, 216, 429
-- 23 --
Lauer, T.R., Ajhar, E., Byun, Yong-Ik, Dressler, A., Faber, S., Grillmair, C., Kormendy, J.,
Richstone, D., Tremaine, S. 1995, AJ, 110, 2622
Magorrian, J. 1998, AJ, 115, In press
Merritt, D. 1996, AJ, 111, 2462
Mihos, J., Hernquist, L. 1994, ApJ Letters, 437, L47
Richstone D. 1998, Nature, In press
Silverman, B.W., 1986, Density Estimation for Statistics and Data Analysis (New York:
Chapman and Hall)
Tremaine, S., Richstone, D., Byun, Youg-Ik, Dressler, A., Faber, S., Grillmair, C.,
Kormendy, J., Lauer, T. 1994, AJ, 107, 634
Weinberg, M. 1997, ApJ, 478, 435
Weinberg, M. 1994a, ApJ, 108, 1398-1402
Weinberg, M. 1994b, ApJ, 108, 1403-1413
Weinberg, M. 1994c, ApJ, 108, 1414-1420
This manuscript was prepared with the AAS LATEX macros v4.0.
-- 24 --
Fig. 1. -- A comparison of data and our model of core parameters of galaxies.
In each
panel, the Lauer et al. (1995) galaxy core data are plotted as squares, and a fit of the core
and the power-law data is plotted as a line. Panel A shows estimated values of η for the
Lauer et al. sample, together with our adopted model. The 5 X's in panel A correspond
to the 5 unique galaxies in Table 1. The other three panels represent the projections of the
core Fundamental Plane for these galaxies. Panel B is a plot of core radius versus absolute
visual magnitude for the Lauer et al data and our model.
In this panel, the stellated X
demonstrates the core radius that Weinberg's 1:100 mass ratio secondary would have if the
primary were to lie on the core Fundamental Plane at Mv = −22. This illustrates that
Weinberg's simulated Fundamental Plane has a shallower slope, as seen by the dashed line.
Panel C plots resolution-limited surface brightness versus absolute visual magnitude of the
Lauer et al sample, with the central surface brightness and absolute visual magnitude of
our models. Panel D plots the resolution limited velocity dispersion versus absolute visual
magnitude for the data, and the central velocity dispersion versus absolute visual magnitude
of our models.
Fig. 2. -- The change in the primary during a 1:100 merger. The density of the primary is
shown as a function of radius before and after the merger. The initial density of the primary
is a dashed line, and the final density is a solid line. The changes are fluctuations due to
small number statistics.
Fig. 3. -- The distance between primary and secondary centers versus secondary crossing
time. We plot the secondary decay trajectory from a treecode experiment as a solid line,
and the analytical dynamical friction derived decay trajectories from two experiments with
varying drag coefficients. The short dashed line represents fdrag = 1.0, and the long dashed
line has fdrag = 0.2.
Fig. 4. -- Test of the method: energy conservation. We plot the total energy (normalized
-- 25 --
so that the mean energy is 1.0) versus the core crossing time of the secondary for the inner
η model secondary, and for the double η model secondary. Energy is constant to within
2 percent for both simulations, although energy conservation is better for the double η
secondary after it relaxes.
Fig. 5. -- The projection of the xy plane of a secondary as it merges with a primary 100
times more massive. Each panel represents a different snapshot of the secondary along its
orbital decay trajectory. The trajectory of the secondary is shown on the bottom of the plot,
as in Figure 3. Most of the envelope particles are unbound after the first pass, and by the
second pass, 90% of the matter is stripped away. However, the inner particles remain bound
to the secondary's potential.
Fig. 6. -- Density profile for the mass ratio 100:1 basic experiment. In the top panel, we
illustrate the change in the inner density profile of the secondary. We plot the density at
a radius r against radius in parsecs. The solid line is the final secondary profile, the thick
dashed line is the initial state of the secondary, and the dotted line is the density profile of
the primary for comparison. In the bottom panel, the thick solid line represents the resulting
remnant, the dotted line corresponds to the final state of the primary, and the thin solid
line represents the final state of the secondary.
Inside the tidal radius, the secondary is
unchanged (refer to § 4.1).
Fig. 7. -- Density profile for the 10:1 basic experiment. See caption for Figure 6.
Fig. 8. -- Density profile for the 2.5:1 basic experiment. See caption for Figure 6.
Fig. 9. -- We plot the secondary's density at radius r versus radius for two drag coefficients.
We adjusted the magnitude of the dynamical friction force by a drag coefficient that we
obtained by fitting the orbital decay time generated by Chandrasekhar to that generated by
a full treecode simulation. This plot indicates that varying the drag coefficient, fdrag, by a
-- 26 --
small amount does not affect the overall result of the merger. The solid line is the initial
secondary, the dashed line is the fdrag = 0.2 case and the dotted line is for fdrag = 0.1. In
both simulations, the secondary remains intact at the center, although the more pericenter
passes, the more the outer envelope is stripped.
Fig. 10. -- Better inner resolution for 100:1. See the caption for Figure 6. Inside the tidal
radius, the secondary is essentially intact, although a comparison with Figure 6 demonstrates
that more mass is lost inside the tidal radius (refer to § 4.3).
Fig. 11. -- Better inner resolution for 10:1. See caption for Figure 6.
Fig. 12. -- Better inner resolution for 2.5:1. See caption for Figure 6. In this experiment, the
density of the secondary at the center has dropped to or below the density of the primary.
Fig. 13. -- κ = 0.2 orbit for 100:1. See caption for Figure 6. The secondary is intact inside
the tidal radius.
Fig. 14. -- κ = 0.6 Orbit for 10:1. See caption for Figure 6. In this experiment, there is a
minor amount of mass loss inside the tidal radius, even for very small radii, but qualitatively,
the secondary survives and overwhelms the inner density profile of the remnant (bottom
panel).
Fig. 15. -- κ = 0.2 orbit for 10:1. See caption for Figure 6.
Fig. 16. -- Overplot of final secondary density for different orbits. We plot the density at a
radius r against radius in parsecs. The top panel is for the 100:1 mass ratio and the bottom
panel illustrates the 10:1 mass ratio. The solid line in each case is the κ = 0.0 case, the short
dashed line is the κ = 0.2 case, and the long dashed line is the κ = 0.6 case.
Fig. 17. -- Treecode comparison for 2.5:1. See caption for Figure 6, noting that in this
experiment, the density profile of the primary changes. The secondary evolves until it is
-- 27 --
everywhere below the initial state of the primary. Refer to § 4.5.
Fig. 18. -- Comparison to Weinberg's results for 10:1. See caption for Figure 6.
In this
experiment, the secondary is destroyed, as it is in Weinberg's paper. The secondary central
density is only a factor of 3 greater than the primary initially, and after the merger, the
density of the secondary is everywhere less than 10 percent greater than the primary. We
regard this as consistent with Weinberg's experiment.
-- 28 --
Table 1. Eta Model Galaxy Parameters
M1
M2
Galaxy
Mv
rcore
a
Mcore
b
ηcore
renv
Menv
ηenv
rhalf
tcore
100:1
Primary
-22.0
263
4.7 x 1010
2.15
4000
3.4 x 1012
2.2
10617
5932
Secondary
-18.0
3.8
1.6 x 108
1.0
300
3.5 x 1010
1.5
508
123
1.34
98.66
4.4 x 10−3
0.996
10:1
Primary
-21.5
155
2.3 x 1010
1.96
4000
1.9 x 1012
2.0
9506
9220
Secondary
-19.5
18.63
1.3 x 109
1.23
1029
1.95 x 1011
1.5
1734
1113
0.117
9.88
6.7 x 10−3
0.993
2.5:1
Primary
-21.5
155
2.3 x 1010
1.96
4000
1.9 x 1012
2.0
9506
9220
Secondary
-20.55
56.6
5.9 x 109
1.61
2000
6.5 x 1011
2.0
4800
13101
2.9 x 10−2
2.48
9.3 x 10−4
0.9908
aradii are in units of pc.
bthe top masses are in units of M⊙, and the bottom masses are normalized such that the total secondary
mass is 1.0
-- 29 --
Table 2. Synopsis of Experiments
Type
Mass Ratio
κ
fdrag
Effect on Secondary
Fiducial
Vary Drag
Inner Eta
Non-Plunging Orbits
Treecode
Weinberg
100:1
10:1
2.5:1
100:1
100:1
10:1
2.5:1
100:1
10:1
2.5:1
10:1
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.2
0.2
0.6
0.0
0.1
0.4
0.1
0.1
0.2
0.1
0.4
0.1
0.1
0.4
0.1
0.1
N/A
0.084
Intact
Intact
Disrupted
Intact
Intact
Intact
Intact
Disrupted
Intact
Intact
Intact
Disrupted
Disrupted
Table 3. King Model Galaxy Parameters
Mass Ratio
Galaxy
Mv
a W0
rcore
b Massb
b
ρ0
10:1
Primary
-21.5
Secondary
-19.5
9.5
9.5
3.15
1.0
10.0
.0032
1.0
.01
aMagnitude is a free parameter chosen to compare the η model results
bnormalized such that the mass and rcore of the secondary is 1.0
3
2.5
2
1.5
1
5.5
5
4.5
4
3.5
3
A
-24
-22
-20
-18
-16
C
-24
-22
-20
-18
-16
3
2
1
0
3
2.5
2
1.5
B
-24
-22
-20
-18
-16
D
-24
-22
-20
-18
-16
-5
-10
-1
0
1
2
Log(r) pc
3
4
400
200
0
0
500
1000
1500
2000
Secondary Crossing Times
1.04
1.02
1
0.98
0.96
1.04
1.02
1
0.98
0.96
0
50
100
150
200
0
50
100
150
200
core crossing times
-5
-10
-5
-10
-1
0
1
2
Log(r) pc
3
4
-5
-10
-5
-10
-1
0
1
2
Log(r) pc
3
4
-5
-10
-5
-10
-1
0
1
2
Log(r) pc
3
4
-5
-10
-15
-1
0
1
2
Log(r) (pc)
3
4
-5
-10
-5
-10
-1
0
1
2
Log(r) pc
3
4
-5
-10
-5
-10
-1
0
1
2
Log(r) pc
3
4
-5
-10
-5
-10
-1
0
1
2
Log(r) pc
3
4
-5
-10
-5
-10
-1
0
1
2
Log(r) pc
3
4
-5
-10
-5
-10
-1
0
1
2
Log(r) pc
3
4
-5
-10
-5
-10
-1
0
1
2
Log(r) pc
3
4
-5
-10
-1
0
1
2
Log(r) pc
3
4
-5
-10
-5
-10
Treecode
Treecode
-1
0
1
2
Log(r) pc
3
4
-5
-10
-5
-10
King Model
King Model
-1
0
1
2
Log(r) pc
3
4
|
astro-ph/0609688 | 3 | 0609 | 2006-11-28T18:35:49 | WASP-1b and WASP-2b: Two new transiting exoplanets detected with SuperWASP and SOPHIE | [
"astro-ph"
] | We have detected low-amplitude radial-velocity variations in two stars, USNO-B1.0 1219-0005465 (GSC 02265-00107 = WASP-1) and USNO-B1.0 0964-0543604 (GSC 00522-01199 = WASP-2). Both stars were identified as being likely host stars of transiting exoplanets in the 2004 SuperWASP wide-field transit survey. Using the newly-commissioned radial-velocity spectrograph SOPHIE at the Observatoire de Haute-Provence, we found that both objects exhibit reflex orbital radial-velocity variations with amplitudes characteristic of planetary-mass companions and in-phase with the photometric orbits. Line-bisector studies rule out faint blended binaries as the cause of either the radial-velocity variations or the transits. We perform preliminary spectral analyses of the host stars, which together with their radial-velocity variations and fits to the transit light curves, yield estimates of the planetary masses and radii. WASP-1b and WASP-2b have orbital periods of 2.52 and 2.15 days respectively. Given mass estimates for their F7V and K1V primaries we derive planet masses 0.80 to 0.98 and 0.81 to 0.95 times that of Jupiter respectively. WASP-1b appears to have an inflated radius of at least 1.33 R_Jup, whereas WASP-2b has a radius in the range 0.65 to 1.26 R_Jup. | astro-ph | astro-ph | Mon. Not. R. Astron. Soc. 000, 1 -- 7 (2006)
Printed 21 October 2018
(MN LATEX style file v2.2)
WASP-1b and WASP-2b: Two new transiting exoplanets
detected with SuperWASP and SOPHIE
A. Collier Cameron1⋆, F. Bouchy12,13, G. H´ebrard12, P. Maxted5, D. Pollacco2,
F. Pont 10, I. Skillen8, B. Smalley5, R. A. Street2, R.G. West 3, D.M. Wilson 5,
S. Aigrain6, D.J. Christian2, W.I. Clarkson4,15, B. Enoch4, A. Evans5,
A. Fitzsimmons2, M. Fleenor16, M. Gillon14, C.A. Haswell4, L. Hebb1, C. Hellier5,
S.T. Hodgkin6, K. Horne1, J. Irwin6, S.R. Kane7, F.P. Keenan2, B. Loeillet11,
T.A. Lister1,5, M. Mayor10, C. Moutou11, A.J. Norton4, J. Osborne3, N. Parley4,
D. Queloz10, R. Ryans2, A.H.M.J. Triaud1, S. Udry10, and P.J. Wheatley 9
1School of Physics and Astronomy, University of St Andrews, North Haugh, St Andrews, Fife KY16 9SS, UK.
2ARC, Main Physics Building, School of Mathematics & Physics, Queen's University, University Road, Belfast, BT7 1NN, UK.
3Department of Physics and Astronomy, University of Leicester, Leicester, LE1 7RH, UK.
4Department of Physics and Astronomy, The Open University, Milton Keynes, MK7 6AA, UK.
5Astrophysics Group, Keele University, Staffordshire, ST5 5BG.
6Institute of Astronomy, University of Cambridge, Madingley Road, Cambridge, CB3 0HA, UK.
7Department of Astronomy, University of Florida, 211 Bryant Space Science Center, Gainesville, FL 32611-2055, USA.
8Isaac Newton Group of Telescopes, Apartado de Correos 321, E-38700 Santa Cruz de la Palma, Tenerife, Spain.
9Department of Physics, University of Warwick, Coventry CV4 7AL, UK.
10Observatoire de Gen`eve, Universit´e de Gen`eve, 51 Ch. des Maillettes, 1290 Sauverny, Switzerland.
11Laboratoire d'Astrophysique de Marseille, BP 8, 13376 Marseille Cedex 12, France
12Institut d'Astrophysique de Paris, CNRS (UMR 7095) -- Universit´e Pierre & Marie Curie, 98bis bvd. Arago, 75014 Paris, France
13Observatoire de Haute-Provence, 04870 St Michel l'Observatoire, France
14Institut d'Astrophysique et de G´eophysique, Universit´e de Li`ege, All´ee du 6 Aout 17, 4000 Li`ege, Belgium
15STScI, 3700 San Martin Drive, Baltimore, MD 21218, USA
16Volunteer Observatory, 10305 Mantooth Lane Knoxville, TN 37932, USA
Accepted 2006 November 28. Received 2006 November 7; in original form 2006 September 22
ABSTRACT
We have detected low-amplitude radial-velocity variations in two stars, USNO-B1.0
1219-0005465 (GSC 02265-00107 = WASP-1) and USNO-B1.0 0964-0543604 (GSC
00522-01199 = WASP-2). Both stars were identified as being likely host stars of tran-
siting exoplanets in the 2004 SuperWASP wide-field transit survey. Using the newly-
commissioned radial-velocity spectrograph SOPHIE at the Observatoire de Haute-
Provence, we found that both objects exhibit reflex orbital radial-velocity variations
with amplitudes characteristic of planetary-mass companions and in-phase with the
photometric orbits. Line-bisector studies rule out faint blended binaries as the cause
of either the radial-velocity variations or the transits. We perform preliminary spectral
analyses of the host stars, which together with their radial-velocity variations and fits
to the transit light curves, yield estimates of the planetary masses and radii. WASP-1b
and WASP-2b have orbital periods of 2.52 and 2.15 days respectively. Given mass esti-
mates for their F7V and K1V primaries we derive planet masses 0.80 to 0.98 and 0.81
to 0.95 times that of Jupiter respectively. WASP-1b appears to have an inflated radius
of at least 1.33 RJup, whereas WASP-2b has a radius in the range 0.65 to 1.26 RJup.
Key words: methods: data analysis -- stars: planetary systems -- techniques: radial
velocities -- techniques: photometric
6
0
0
2
v
o
N
8
2
3
v
8
8
6
9
0
6
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
⋆ E-mail:[email protected]
c(cid:13) 2006 RAS
2
A. Collier Cameron et al
1
INTRODUCTION
Extra-solar planets that transit their parent stars are of
key interest because their masses and radii can be deter-
mined directly, providing clues to their internal compo-
sitions (Guillot et al. 2006). They define the mass-radius-
separation relation for irradiated giant planets (Mazeh et al.
2005). They provide unique insights into their thermal prop-
erties (e.g. Charbonneau et al. 2005; Deming et al. 2005;
Deming et al. 2006) and the chemical compositions of their
atmospheres (Charbonneau et al. 2002; Vidal-Madjar et al.
2003; Vidal-Madjar et al. 2004).
The first exoplanet found to exhibit transits, HD
209458b (Charbonneau et al. 2000; Henry et al. 2000), was
initially discovered using the radial-velocity method that has
to date yielded the vast majority of the 210 known exo-
planets. Its inflated radius (Brown et al. 2001) presents a
challenge to theories of the structure and evolution of irra-
diated exoplanets. New radial-velocity discoveries are rou-
tinely subjected to careful photometric followup at around
the predicted times of transit. Indeed, recent radial-velocity
surveys targeting bright stars of high metallicity with the
specific goal of discovering new hot Jupiters have revealed
two new transiting planets in the last two years (Sato et al.
2005; Bouchy et al. 2005). Transiting planets thus comprise
roughly 10% of all the hot Jupiters with orbital periods un-
der 5 days or so, which is consistent with expectations for
randomly-oriented orbits.
The complementary approach is to look for transits
first, then to seek evidence of their planetary nature in
radial-velocity followup observations. The OGLE project pi-
oneered this approach, which has to date yielded five transit-
ing exoplanet candidates for which radial-velocity variations
have been detected with amplitudes indicative of planetary-
mass companions (Konacki et al. 2003; Bouchy et al. 2004,
Konacki et al. 2004; Pont et al. 2004; Konacki et al. 2005).
Several
teams have embarked on ultra-wide field
searches for transiting exoplanets. In this complementary
approach, small-aperture CCD imaging systems coupled to
commercial camera optics secure light curves of millions
of stars. The TrES survey has so far yielded two transit-
ing planets (Alonso et al. 2004; O'Donovan et al. 2006b).
The XO survey and HAT surveys have so far found one
each (McCullough et al. 2006; Bakos et al. 2006). Here we
present the first two transiting exoplanets from the Super-
WASP survey. The hardware, data analysis pipeline and
archive methodology for this project are described in detail
by Pollacco et al. (2006).
In a series of recent papers, Christian et al. (2006),
Street et al. (2006) and Lister et al. (2006) presented transit
candidates from the inaugural 2004 May -- September observ-
ing season of the five SuperWASP wide-field survey cameras
that were operating at that time. The light curves of some
1.1 × 106 stars in the magnitude range 8 < V < 13 were
searched for periodic shallow transits, and several dozen
plausible candidates were identified for detailed followup.
The candidate selection methodology is described in detail
in these papers and by Collier Cameron et al. (2006).
Table 1. Target list for SOPHIE radial-velocity followup pro-
gramme. The 1SWASP identifiers give the J2000 stellar coor-
dinates. The second column gives the number Ncl of SOPHIE
observations that were required to classify and eliminate each
non-planetary system, and the total number of observations for
planet-bearing systems. Half the astrophysical false positives were
eliminated after a single observation. Five more were eliminated
after a second observation, leaving 8 narrow, single-lined targets.
1SWASP ID
Ncl
SOPHIE followup classification
J161732.90+242119.0
J165949.13+265346.1
J174118.30+383656.3
J181252.03+461851.6
J183104.01+323942.7
J184303.62+462656.4
J203054.12+062546.4
J204125.28+163911.8
J205027.33+064022.9
J205308.03+192152.7
J210318.01+080117.8
J211608.42+163220.3
J214151.03+260158.5
J215802.14+253006.1
J222317.60+130125.8
J223320.44+370139.1
J223651.20+221000.8
J234318.41+295556.5
J002040.07+315923.7
J005225.90+203451.2
J010151.11+314254.7
J025500.31+281134.0
J031103.19+211141.4
J051221.34+300634.9
SB1
SB2
Late-type giant
Planet host: WASP-2
Late-type giant
3 No detectable RV variation
1 Rapid rotator
2
2 No detectable RV variation
1
1
9
1
3 No detectable RV variation
1
4
1
1
1
1
1
2
2
7
3 No detectable RV variation
1 Rapid rotator
4 No detectable RV variation
1 Rapid rotator
2 Multiple system
SB2
Line bisector variable
SB2
Late-type giant
SB2
Late-type giant
SB2
SB2
SB1
Planet host: WASP-1
2 SOPHIE AND OFXB OBSERVATIONS
We conducted a radial-velocity survey of a sample of high-
priority SuperWASP transit candidates, using the newly-
commissioned SOPHIE spectrograph (Bouchy et al. 2006)
on the 1.93-m telescope at the Observatoire de Haute-
Provence during the four nights from 2006 August 31 to
2006 September 3, and between September 12 and 19 dur-
ing the science verification phase of SOPHIE. SOPHIE is a
bench-mounted, fibre-fed spectrograph built on the same de-
sign principles as the HARPS instrument (Pepe et al. 2004)
on the ESO 3.6-m telescope at La Silla. The spectrograph's
thermal environment is carefully controlled, with the aim of
achieving radial-velocity measurements with stability better
than 2 m s−1. For the targets studied here, a radial-velocity
precision of 10 to 15 m s−1 is adequate to establish or re-
ject the planetary nature of a transit candidate. We there-
fore elected to use SOPHIE's High-Efficiency (HE) mode,
which has resolving power λ/∆λ = 35000. The CCD de-
tector records 39 spectral orders spanning the wavelength
range from 387 to 694 nm. Radial velocities are determined
by cross-correlation with a mask spectrum matched to the
spectral type of the target (Baranne et al. 1996; Pepe et al.
c(cid:13) 2006 RAS, MNRAS 000, 1 -- 7
Transiting exoplanets WASP-1b and WASP-2b
3
Table 2. Journal of radial-velocity measurements of WASP-1 and WASP-2. The 1SWASP identifiers give the J2000 stellar coordinates
of the photometric apertures; the USNO-B1.0 number denotes the star for which the radial-velocity measurements were secured. The
uncertainties given here include 10.0 m s−1 systematic error in added in quadrature to the formal photon-noise error. The fifth and
sixth columns give the FWHM of the CCF dip and the contrast of the dip as a fraction of the weighted mean continuum level. The
signal-to-noise ratio near 550 nm is given in column 6, and the spectral type of the cross-correlation mask in column 7.
HJD
texp
(s)
Vr
km s−1
FWHM Contrast
km s−1
%
S:N
Mask
Sp. type
Notes
1SWASP J002040.07+315923.7: USNO-B1.0 1219-0005465 = GSC 02265-00107 = WASP-1
2453979.6311
2453980.5558
2453981.5649
2453981.6752
2453982.4167
2453982.5843
2453982.6758
2400 −13.425 ± 0.011
2400 −13.484 ± 0.012
2400 −13.587 ± 0.016
1219 −13.520 ± 0.021
2100 −13.376 ± 0.014
2100 −13.412 ± 0.013
2407 −13.398 ± 0.012
During transit
Variable cloud
Variable cloud
37.7
36.5
25.0
21.4
31.5
33.3
35.0
25.3
21.7
17.1
13.8
17.8
19.3
22.0
9.7
9.7
9.8
9.8
9.8
10.0
10.0
G2
G2
G2
G2
G2
G2
G2
1SWASP J203054.12+062546.4: USNO-B1.0 0964-0543604 = GSC 00522-01199 = WASP-2
2453981.5065
2453982.3786
2453982.4962
2453991.3817
2453991.5102
2453996.3529
2453996.4301
2453997.3824
2453998.3415
1906 −28.125 ± 0.186
2500 −27.711 ± 0.012
2500 −27.736 ± 0.013
1200 −27.780 ± 0.011
1200 −27.812 ± 0.011
1200 −28.037 ± 0.011
1200 −28.020 ± 0.012
1500 −27.723 ± 0.012
1200 −27.987 ± 0.011
12.4
37.3
39.4
45.0
41.6
37.5
35.4
36.7
40.2
2.2
19.1
17.0
30.2
29.8
25.0
24.4
24.5
25.3
K5
K5
K5
K5
K5
K5
K5
K5
K5
8.9
93
9.0
8.8
8.8
9.0
9.0
9.0
9.0
During transit
Variable to heavy cloud
2002). Automatic data reduction at the telescope allows
highly efficient candidate selection and data assessment.
The target list was drawn from the candidate papers of
Christian et al. (2006), Street et al. (2006) and Lister et al.
(2006), and from further candidate lists covering different
regions of the sky for which candidate papers are currently
in preparation. Table 1 gives a brief summary of all targets
observed and the outcomes of the SOPHIE observations.
A full investigation of the natures of the other objects
observed during this radial-velocity study is currently in
progress. The results of this investigation and its implica-
tions for improving our pre-selection criteria will be pub-
lished in a companion paper. Here we present the two targets
in the survey sample that have single, narrow-lined cross-
correlation functions (CCF) with radial velocity variations
in agreement with an oscillation precisely phased with the
ephemeris predicted from SuperWASP data. The amplitudes
of these radial velocity oscillations are less than a few hun-
dred m s−1 and no significant line-bisector variations are
detected (see Section 5). Thus we can conclude that each
of these two stars harbours a transiting planet. The radial-
velocity measurements for the two stars are listed in Table
1.
A complete transit of WASP-2 was observed using a
CCD camera with R-band filters on the 60-cm telescope
of the Observatoire Fran¸cois-Xavier Bagnoud at St-Luc
(OFXB), on the night of 2006 September 12/13 UT. A full
transit of WASP-1 was observed on the morning of 2006
October 2 UT using an SBIG ST10XME CCD camera with
R-band filter on the 0.35-m Schmidt-Cassegrain Telescope
at the Volunteer Observatory at Knoxville, Tennessee.
c(cid:13) 2006 RAS, MNRAS 000, 1 -- 7
3 STELLAR PARAMETERS
The extracted SOPHIE spectra were used for a preliminary
analysis using the uclsyn spectral synthesis package and
ATLAS9 models without convective overshooting (Castelli
et al. 1997). The Hα, Nai D and Mgi b lines were used
as diagnostics of both Teff and log g. The abundances do
not appear to be substantially different from solar. We used
these values to infer the radii and masses of the stars, as
listed in Table 3. For WASP-1, comparison with the stellar
evolution models of Girardi et al. (2000) gives maximum-
likelihood values M∗ = 1.15M⊙ and R∗ = 1.24R⊙, but
many models with 1.06 < M∗/M⊙ < 1.39 and 1.04 <
R∗/R⊙ < 1.92 satisfy the spectroscopic constraints, because
the main sequence is very wide in that temperature range.
The radius and mass estimates for WASP-2 are better con-
strained. We have also used the available BV and 2MASS
photometry to estimate Teff using the Infrared Flux Method
(Blackwell & Shallis 1977), which gave results in agreement
with that obtained from the spectral analysis.
4 PLANETARY PARAMETERS
The formal precision of the radial-velocity observations de-
pends on the signal-to-noise ratio of the spectrum, the sharp-
ness and density of the stellar lines, and the scattered-light
background in the instrument. The formal errors on the ve-
locity measures are given by the semi-empirical estimator
σRV = 1.7√FWHM/(S:N*Contrast). The Contrast param-
eter quantifies the contrast of the CCF peak against the
additional background signal from light leakage into the as-
4
A. Collier Cameron et al
Table 3. Stellar parameters derived from preliminary spectral
analyses with uclsyn. Masses and radii are derived from the
Padua models of Girardi et al. (2000) assuming [Fe/H]= 0.1 ± 0.2
to reflect the uncertainty in the metallicity.
Parameter
WASP-1
WASP-2
GSC
WASP V (mag)
Spectral type
Teff (K)
log g
MV (mag)
M∗/M⊙
R∗/R⊙
02265-00107
00522-01199
11.79
F7V
6200 ± 200
4.3 ± 0.3
3.9 ± 0.4
1.15+0.24
−0.09
1.24+0.68
−0.20
11.98
K1V
5200 ± 200
4.3 ± 0.3
6.2 ± 0.5
0.79+0.15
−0.04
0.78 ± 0.06
yet incomplete spectrograph enclosure. Because we did not
use simultaneous thorium-argon wavelength calibration, the
accuracy of the radial-velocity measurements is limited by
the stability of the spectrograph over the 2 to 3 hours be-
tween successive thorium-argon calibration exposures, con-
sidering that its thermal control was not yet optimized. Tests
performed during the commissioning of the instrument indi-
cate that the velocity drift during a night is typically in the
range ±10 m s−1. Taking into consideration additional un-
certainties coming from the wavelength solution of the H.E.
mode and the guiding noise, we estimated that during our
run, the systematic RV errors were 10 m/s. Although the
wavelength calibration may vary only slowly during a given
night, the drifts may reasonably be expected to be uncor-
related from one night to the next. Guiding noise should
be uncorrelated even between successive observations. Since
most targets were observed only once per night, we treat
the additional systematic error as uncorrelated. We there-
fore used the quadrature sum of the formal and systematic
errors (Table 2) for all model fits.
The transits of WASP-1 and WASP-2 can be timed with
a precision of about 20 minutes from the 2004 SuperWASP
data set. The first and last transits in this data set are sep-
arated by about 120 days, so the accumulated uncertainties
in the transit timings at the epoch of the 2006 SOPHIE,
OFXB and Volunteer Observatory observations are no more
than a few hours. There is no ambiguity in number of cycles
between the 2004 and 2006 data sets. The OFXB and Volun-
teer Observatory transit observations establish the improved
photometric ephemerides in Table 4.
We estimated the stellar and planetary radii and the
planetary masses by minimising χ2 for the photometric and
radial-velocity measurements simultaneously with respect
to the analytic model of Mandel & Agol (2002) for small
planets (Rp/R∗ < 0.1), assuming that the planets have cir-
cular orbits. In modelling the SuperWASP photometry we
used linear limb darkening coefficients u = 0.51 and 0.63
(van Hamme 1993) for WASP-1 and WASP-2 respectively.
For the more precise data from OFXB and Volunteer Ob-
servatory we used the 4-coefficient model of Claret (2000).
We used a bootstrap analysis to estimate the errors on
the parameters of the model. This entails taking the residu-
Table 4. Results of simultaneous minimum-χ2 circular-orbit fits
to the photometric and radial-velocity data for WASP-1 and
WASP-2. The parameters of the lightcurve model are given in
terms of the radius of the star and planet (R∗ and Rp, respec-
tively), the separation of the stars (a) and the inclination (i). The
total number of degrees of freedom, including the NRV radial-
velocity measurements, is Ndf . The contribution of the radial-
velocity data to the value of χ2 is χ2
RV, using the errors given
in Table 2. Data in transit are given reduced weight. Standard
errors on the parameters are derived using a bootstrap analysis
as described in the text.
Parameter
WASP-1b
WASP-2b
Transit HJD
Period (days)
γ (km s−1)
K1 (m s−1)
a (AU)
b = a cos i/R∗
Rp/R∗
R∗/a
Ndf
χ2
NRV
χ2
RV
M∗(M⊙)
Mp/MJup
Rp/RJup
2453912.514 ± 0.001
2453991.5146 ± 0.0044
2.51995 ± 0.00001
−13.503 ± 0.009
115 ± 11
0.0369 -- 0.0395
0 -- 0.8
0.093 -- 0.104
0.168 -- 0.260
961
1420 -- 1449
7
11.6
2.152226 ± 0.000004
−27.863 ± 0.007
155 ± 7
0.0296 -- 0.0318
0 -- 0.8
0.119 -- 0.140
0.086 -- 0.132
1013
1627.2 -- 1647.1
9
13.4
(1.06 -- 1.39)
(0.73 -- 0.94)
(0.80 -- 0.98)±0.11
(0.81 -- 0.95)±0.04
1.33 -- 2.53
0.65 -- 1.26
als from the optimum light-curve fit, applying an arbitrary
phase shift and restoring the model transit at phase zero.
This preserves both outliers and the correlated noise char-
acteristics of the WASP data. The synthetic radial-velocity
data are generated by sampling the best fit radial-velocity
curve at the observed phases and adding gaussian random
deviates with the appropriate standard error. These syn-
thetic data are then fitted repeatedly, to recover the distri-
butions of the fitted parameter values.
There is a strong degeneracy between the impact pa-
rameter b = a cos i/R∗ and the parameter R∗/a when fitting
planet transit lightcurves of the quality presented here. We
therefore present the results of minimum-χ2 fits with the
value of b fixed at values b = 0 (corresponding to the mini-
mum radius of the star) and b = 0.8 for both stars. Models
with b > 0.8 give noticeably worse fits to the lightcurves.
The radial-velocity measurements and minimum-χ2 fits
are shown as a function of orbital phase in Fig. 1. The sinu-
soidal variation in radial velocity is clearly seen. The light
curves and minimum-χ2 fits to the phases around the tran-
sits are shown in Fig. 2. The parameters of the best fits
are given in Table 4. An additional 50 m s−1 was added in
quadrature to the uncertainties of radial-velocity observa-
tions during transits, to allow for the Rossiter-McLaughlin
effect (Rossiter 1924; McLaughlin 1924).
In spite of this, the contribution of the radial-velocity
data to the total value of χ2 in both stars is larger than
expected. WASP-1 yields a spectroscopic χ2
s = 11.6 for 5
degrees of freedom, while WASP-2 gives χ2
s = 13.4 for 7
c(cid:13) 2006 RAS, MNRAS 000, 1 -- 7
Transiting exoplanets WASP-1b and WASP-2b
5
Figure 1. Radial velocities of WASP-1and WASP-2 plotted against barycentric Julian date (upper) and folded on the photometric
ephemerides (lower). In all panels the solid line represents the best-fitting circular-orbit solution.
degrees of freedom. More extensive observations will be re-
quired to establish the cause of this small additional "jit-
ter" in the radial-velocity measurements. A common cause
of such radial-velocity "jitter" is rotationally-modulated dis-
tortion of the line profiles arising from magnetic activity in
the stellar chromosphere and photosphere. While the S:N in
the bluest orders of the SOPHIE spectra is insufficient for us
to measure the chromospheric emission cores in the Ca II HK
lines in any meaningful way, both stars have narrow CCFs
that suggest v sin i values less than ∼ 5 km s−1 and hence
low to moderate activity levels. We allow for this additional
variability in the bootstrap error analysis by adding 12 m
s−1 of additional radial-velocity jitter to the synthetic radial-
velocity data. We note that this amount of jitter is within
the ranges determined empirically for F and K stars with low
to moderate levels of chromospheric activity (Wright 2005),
and conclude that chromospheric activity is a likely cause.
Also given in Table 4 are the masses and radii of WASP-
1b and WASP-2b derived using the best-fit model parame-
ters. Despite the ambiguities in the light-curve solution it is
clear that the data presented show that WASP-1 and WASP-
2 have planetary-mass companions with gas-giant radii.
5 FAINT BLENDED-BINARY SCENARIOS
Although many types of stellar binary system can mimic a
planet-like transit signal, most are easily eliminated. Strong
rotational broadening implies tidal synchronisation by a
massive companion. Grazing double or single-lined stel-
lar binaries reveal their nature after one or two observa-
tions. Triple systems can, however, mimic transiting planets
in a way that is difficult to eliminate (Torres et al. 2004;
O'Donovan et al. 2006a). A bright single F, G or K star
with a much fainter, physically associated eclipsing-binary
companion will apparently produce shallow eclipses. If the
c(cid:13) 2006 RAS, MNRAS 000, 1 -- 7
Figure 2. Transit profiles of WASP-1 (upper) and WASP-2
(lower), fitted with the optimum limb-darkened models with pa-
rameters as in Table 4. Small symbols denote 2004 season Super-
WASP photometry; filled circles show the Volunteer Observatory
light curve of 2006 October 1 (WASP-1) and the OFXB light
curve of 2006 September 12 (WASP-2).
bright primary has narrow absorption lines, the faint bi-
nary's lines will produce apparent periodic velocity shifts in
the combined spectrum, in phase with the photometric or-
bit. The apparent shift occurs because the broadened lines
of the tidally-synchronised primary of the eclipsing binary
6
A. Collier Cameron et al
are Doppler shifted by orbital motions into the wings of the
composite line profile.
Several tests can reveal this type of system. Because the
spectral types of the bright star and the cooler eclipsing bi-
nary are very different, cross-correlation masks of different
spectral types tend to yield different orbital velocity ampli-
tudes (Santos et al. 2002). We computed the CCF with dif-
ferent masks without significant change in the radial-velocity
values.
Because the apparent Doppler shift arises through a
variable asymmetry in the line wings, line-bisector analysis
of the cross-correlation function is also an effective probe for
this type of system (Queloz et al. 2001). When the fainter,
moving spectrum is broadened by rotation, the induced ve-
locity shift is sufficiently weak that the apparent velocity
variation tends to be small unless the binary's lines are
strong enough to give a clearly-visible asymmetry in the
combined line profile. The change in line-bisector velocity
from the wings to the core of the line tends to be greater than
the induced velocity amplitude. We measured the asymme-
tries of the cross-correlation function peaks using the line-
bisector method of Queloz et al. (2001). For WASP-1 and
WASP-2 we found the scatter in bisector velocities from the
wings to the core of the CCF profile to be substantially
less than the measured orbital velocity amplitude, and un-
correlated with orbital phase. A third candidate, 1SWASP
J210318.01+080117.8, failed this test and was eliminated as
a planet candidate.
Therefore, we can eliminate the scenario "bright single
star plus faint, late-type short-period eclipsing binary" with
confidence for WASP-1 and WASP-2. While other more in-
tricate scenarios are possible, we were not able to contrive
any that could explain both the photometric and the veloc-
ity signals while remaining credible.
At the 380 and 140 pc distances of WASP-1 and WASP-
2, it should be possible to resolve such binary companions
at separations of a few tens of AU or more with the help
of adaptive optics. As an additional line of defence against
this type of astrophysical false positive, we secured high-
resolution H-band images of both targets with the NAOMI
adaptive-optics system on the 4.2-m William Herschel Tele-
scope on the nights of 2006 September 6 and 7. Images
with corrected FWHM = 0.25 arcsec reveal that WASP-1
has a stellar companion 4.7 arcsec to the north and 3.7
magnitudes fainter at H. The SOPHIE fibre aperture has a
diameter of 3 arcsec. Seeing and guiding errors may allow
some of the companion's light into the fibre, but we can be
fairly confident that the contamination is not significant at
visible wavelengths. Even if the companion were an eclips-
ing binary, it would be unlikely to mimic the radial-velocity
signature of a planet, and indeed the line-bisector analysis
eliminates this possibility. NAOMI H-band images taken on
2006 September 7 with corrected FWHM=0.2 arcsec, and
using the OSCA coronagraph system, show that WASP-2
has a stellar companion 2.7 magnitudes fainter at H located
0.7" to the east. This falls within the SOPHIE fibre aper-
ture. Additional NAOMI images secured during transit on
2006 September 10 20:00 to 20:20 UT with 0.2 arcsec cor-
rected FWHM showed no sign of the ∼1.5-mag deep eclipse
in the companion that would be needed to mimic a transit.
Future AO observations should reveal whether these faint
stellar companions (which are the only objects visible be-
Figure 3. Mass-orbital separation diagram for the 14 known
transiting exoplanets. WASP-1b and WASP-2b follow the general
trend with only high-mass planets surviving at small separations.
(From http://obswww.unige.ch/∼pont/TRANSITS.htm)
sides WASP-1 and WASP-2 in their respective NAOMI fields
of view) are chance alignments or common-proper-motion
companions.
6 DISCUSSION AND CONCLUSIONS
We have detected the presence of radial-velocity varia-
tions in two exoplanetary transit candidates. Our prelim-
inary analysis yields masses between 0.8 and 1.0 MJup
for both planets. WASP-1b and WASP-2b lie between
the "hot Jupiters" and the "very hot Jupiters" in the
mass-orbital separation diagram for transiting hot Jupiters,
(Fig. 3).Their intermediate masses appear consistent with
the general trend toward high masses at the smallest orbital
separations noted by Mazeh et al. (2005). WASP-2b in par-
ticular lies close to the minimum separation at which planets
in this mass range survive, making it a good candidate for
future mass-loss studies. The radius of WASP-1b, which or-
bits an F7V star, is poorly constrained by the SuperWASP
data alone, but appears to be at least 1.33 RJup. WASP-1b
seems thus to be an expanded, low-density planet, similar
to HD 209458b and HAT-P-1b (Bakos et al. 2006). Addi-
tional photometry of WASP-2b, which orbits a K1 dwarf
at a slightly smaller orbital separation yields a radius close
to that of Jupiter, suggesting a substantially higher density.
Additional high-precision photometry is needed to refine the
radii and densities of both planets; indeed, precise radius es-
timates have recently been derived from high-precision pho-
tometry by Shporer et al. (2006) and Charbonneau et al.
(2006), confirming the oversized nature of WASP-1b.
ACKNOWLEDGMENTS
We are grateful to all the staff of Observatoire de Haute
Provence for their efforts, their efficiency and their support
on the new instrument SOPHIE and for the photometric
observation with the 1.20-m telescope, and in particular H.
Le Coroller and R. Giraud. We acknowledge the award of
Director's Discretionary Time at Haute-Provence for addi-
tional observations of WASP-2 in SOPHIE science verifica-
tion time, without which the detection could not have been
c(cid:13) 2006 RAS, MNRAS 000, 1 -- 7
Transiting exoplanets WASP-1b and WASP-2b
7
Henry G. W., Marcy G. W., Butler R. P., Vogt S. S., 2000,
ApJ, 529, L41
Konacki M., Torres G., Jha S., Sasselov D. D., 2003, Na-
ture, 421, 507
Konacki M., et al., 2004, ApJ, 609, L37
Konacki M., Torres G., Sasselov D. D., Jha S., 2005, ApJ,
624, 372
Lister T. A., et al., 2006, MNRAS, Submitted
Mandel K., Agol E., 2002, ApJ, 580, L171
Mazeh T., Zucker S., Pont F., 2005, MNRAS, 356, 955
McCullough P. R., et al., 2006, ApJ, 648, 1228
McLaughlin D. B., 1924, ApJ, 60, 22
O'Donovan F. T., et al., 2006b, ApJ, 651, L61
O'Donovan F. T., et al., 2006a, ApJ, 644, 1237
Pepe F., et al., 2002, A&A, 388, 632
Pepe F., Mayor M., Queloz D., et al., 2004, A&A, 423, 385
Pollacco D. L., et al., 2006, PASP, 118, 1407
Pont F., Bouchy F., Queloz D., Santos N. C., Melo C.,
Mayor M., Udry S., 2004, A&A, 426, L15
Queloz D., et al., 2001, A&A, 379, 279
Rossiter R. A., 1924, ApJ, 60, 15
Santos N. C., et al., 2002, A&A, 392, 215
Sato B., et al., 2005, ApJ, 633, 465
Shporer A., Tamuz O., Zucker S., Mazeh T., 2006, MNRAS,
Submitted (arXiv:astro-ph/0610556 )
Street R. A., et al., 2006, MNRAS, Submitted
Torres G., Konacki M., Sasselov D. D., Jha S., 2004, ApJ,
614, 979
van Hamme W., 1993, AJ, 106, 2096
Vidal-Madjar A., et al., 2003, Nature, 422, 143
Vidal-Madjar A., et al., 2004, ApJ, 604, L69
Wilson D. M., et al., 2006, PASP, 118, 1245
Wright J. T., 2005, PASP, 117, 657
This paper has been typeset from a TEX/ LATEX file prepared
by the author.
secured. We extend our special thanks to the team of the Ob-
servatoire Fran¸cois-Xavier Bagnoud at St-Luc, and in par-
ticular to Brice-Olivier Demory and Fr´ed´eric Malmann. Ad-
ditional partial observations of the 2006 September 12 tran-
sit of WASP-2 were secured at short notice with the OHP
1.2-m telescope. The WASP project is funded and operated
by Queen's University Belfast, the Universities of Keele, St
Andrews and Leicester, the Open University, the Isaac New-
ton Group, the Instituto de Astrofisica de Canarias, the
South African Astronomical Observatory and by PPARC. In
recognition of the considerable regional support given to the
WASP project on La Palma, we would like to associate the
pseudonym Garafia-1 with the planet of WASP-1. AMHJT
was supported by an Undergraduate Research Bursary from
the the Nuffield Foundation. 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 In-
stitute of Technology, funded by the National Aeronautics
and Space Administration and the National Science Foun-
dation. This research has made use of the VizieR catalogue
access tool, CDS, Strasbourg, France. We are grateful to the
referee, Dr. Chris Tinney, for numerous helpful suggestions
to improve and clarify the manuscript.
REFERENCES
Alonso R., et al., 2004, ApJ, 613, L153
Bakos G. A., et al., 2006, ApJ, In press (arXiv: astro-
ph/0609369)
Baranne A., et al., 1996, A&AS, 119, 373
Blackwell D. E., Shallis M. J., 1977, MNRAS, 180, 177
Bouchy F., Pont F., Santos N. C., Melo C., Mayor M.,
Queloz D., Udry S., 2004, A&A, 421, L13
Bouchy F., et al., 2005, A&A, 444, L15
Bouchy F., The Sophie Team 2006, in Arnold L., Bouchy
F., Moutou C., eds, Tenth Anniversary of 51 Peg-b: Status
of and prospects for hot Jupiter studies, pp 319 -- 325
Brown T. M., Charbonneau D., Gilliland R. L., Noyes
R. W., Burrows A., 2001, ApJ, 552, 699
Castelli F., Gratton R. G., Kurucz R. L., 1997, A&A, 318,
841
Charbonneau D., Brown T. M., Latham D. W., Mayor M.,
2000, ApJ, 529, L45
Charbonneau D., Brown T. M., Noyes R. W., Gilliland
R. L., 2002, ApJ, 568, 377
Charbonneau D., et al., 2005, ApJ, 626, 523
Charbonneau D., Winn J. N., Everett M. E., Latham
D. W., Holman M. J., Esquerdo G. A., O'Donovan F. T.,
2006, ApJ, Submitted (arXiv:astro-ph/0610589)
Christian D. J., et al., 2006, MNRAS, 372, 1117
Claret A., 2000, A&A, 363, 1081
Collier Cameron A., et al., 2006, MNRAS, In press. (arXiv:
astro-ph/0609418)
Deming D., Harrington J., Seager S., Richardson L. J.,
2006, ApJ, 644, 560
Deming D., Seager S., Richardson L. J., Harrington J.,
2005, Nature, 434, 740
Girardi L., Bressan A., Bertelli G., Chiosi C., 2000, A&AS,
141, 371
Guillot T., et al., 2006, A&A, 453, L21
c(cid:13) 2006 RAS, MNRAS 000, 1 -- 7
|
0803.1432 | 1 | 0803 | 2008-03-10T15:06:43 | GINS: a new tool for VLBI Geodesy and Astrometry | [
"astro-ph"
] | In the framework of the "Groupe de Recherches de G\'eod\'esie Spatiale" (GRGS), a rigorous combination of the data from five space geodetic techniques (VLBI, GPS, SLR, LLR and DORIS) is routinely applied to simultaneously determine a Terrestrial Reference Frame (TRF) and Earth Orientation Parameters (EOP). This analysis is conducted with the software package GINS which has the capability to process data from all five techniques together. Such a combination at the observation level should ultimately facilitate fine geophysical studies of the global Earth system. In this project, Bordeaux Observatory is in charge of the VLBI data analysis, while satellite geodetic data are processed by other groups. In this paper, we present (i) details about the VLBI analysis undertaken with GINS, and (ii) the results obtained for the EOP during the period 2005-2006. We also compare this EOP solution with the IVS (International VLBI Service for geodesy and astrometry) analysis coordinator combined results. The agreement is at the 0.2 mas level, comparable to that of the other IVS analysis centers, which demonstrates the capability of the GINS software for VLBI analysis. | astro-ph | astro-ph |
GINS: a new tool for VLBI Geodesy and
Astrometry
G. Bourda1, P. Charlot1, R. Biancale2
(1) Observatoire Aquitain des Sciences de l'Univers - Universit´e Bordeaux I
Laboratoire d'Astrophysique de Bordeaux - UMR5804/CNRS - Floirac, France
(2) Centre National d'Etudes Spatiales - Groupe de Recherche de G´eod´esie Spatiale
Toulouse, France
In the framework of the Groupe de
Abstract.
Recherches de G´eod´esie Spatiale (GRGS), a rig-
orous combination of the data from five space
geodetic techniques (VLBI, GPS, SLR, LLR and
DORIS) is routinely applied to simultaneously
determine a Terrestrial Reference Frame (TRF)
and Earth Orientation Parameters (EOP). This
analysis is conducted with the software package
GINS which has the capability to process data
from all five techniques together. Such a combi-
nation at the observation level should ultimately
facilitate fine geophysical studies of the global
Earth system.
In this project, Bordeaux Ob-
servatory is in charge of the VLBI data analy-
sis, while satellite geodetic data are processed by
other groups. In this paper, we present (i) details
about the VLBI analysis undertaken with GINS,
and (ii) the results obtained for the EOP dur-
ing the period 2005–2006. We also compare this
EOP solution with the IVS (International VLBI
Service for geodesy and astrometry) analysis co-
ordinator combined results. The agreement is at
the 0.2 mas level, comparable to that of the other
IVS analysis centers, which demonstrates the ca-
pability of the GINS software for VLBI analysis.
reference
Keywords.
systems, Earth rotation, VLBI
astrometry,
geodesy,
1 Introduction
The software package GINS (G´eod´esie par
Int´egrations Num´eriques Simultan´ees)
is a
multi-technique software initially developed by
the GRGS/CNES (Groupe de Recherches de
G´eod´esie Spatiale – Centre National d'Etudes
Spatiales, Toulouse, France) for analysing satel-
lite geodetic data, and extended at later stages
for analysing data from other space geodetic
techniques (Meyer et al., 2000). Currently, GPS
(Global Positionning System), DORIS (Doppler
Orbitography and Radio-positioning Integrated
by Satellite), SLR (Satellite Laser Ranging),
LLR (Lunar Laser Ranging) and VLBI (Very
Long Baseline Interferometry) observations can
be processed with GINS. The parameters that
can be estimated comprise satellite orbits around
the Earth or another body of the solar system,
gravity field coefficients, Earth Orientation Pa-
rameters (EOP), station coordinates, or other
geophysical parameters. The well-known GRIM5
and EIGEN gravity field models were produced
with GINS in particular (Biancale et al., 2000;
Reigber et al., 2002).
A rigorous combination of all the above space
geodetic data has been developed to estimate sta-
tion coordinates and EOP simultaneously from
all techniques in the framework of the IERS (In-
ternational Earth Rotation and Reference Sys-
tems Service) multi-technique combination pilot
project.
In this analysis, observations of the
different astro-geodetic techniques (VLBI, GPS,
SLR, LLR and DORIS) are first processed sepa-
rately using GINS. The weekly datum-free nor-
mal equation matrices derived from the anal-
yses of the different techniques are then com-
bined to estimate station coordinates and EOP
(Coulot et al., 2007; Gambis et al., 2007). Re-
sults are made available to the IERS in the form
of SINEX files. In this project, the VLBI data
are analysed in Bordeaux, while the satellite
geodetic data are processed either in Toulouse
(for GPS, DORIS and LLR) or Grasse (for SLR),
with the final combination produced at Paris Ob-
servatory (see Figure 1 for more details about the
project organization).
The strength of the method is in the use of
a unique software for all techniques with iden-
tical and up-to-date models and standards, en-
suring homogeneous and reliable combined prod-
ucts. In addition, the solution benefits from com-
plementary constraints brought by the various
techniques.
Figure 1. Organization of the coordinated project of the GRGS for multi-technique combination at the
observation level. CLS (Collecte Localisation Satellite) is a private company funded in particular by the
CNES.
In this paper, we present an overview of the
analyses undertaken with this new VLBI soft-
ware, the results obtained for the EOP from 2005
to 2006, and the comparisons made with the IVS
(International VLBI Service for geodesy and as-
trometry) analysis coordinator combined results.
These comparisons indicate that GINS is at the
level of the other VLBI analysis software pack-
ages.
2 VLBI analysis with GINS: data
and modeling
Since 2005 the regular weekly VLBI data ac-
quired by the IVS have been routinely processed
with the GINS software in order to estimate the
EOP and the VLBI station positions. These data
include both the IVS intensive sessions (i.e. one-
hour long daily experiments) and the so-called
IVS-R1 and IVS-R4 sessions (i.e.
two 24-hour
experiments per week). Overall, a total of 20 sta-
tions have been used in such sessions.
Based on these data, weekly normal matrices
are produced for combination with the data ac-
quired by the other space geodetic techniques
(GPS, SLR, LLR and DORIS). The free VLBI
parameters include station positions and the five
EOP (Xp, Yp, U T 1 − U T C, dψ, dǫ) along with
clock and troposphere parameters. The clocks
are modeled using piecewise continuous linear
functions with breaks every two hours. The tro-
pospheric zenith delays are modeled in a similar
way except that breaks are applied every hour.
Continuity constraints of 10 µs and 10 cm are
applied to the clock and troposphere breaks, re-
spectively. The a priori EOP series used is the
IERS C04 series, the a priori Terrestrial Ref-
erence Frame (TRF) is VTRF2005 (Nothnagel,
2005), while the celestial frame is fixed to the
ICRF (International Celestial Reference Frame;
Ma et al., 1998; Fey et al., 2004).
The following tidal and atmospheric models
are also used in the analysis:
◦ IERS Conventions 2003 for solid Earth tides
and pole tide models (McCarthy and Petit,
2003),
◦ FES2004 for oceanic tides and oceanic load-
ing models (Lyard et al., 2006),
◦ 6h-ECMWF (European Center for Mete-
2
IVS-R1
IVS-R4
3
2.5
2
1.5
1
0.5
)
m
c
(
s
e
u
l
a
V
0
2005
2006
Years
2007
Figure 2. Post-fit weighted RMS delay residuals
with GINS for the IVS-R1 and IVS-R4 sessions con-
ducted in 2005-2006.
orological Weather Forecast) atmospheric
pressure fields only over continents (inverse
barometer hypothesis) for atmospheric load-
ing model,
◦ Niell tropospheric mapping functions (Niell,
1996).
3 VLBI analysis with GINS: results
and comparison
section, we present
In this
the VLBI-only
EOP results obtained based on a fixed TRF
(VTRF2005) and compare these to the IVS com-
bined EOP series. One set of EOP (Xp, Yp,
U T 1 − U T C, dψ, dǫ) was estimated for every
24-hour session.
(see RMS values in Table 1):
◦ 0.20 mas for the polar motion coordinates,
◦ 0.15 mas for the celestial pole offsets, and
◦ 10 µs for the Earth's angle of rotation.
These differences may partly arise from us-
ing different TRF in the two analyses: the IVS
analysis coordinator used ITRF2000, whereas we
fixed the TRF to VTRF2005. Another point to
highlight is that these comparisons used non-
weighted RMS to evaluate the differences be-
tween the EOP series obtained with GINS and
those published by the IVS analysis coordinator.
Such values are generally larger that the more
common "weighted RMS".
4 Conclusions and prospects
GINS is a new VLBI analysis software in the IVS
community. We showed that VLBI analyses un-
dertaken with GINS lead to EOP results that
agree at the level of 150–200 µas with respect
to the IVS combined series results. This is com-
parable to the other IVS analysis centers, which
demonstrates the capability of the GINS software
for VLBI-only analysis.
Another strength of GINS is the possibility
of analysing observations of five space geode-
tic techniques (VLBI, GPS, LLR, SLR and
DORIS) alltogether.
Such a combination at
the observation level
is one of the goals of
the IAG (International Association of Geodesy)
project GGOS (Global Geodetic Observing Sys-
tem; Rummel et al., 2005).
In the future, further developments and inves-
tigations are planned to refine such VLBI analy-
sis with GINS:
◦ To improve the relative weighting of the
◦ To adjust tropospheric gradients, together
with zenithal tropospheric delays.
VLBI observations.
Figure 2 shows the post-fit weighted RMS
(Root Mean Square) delay residuals obtained
with GINS for the IVS-R1 and IVS-R4 sessions
in 2005–2006. The RMS average over this period
is 1.08 cm (i.e. 36 ps) for the IVS-R1 sessions,
and 0.97 cm (i.e. 32 ps) for the IVS-R4 sessions.
the EOP series as de-
Figure 3 shows
of
rived with GINS,
the
on the basis
section.
analysis described in the previous
to
The
are plotted with respect
the IVS combined series (ivs06q3e.eops;
see
http://vlbi.geod.uni-bonn.de/IVS-AC/combi-eops/QUAT/HTML/start q.html).
Table 1 summarizes the statistics for these series,
plus those with respect to the IERS C04 series.
Our current VLBI-only EOP results agree with
the IVS combined series at the following levels
results
◦ To validate the underlying models in GINS
by carrying out detailed comparisons with
those implemented in the JPL (Jet Propul-
sion Laboratory) VLBI software MODEST
(Sovers and Jacobs, 1996).
◦ To submit to the IVS analysis coordinator
the EOP results obtained with GINS for fur-
ther evaluation.
3
Table 1. Mean and RMS differences for each of the five EOP series (Xp, Yp, U T 1 − U T C, dǫ, dψ sin ǫ) derived
with GINS when compared to (1) the IERS C04 series, and (2) the IVS combined series.
EOP
wrt
Xp
µas
Yp U T 1 − U T C
µas
µs
Mean C04
-201
RMS
IVS
C04
IVS
47
216
212
343
-95
215
211
3.4
2.3
9.9
10.2
dǫ
µas
49
36
150
145
dψ sin ǫ
µas
7
-62
139
139
◦ To adjust the radiosource coordinates to in-
vestigate the source position variability and
ultimately to produce a celestial reference
frame with GINS.
Acknowledgements
G. Bourda is grateful to the CNES (Centre Na-
tional d'Etudes Spatiales, France) for the post-
doctoral position granted at Bordeaux Observa-
tory. She wishes also to the Advisory Board of
the Descartes-Nutations prize for supporting the
journey to Vienna in order to present this work.
References
Biancale R., Balmino G., Lemoine J.-M., Marty J.-
C., Moynot B., Barlier F., Exertier P., Laurain O.,
Gegout P., Schwintzer P., Reigber C., Bode A.,
Konig R., Massmann F.-H., Raimondo J.-C.,
Schmidt R. & Zhu S. Y., 2000, "A new global
Earth's gravity field model from satellite orbit
perturbations: GRIM5-S1", Geophys. Res. Lett.,
27, 3611–3614
Coulot D., Berio P., Biancale R., Loyer S.,
Soudarin L. & Gontier A.-M., 2007, "Toward
a direct combination of
space-geodetic tech-
niques at the measurement level: Methodol-
ogy and main issues", J. Geophys. Res., 112,
10.1029/2006JB004336
Fey A.L., Ma C., Arias E.F., Charlot P., Feissel-
Vernier M., Gontier A.-M., Jacobs C.S., Li J. &
MacMillan D.S., 2004, "The Second Extension
of the International Celestial Reference Frame:
ICRF-EXT.1", AJ, 127, 3587–3608
Gambis D., Biancale R., Carlucci T., Lemoine J.-
M., Marty J.-C., Bourda G., Charlot P., Loyer S.,
Lalanne T. & Soudarin L., 2007, "Combination
of Earth Orientation Parameters and terrestrial
frame at the observational level", IAG Springer
Series (submitted)
Lyard F., Lef`evre F., Letellier T. & Francis O., 2006,
"Modelling the global ocean tides: a modern in-
4
sight from FES2004", Ocean Dynamics, 56, 394–
415
Ma C., Arias E.F., Eubanks T.M., Fey A.L., Gon-
tier A.-M., Jacobs C.S., Sovers O.J., Archi-
nal B.A. & Charlot P., 1998, "The International
Celestial Reference Frame as Realized by Very
Long Baseline Interferometry", AJ, 116, 516–546
McCarthy D.D. & Petit G., 2003, "IERS Conven-
tions (2003)", IERS Technical Note 32, Frankfurt
am Main: Verlag des Bundesamts fr Kartogra-
phie und Geodsie, paperback, ISBN 3-89888-884-
3 (print version)
Meyer U., Charlot P. & Biancale R., 2000, "GINS:
A new Multi-Technique Software for VLBI Analy-
sis", International VLBI Service for Geodesy and
Astrometry 2000 General Meeting Proceedings,
edited by Nancy R. Vandenberg and Karen D.
Baver, NASA/CP-2000-209893
Niell A.E., 1996, "Global mapping functions for the
atmosphere delay at radio wavelengths", J. Geo-
phys. Res., 101, 3227-3246, 10.1029/95JB03048
Nothnagel A., 2005, "VTRF2005 - A combined VLBI
Terrestrial Reference Frame", Proceedings of the
17th Working Meeting on European VLBI for
Geodesy and Astrometry, Noto, Italy
Reigber Ch., Balmino G., Schwintzer P., Biancale R.,
Bode A., Lemoine J.-M., Koenig R., Loyer S.,
Neumayer H., Marty J.-C., Barthelmes F., Per-
osanz F. & Zhu S.Y., 2002, "A high qual-
ity global gravity field model
from CHAMP
GPS tracking data and Accelerometry (EIGEN-
1S)", Geophysical Research Letters,
29(14),
10.1029/2002GL015064
Rummel R., Rothacher M. & Beutler G., 2005, "In-
tegrated Global Geodetic Observing System (IG-
GOS) – science rationale", Journal of Geodynam-
ics, 40(4–5), 357–362, 10.1016/j.jog.2005.06.003
Sovers O.J. & Jacobs C.S., 1996, "Observation Model
and Parameter Partials for the JPL VLBI Param-
eter Estimation Software MODEST ", JPL Publi-
cation 83-39, Rev. 6, August 1996
Polar Motion Coordinates wrt IVS combined series
Xp
Yp
2006
Years
2007
Celestial Pole Offsets wrt IVS combined series
dEps
dPsi*sin(Eps)
2006
Years
2007
UT1-UTC wrt IVS combined series
1
0.5
0
-0.5
)
s
a
m
(
s
e
u
l
a
V
-1
2005
1
0.5
0
-0.5
)
s
a
m
(
s
e
u
l
a
V
-1
2005
)
s
d
n
o
c
e
s
o
r
c
i
m
(
s
e
u
l
a
V
40
20
0
-20
-40
2005
2006
Years
2007
Figure 3. VLBI Earth Orientation Parameters (Xp,
Yp, dǫ, dψ sin ǫ, U T 1 − U T C) estimated with GINS,
compared to the IVS analysis coordinator combined
series (ivs06q3e.eops) between 2005 and 2006.
5
|
astro-ph/0503699 | 1 | 0503 | 2005-03-31T14:16:40 | The basic parameters of gamma-ray-loud blazars | [
"astro-ph"
] | We determined the basic parameters, such as the central black hole mass ($M$), the boosting factor (or Doppler factor) ($\delta$), the propagation angle ($\Phi$) and the distance along the axis to the site of $\gamma$-ray production ($d$) for 23 $\gamma$-ray-loud blazars using their available variability timescales. In this method, the absorption effect depends on the $\gamma$-ray energy, emission size and property of the accretion disk. Using the intrinsic $\gamma$-ray luminosity as a fraction $\lambda$ of the Eddington luminosity, $L^{in}_{\gamma}=\lambda L_{Ledd.}$ and the optical depth equal to unity, we can determine the upper limit of the central black hole masses. We found that the black hole masses range between $10^{7}M_{\odot}$ and $10^{9}M_{\odot}$ when $\lambda$ = 0.1 and 1.0 are adopted. Since this method is based on gamma-ray emissions and the short time-scale of the sources, it can also be used for central black hole mass determination of high redshift gamma-ray sources. In the case of the upper limit of black hole mass there is no clear difference between BLs and FSRQs, which suggests that the central black hole masses do not play an important role in the evolutionary sequence of blazars. | astro-ph | astro-ph | Astronomy & Astrophysics manuscript no. jhfan
(DOI: will be inserted by hand later)
November 2, 2018
5
0
0
2
r
a
M
1
3
1
v
9
9
6
3
0
5
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
The basic parameters of γ-ray-loud blazars
J.H. Fan
1. Center for Astrophysics, Guangzhou University, Guangzhou 510400, China, e-mail: [email protected]
2. Chinese Academy of Science-Peking University Joint Beijing Astrophysical Center(CAS-PKU.BAC), Beijing,
China
3. National Astronomical Observatory, Chinese Academy of Sciences, Beijing, China
Received Oct 25, 2003/ Accepted Jan. 11, 2005
Abstract. We determined the basic parameters, such as the central black hole mass (M ), the boosting factor (or
Doppler factor) (δ), the propagation angle (Φ) and the distance along the axis to the site of γ-ray production (d)
for 23 γ-ray-loud blazars using their available variability timescales. In this method, the absorption effect depends
on the γ-ray energy, emission size and property of the accretion disk. Using the intrinsic γ-ray luminosity as a
fraction λ of the Eddington luminosity, Lin
γ = λLLedd. and the optical depth equal to unity, we can determine
the upper limit of the central black hole masses. We found that the black hole masses range between 107M⊙ and
109M⊙ when λ = 0.1 and 1.0 are adopted. Since this method is based on gamma-ray emissions and the short
time-scale of the sources, it can also be used for central black hole mass determination of high redshift gamma-ray
sources. In the case of the upper limit of black hole mass there is no clear difference between BLs and FSRQs,
which suggests that the central black hole masses do not play an important role in the evolutionary sequence of
blazars.
Key words. Galaxies:quasars-galaxies:BL Lacertae objects-galaxies:jet- galaxies:γ-rayst
1. Introduction
The EGRET instrument at CGRO has detected many
blazars (i.e. flat spectrum radio quasars (FSRQs) and BL
Lacertae objects (BLs)). Blazars emit most of their bolo-
metric luminosity in γ-rays (E > 100 MeV) (Hartman et
al. 1999). Many γ-ray emitters are also superluminal radio
sources (von Montigny et al. 1995). These objects share
some common properties, such as luminous γ-ray emis-
sion and strong variability in the γ-ray and other bands
on timescales from hours to days (see below). These facts
suggest that γ-ray emission in blazars is likely arise from
a jet. To explain its observational properties, a beaming
(black hole + accretion disk + jet) model has been pro-
posed. In the beaming model, a supermassive black hole is
surrounded by an accretion disk. Many authors have tried
to estimate the masses using different methods, (1) the
reverberation mapping technique (e.g. Wandel, Peterson
& Malkan 1999; Kaspi et al. 2000), (2) the gas and stellar
dynamics technique (see Genzel et al. 1997; Magorrian et
al. 1998; Kormendy & Gebhardt 2001), (3) the variability
time-scale technique (Fan et al. 1999; Cheng et al. 1999),
(4) the broad-line width technique (Dibai 1984; Wandel &
Yahil 1985; Padovani & Rafanelli 1988; Laor 1998; Mclure
offprint
Send
[email protected]
requests
to: Dr.
J.H.
Fan,
e-mail:
& Dunlop 2001; Vestergaard 2000 based on the assump-
tion that the clouds in the broad-line region (BLR) are
gravitationally bound and orbiting with Keplerian veloci-
ties). The central black hole mass is also found to be cor-
related with bulge luminosities (Kormendy & Richstone
1995), the bulge mass (Magorrian et al. 1998), the bulge
velocity dispersion (Ferrarese & Merritt 2000; Gebhardt
et al. 2000; Ferrarese et al. 2001) and the radio power
(Franceschini et al. 1998). However, Ho (2002) found that
the radio continuum power either integrated for the whole
galaxy or isolated for the core is poorly correlated with
the central black hole mass. The tight MBH − σ correla-
tion can be used for black hole mass determination. This
relation is also used by Barth et al. (2002) and Wu et al.
(2002) for black hole mass determination for Mrk 501 and
other AGNs. Cao (2002) estimated the black hole mass
for a sample of BL Lacertae objects based on the assump-
tion that broad emission lines are emitted from clouds
ionized by the radiation of the accretion disk surrounding
the black hole.
Since there is a large number of soft photons around
the central black hole, it is generally believed that the
escape of high energy γ-rays from the AGN depends on
the γ − γ pair production process. Therefore, the opac-
ity of γ − γ pair production in γ-ray-loud blazars can be
used to constrain the basic parameters. Becker & Kafatos
2
J.H. Fan: The basic parameters of γ-ray-loud blazars
γ = ΩD2F obs
γ
γ
(> 100M eV ), where F obs
(1995) have calculated the γ-ray optical depth in the X-
ray field of an accretion disk and found that the γ-rays
should preferentially escape along the symmetry axis of
the disk, due to the strong angular dependence of the
pair production cross section. The phenomenon of γ − γ
"focusing" is related to the more general issue of γ − γ
transparency, which sets a minimum distance between
the central black hole and the site of γ-ray production
(Bednarek 1993, Dermer & Schlickeiser 1994, Becker &
Kafatos 1995, Zhang & Cheng 1997). So, the γ-rays are
focused in a solid angle, Ω = 2π(1−cosΦ), suggesting that
the apparent observed luminosity should be expressed as
Lobs
and D are ob-
served γ-ray energy flux and luminosity distance respec-
tively. The observed γ-rays from an AGN require that the
jet almost points towards us and that the optical depth
τ is not greater than unity. The γ-rays are from a solid
angle, Ω, instead of being isotropic. In this sense, the non-
isotropic radiation, absorption and beaming (boosting) ef-
fects should be considered when the properties of a γ-ray-
loud blazars are discussed. In addition, the variability time
scale may carry the information about the γ-ray emission
region. These considerations require a new method to esti-
mate the central black hole mass and other basic parame-
ters of a γ-ray-loud blazar, which is the focus of the present
paper. In section 2, we introduce the method used to es-
timate the black hole mass and three other parameters
(the Doppler factor, the propagation angle of the γ-rays
and their emission distance at the symmetric axis above
the accretion disk). In section 3, we present the discussion
and a brief summary of the paper.
H0 = 75 km s−1 Mpc−1, and q0 = 0.5 are adopted
throughout the paper.
2. Mass estimation method and result
2.1. Method
Here we describe our method of estimating the basic pa-
rameters, namely, the central black hole mass (M ), the
boosting factor (or Doppler factor) (δ), the propagation
angle (Φ) and the distance along the axis to the site of
the γ-ray production (d) for γ-ray-loud blazars with short
timescale variabilities (see Cheng et al. 1999 for detail).
To do so, we consider a two-temperature disk (See Fig.
1). The γ-ray observations suggest that the γ-rays are
strongly boosted. From the high energy γ-ray emission
we know that the optical depth of γ-γ pair production
should not be larger than unity. In addition, the observed
short-time scale gives some information about the size of
emitting region. This can be used to constrain the basic
parameters of a γ-ray-loud blazar as in the following.
Optical depth Based on the paper by Becker &
Kafatos (1995), we can obtain an approximate empirical
formula for the optical depth for a two-temperature disk
case at an arbitrary angle, Φ (see Cheng et al. 1999),
Z
Φ
α
A
d
O
X−ray
Θ
γ−ray
interaction point
ρ
η
β
R
n
X
X−ray emitting disk
R0
R ms
Fig. 1. Schematic diagram of γ-ray propagation above
a two-temperature disk surrounding a supermassive black
hole. γ-rays interact with the soft X-ray photons produced
at all points on the disk. The interaction angle between the
γ-ray and X-ray photons is Θ, the angle between the γ-ray
trajectory and the z-axis is Φ. η is the distance between
the photon-photon interaction point and the soft photon
emission point in the disk.
where k is
τγγ(M7, Φ, d) = 9×Φ2.5(
d
Rg
)−
2αX +3
2 +kM −1
7 (
d
Rg
)−2αX−3 , (1)
k = 4.61 × 109 Ψ(αX )(1+z)3+αX F ′
(2αX +1)(2αX +3)
0(1+z−√1+z)2
×
( R0
Rg
[
)2αX +1−( Rms
)−1−( R0
( Rms
Rg
Rg
Rg
J.H. Fan: The basic parameters of γ-ray-loud blazars
3
)2αX +1
)−1
]( Eγ
4mec2 )αX ,
(2)
Substituting Eqs. (6) and (5) into Eq. (1), we obtain
a relation for τγγ(Φ, M, Liso),
M7 is the black hole mass in units of 107M⊙, Ψ(αX ) a
function of the X-ray spectral index, αX , F ′0 the X-ray flux
parameter in units of cm−2 s−1, me the electron mass, c
the speed of light, Rg = GM
c2 Schwarzschild radius, Eγ the
average energy of the γ-rays. The inner and outer radii of
the hot region of a two-temperature accretion disk (Becker
& Kafatos 1995) are R0 and Rms respectively. Eq. (1)
shows that the optical depth depends on d, Φ and M .
Time scale and the site of γ−ray production The time
scale gives us information the emitting region (or the dis-
tance along the axis to the site of γ-ray production),
d = cδ∆TD
1+z . For convenience, we express d in the form
of (∆TD in units of days).
d
Rg
= 1.73 × 103 ∆TD
1 + z
δM −1
7
(3)
1
Γ(1−βcosΦ) is the boosting factor, (Γ = (1 −
where δ =
β2)−1/2 is the bulk Lorentz factor and β is the bulk ve-
locity in unit of the speed of light c.
γ−Ray luminosity
In a relativistic beaming model,
the observed luminosity is correlated with the intrinsic
one in the frame comoving with the relativistic jet by
Lobs
γ =
δαγ +4
(1 + z)αγ−1 Lin
γ
where αγ is γ-ray spectral index. As mentioned above, the
observed γ-ray flux, F obs
(> 100M eV ), which is in units
of ergs cm−2 s−1, can be expressed as a function of the
intrinsic luminosity Lin
γ , the Doppler factor δ, the lumi-
nosity distance D, and the solid angle Ω (or propagation
angle Φ)
γ
F obs
γ
(> 100MeV) = (1 + z)1−αγ δαγ +4Lin
γ /ΩD2
τγγ(Φ, M7, Liso) = [9 × Φ2.5(1 − cosΦ)−
2αX +3
2αγ +8
+ kM −1
7 A−
2αX +3
2
(1 − cosΦ)−
2αX +3
αγ +4 ]A−
2αX +3
2
.
(7)
From the high energy γ-ray emission, we know that the
optical depth of γ-γ pair production should not be larger
than unity. So, we can assume τγγ(Φ, M7, Liso) = 1.0
for our purposes. However, there are two variables in the
equation of τγγ(Φ, M7, Liso) = 1.0, so one should have
one more equation to determine the basic parameters.
Fortunately, the τγγ(Φ, M, Liso) shows a minimum for a
certain mass M7 and angle Φ. For a given mass, M7, the
dependence of τγγ(Φ, M, Liso) on Φ is illustrated in Figure
2, in which we show the case of 0208-512. For the source,
the relevant values are αX = 1.04, αγ = 0.69, k = 6.41,
Liso = 2.0 × 1048ergss−1, z = 1.003, λ = 0.1, and ∆T =
134.4 hours respectively. In this sense, if we assume that
the minimum value of τγγ(Φ, M, Liso) is equal to 1.0, then
we will have the relation ∂τγγ
∂Φ M = 0.
Equation (7) gives
∂τγγ
∂Φ
M = [22.5Φ1.5(1 − cosΦ) − 9 ×
2αX + 3
2αγ + 8
Φ2.5sinΦ
−
2αX + 3
αγ + 4
kM −1
7 A−
2αX +3
2
(1 − cosΦ)−
2αX +3
2αγ +8 sinΦ]×
×[(1 − cosΦ)−
2αX +2αγ +11
2αγ +8 A−
2αX +3
2
],(8)
then, ∂τγγ
∂Φ M = 0 suggests that
22.5Φ1.5(1 − cosΦ) − 9 ×
−
2αX + 3
αγ + 4
kM −1
7 A−
2αX +3
2
(1 − cosΦ)−
Φ2.5sinΦ
2αX + 3
2αγ + 8
2αX +3
2αγ +8 sinΦ = 0 (9)
Under this consideration, we can finally get four rela-
If we define an isotropic luminosity as Liso = 4π D2F obs
100M eV ), we have
γ
(>
tions,
λ2.52 δαγ +4
L45
iso =
(1 − cosΦ)(1 + z)αγ−1 M7 ,
where Lin
γ = λLEdd = λ1.26 × 1045M7 is adopted, λ is
a parameter depending on specific γ-ray emission models,
iso is the isotropic luminosity in units of 1045 ergs s−1.
L45
(4)
From equation (4), we can get the Doppler factor
d
Rg
= 1.73 × 103 ∆TD
1 + z
δM −1
7
λ2.52 δαγ +4
L45
iso =
(1 − cosΦ)(1 + z)αγ−1 M7
)−2αX−3 = 1
2 + kM −1
7 (
2αX +3
d
Rg
9 × Φ2.5(
d
Rg
)−
22.5Φ1.5(1 − cosΦ) − 9 ×
Φ2.5sinΦ
2αX + 3
2αγ + 8
2αX +3
2αγ +8 sinΦ = 0 (10)
δ = (
L45
iso(1 − cosΦ)(1 + z)αγ−1
λ2.52M7
1
αγ +4
)
.
(5)
−
2αX + 3
αγ + 4
kM −1
7 A−
2αX +3
2
(1 − cosΦ)−
Substituting Eqs. (5) into Eq. (3), we can get a relation
for d(Φ, M, Liso),
d(Φ, M, Liso) = ARg(1 − cosΦ)
1
αγ +4
.
(6)
where
A = 1.73 × 103∆TD(1 + z)− 5
αγ +4 M −
7
αγ +5
αγ +4
(
L45
iso
λ2.52
)
1
αγ +4
in which, there are four basic parameters. So, for a source
with available data in the X-ray and γ-ray bands, the
masses of the central black holes, M7, the Doppler fac-
tor, δ, the distance along the axis to the site of the γ-ray
production, d, and the propagation angle with respect to
the axis of the accretion disk, Φ, can be derived from Eq.
(10), where Rms = 6Rg, R0 = 30Rg, and Eγ= 1GeV are
adopted.
4
J.H. Fan: The basic parameters of γ-ray-loud blazars
ferent sources and/or different observational periods, we
use the doubling timescale, ∆TD = (Fminimum/∆F )∆T ,
as the variability timescale, where ∆F = Fmaximum −
Fminimum is the variation of the flux over the time ∆T .
There are few simultaneous X-ray and γ-ray band obser-
vations, so the data considered here are not simultane-
ous. The γ-ray data by Hartman et al. (1999) are used to
calculate the γ-ray luminosity. The X-ray data are taken
from recent publications(See col. 5 in table 1). Except for
the two sources 1226+023 (Courviosier et al. 1988) and
2230+114 (Pica et al. 1988), the doubling time is taken
from Dondi & Ghisellini (1995).
The intrinsic γ-ray luminosity is unknown, so we as-
sume it to be close to the Eddington luminosity, say
λLEdd.. In the present work, λ = 0.1 and 1.0 are adopted
for the calculations. From the available X-ray and γ-ray
data, we can estimate the central black hole mass, M7
(see Table 1) and three other parameters (Φ, δ, d, these
values are not listed in Table 1). Since short-term time
scales and γ-ray emissions are included in our considera-
tion, only objects with those values can be involved in the
present paper. However, at present, short term timescales
are available only for 23 γ-ray loud blazars as listed in
Table 1, in which Col. 1 gives the name, Col. 2 the red-
shift, Col. 3 the identification where Q stands for a flat
spectral radio quasar and B for a BL Lacertae object,
Col. 4 the 1 KeV X-ray flux density in units of µ Jy, Col.
5 the reference for Col. 4, Col. 6 the X-ray spectral index
αX (the averaged value of < αX >= 0.67 (Comastri et al.
1997) is adopted for FSRQs and αOX = 1.31 is used for
αX for BLs if their X-ray spectral indices are unknown,
as done by Ghisellini et al. (1998)), Col. 7 the reference
for Col. 6, Col. 8 the flux F(>100MeV) in units of 10−6
photon cm−2 s−1, Col. 9 the γ-ray spectral index αγ =
1.0 is adopted for 0537-441 (see Fan et al. 1998); The
data in col no. (7) and (8) are mainly from by Hartman
et al. (1999) except for Mkn 501, which showed a flux of
F (> 100MeV) = (0.32 ± 0.13) × 10−6 photons cm−2 s−1
with a photon index of 1.6±0.5 during a 1996 multiwave-
length campaign (see Kataoka et al. 1999), Col. 10 the
doubling time scale in units of hours, Col. 11 reference for
Col. 10, Col. 12 the observed isotropic luminosity in units
of 1048erg s−1, Col. 13, the central black hole mass in units
of 107M⊙ (λ = 1), Co1. 14, the central black hole mass
in units of 107M⊙ (λ = 0.1). The references in Table 1 (
B97: Bloom et al. 1997; C97: Comastri et al. 1997; Ch99:
Chiappetti et al. 1999; DG: Dondi & Ghisellini 1995; F98:
Fan et al. 1998; Fo98: Fossati et al. 1998; H96: Hartman
1996; H96b: Hartman et al. 1996; H: Hartman et al. 1999;
K93: Kniffen et al. 1993; L98: Lawson & McHardy 1998;
M93: Mattox et al. 1993; M96: Madejski et al. 1996; M97:
Mattox et al. 1997; P96: Perlman et al. 1996; P88: Pica et
al. 1988; Q96: Quinn et al. 1996; S96: Stacy et al. 1996;
U97: Urry et al. 1997; W98: Wehrle et al. 1998)
Fig. 2. Plot of the optical depth against the angle Φ. For
illustration, we used the data of 0208-512, the solid curve
stands for the M7 = 95 case, while the dash-dotted curve
for M7 = 60 and the dashed curve for M7 = 130. Here
the relevant values are αX = 1.04, αγ = 0.69, k = 6.41,
Liso = 2.0 × 1048ergss−1, z = 1.003, λ = 0.1, and ∆T =
134.4 hours respectively.
2.2. Results
Since we are interested in the variability timescale, we
present here only the γ-ray-loud blazars with detected
short time scale variability. Since the variability timescale
corresponds to the different amplitude of variation for dif-
J.H. Fan: The basic parameters of γ-ray-loud blazars
5
3. Discussion
The central black hole plays an important role in the ob-
servational properties of AGNs and has drawn much atten-
tion. It may also shed some light on the evolution (Wang
et al. 2001; Barth et al. 2002; Cao 2002). There are sev-
eral methods for black hole mass determinations although
consensus has not been reached. In the present work, we
proposed a new method to estimate the central black hole
mass. It is constrained by the optical depth of the γ − γ
pair production and can be used if X-ray and γ-ray emis-
sions and short time-scale are known. This method can
be used to determine the central black hole mass of high
redshift gamma-ray sources. It is an approximate empir-
ical method, which is obtained from the data/figures of
Becker & Kafatos (1995). The mass determined in the
present paper corresponds to an optical depth of unity
and therefore the results correspond to the upper limit of
the central black hole mass. The main difference between
our consideration and others is that we assumed that the
γ-rays originate from a cone while others think that the γ-
rays are isotropic. However, our results are consistent with
those obtained by others as discussed in the following.
In our consideration, the estimated mass upper limits
for a sample presented here are in the range of 107M⊙ to
109M⊙, (0.57 ∼ 60) × 107M⊙ for λ = 1.0 and (0.87 ∼
95.0) × 107M⊙ for λ = 0.1. The real value of λ will cause
an uncertainty in the mass, but the uncertainty caused
by λ will be negligible. When λ decreases by a factor of
10, the mass increases by a factor of ∼ 1.5. The results
obtained with the present method are independent of the
γ-ray emission mechanism although it will depend the X-
ray emission mechanism.
For illustration, we will compare the mass estimation
for two sources, 3C 279 and Mkn 501. Several groups of
authors have estimated their masses.
3C 279 This quasar displayed two outbursts, one in
1991 and another in 1996. The upper limit of the central
black hole masses obtained from these outbursts are sim-
ilar. For λ = 1.0 the masses were M = 6.57 × 107M⊙
and M = 5.25 × 107M⊙ whereas for λ = 0.1, M =
10.2 × 107M⊙ and M = 8.12 × 107M⊙ for the 1991 and
1996 outbursts respectively. To fit the 3C 279 multiwave-
length energy spectrum corresponding to the 1991 γ-ray
flare, Hartman et al. (1996) used an accreting black hole
of 108M⊙; our result of (6.57−10.2)×107M⊙ is consistent
with their.
Mkn 501 TeV and X-ray emission result shows that
there is a possible period od 23 days in the light curves
of Mkn 501 (Hayashida et al. 1998; Kranich et al. 1999),
which may suggest binary black holes at center (Rieger &
Mannheim 2000, 2003; Villata & Raiteri 1999). The black
hole mass has been determined by many authors (0.2 −
3.4) × 109M⊙ (Barth et al. 2002; Kormendy & Gehbardt
2001; Merritt & Ferrarese 2001; Rieger & Mannheim
2003). Our present result shows that the central black
hole mass upper limit is (4.01 ∼ 6.27) × 107M⊙. This re-
sult is lower than those claimed by the authors mentioned
above, but this difference is probably from the facts that
(1) the methods used to estimate the mass by different
authors are based on different assumptions, (2) there is a
binary black hole system at the center of the BL Lacertae
objects, our result corresponds to the less massive black
hole. A binary black hole system is successfully used in the
explanation of the periodic variability of the BL Lacertae
object OJ 287 by Sillanpaa et al. (1988), who claimed that
the ratio of the secondary black hole mass to the primary
black hole mass is 0.004. Recently, Komossa et al. (2003)
reported that there is a binary black hole system in NGC
6240. If there is a binary black hole system at the center
of Mkn 501, and the more massive black hole of ∼ 109M⊙
corresponds to the primary black hole and the less mas-
sive black hole of 107M⊙ corresponds to the secondary,
then the ratio is 0.01, which is consistent with the result
for OJ 287 by Sillanpaa et al. (1988). In addition, our
result is consistent with that of De Jager et al. (1999),
(1 − 6) × 107M⊙.
BLs and FSRQs are two subclasses of blazars. From
the observational point of view, except for the emis-
sion line properties (the emission line strength in FSRQs
is strong while that in BL is weak or invisible), other
observational properties are very similar between them
(Fan 2002). Their relationship has drawn much atten-
tion (e.g. Sambruna et al. 1996; Scarpa & Falomo, 1997;
Ghisellini et al. 1998; D'Elia & Cavaliere 2000; Bottcher &
Dermer 2002; Fan 2002; Ciaramella et al. 2004). Bottcher
& Dermer (2002) proposed that the accretion rate rather
than the central black hole masses play an important role
in the evolutionary sequence (from FSRQs evolving to
BLs) of blazars. The accretion rate in FSRQs is much
larger than in BLs, there is not much gas in BLs to fuel
the central black hole. In the present paper, if we consider
BLs and FSRQs separately, the distribution of the mass
upper limits is not very show much different (see Fig. 3),
their average masses are logM = 8.06 ± 0.54 for FSRQs,
and logM = 8.13 ± 0.46 for BLs. There is no difference in
black hole mass between BLs and FSRQs. Recently, Wu et
al. (2002) also found that there is no mass difference be-
tween different subclasses of AGNs. This result suggests
that the central black hole mass play a less important role
in the evolutionary sequence as pointed out by Bottcher
& Dermer (2002). To investigate the evolutionary process
further, one should take into account the black-hole spin
and the accretion processes since the former effect is im-
portant for non-thermal radio emission while the latter
will change the properties of the outflow.
The high luminosity, rapid variability and superlumi-
nal motion observed in some γ-ray-loud blazars suggest
that the γ-ray emission is strongly beamed, therefore, one
can assume that the beamed emissions arise from a cer-
tain solid angle. In the present paper, our calculation show
that the propagation angle Φ and Doppler factor δ are in
the range of 9.◦68 to 61.◦14 and 0.12 to 3.31 for λ = 1.0.
For λ = 0.1 the value of Φ ranges from 8.◦91 to 56.◦49 and
δ 0.16 to 4.6. In the isotropic emission case, the γ-rays can
be detected at any angle, but in the case of non-isotropic
6
J.H. Fan: The basic parameters of γ-ray-loud blazars
Table 1. Black hole mass for 23 γ-ray-loud blazars
N ame
(1)
0208-512
0219+428
0235+164
0420-014
0458-020
0521-365
0528+134
0537-441
0716+714
0735+178
0829+046
0836+710
1101+384
1226+023
1253-055
1253-055
1510-089
1622-297
1633+382
1652+399
2155-304
2200+420
2230+114
2251+158
z
(2)
1.003
0.444
0.94
0.915
2.286
0.055
2.07
0.894
0.3
0.424
0.18
2.17
0.031
0.158
0.537
0.538
0.361
0.815
1.814
0.033
0.117
0.07
1.04
0.859
ID f1KeV
(4)
(3)
Q
B
B
Q
Q
B
Q
B
B
B
B
Q
B
Q
Q
Q
Q
Q
Q
B
B
B
Q
Q
0.61
1.56
2.5
1.08
0.1
1.78
0.65
0.81
1.35
0.248
1.07
0.819
37.33
12.07
2.43
2.0
0.718
0.08
0.42
10.1
0.058
1.84
0.486
1.08
Ref
(5)
C97
Fo98
M96
Fo98
DG
DG
C97
C97
Fo98
Fo98
DG
Fo98
Fo98
Fo98
H96b
L98
Fo98
M97
C97
C97
U97
P96
Fo98
Fo98
αX
(6)
Ref
(7)
C97
1.04
1.6
Fo98
1.01 M96
0.67
C97
C97
0.67
DG
0.68
C97
0.54
1.16
C97
Fo98
1.77
Fo98
1.34
C97
0.67
Fo98
0.42
2.10
Fo98
0.81
Fo98
0.68 H96b
L98
0.78
0.90
Fo98
C97
0.67
C97
0.53
C97
1.60
U97
1.25
1.31
P96
C97
0.67
0.62
Fo98
F
(8)
9.1
0.25
0.65
0.64
0.68
0.32
3.08
2.0
0.46
0.30
0.34
0.33
0.27
0.09
2.8
11.
0.49
17.
0.96
0.32
0.34
1.71
0.51
1.16
αγ
(9)
0.69
1.01
1.85
1.44
1.45
1.63
1.21
1.0
1.19
1.6
1.47
1.62
0.57
1.58
1.02
0.97
1.47
0.87
0.86
0.68
0.56
0.68
1.45
1.21
∆TD
(10)
Ref
(11)
134.4
30.0
72
33.6
144.0
72
24.
16.
1.92
28.8
24.
24.
1.92
24
12.
6.
S96
DG
M96
DG
DG
DG
DG
H96
DG
DG
DG
DG
DG
C88
K93
W98
DG
57.6
4.85 M97
M93
16.
Q96
6.
Ch99
3.3
3.2
B97
P88
48
1.92
DG
L48
iso
(12)
2.0
0.018
2.0
0.175
1.424
0.0002
18.4
3.01
0.013
0.013
0.003
0.548
0.0001
0.003
1.34
5.75
0.018
26.9
9.72
0.0003
0.003
0.019
0.184
0.32
M1 M0.1
(14)
(13)
60.9
19.73
36.75
10.98
31.48
31.1
4.52
10.46
2.22
20.03
12.14
1.67
1.5
8.9
6.57
5.25
28.03
5.44
3.43
4.01
3.26
11.0
15.33
0.57
95.0
29.80
53.69
16.43
47.2
46.44
6.97
15.96
3.28
29.37
18.2
2.53
2.31
13.27
10.2
8.12
41.9
8.51
5.42
6.27
5.11
15.95
22.95
0.87
1
emission, the emission is produced in the cone of a solid
angle of Ω, then γ-rays will not be detected at any angle.
From the observational properties such as high luminosity,
rapid γ-ray variability and superluminal motion, boosting
effects should be present in the γ-ray sources. So, the op-
tical depth should depend on the angle Φ, the boosting
factor δ, the central black hole masses and the distance
along the axis to the site of γ-ray production, d, which
αγ +4 .
can be expressed as d(Φ, M, Liso) = ARg(1 − cosΦ)
If the angle Φ is too small, then d is small, therefore, the
site of the γ-ray production is near the center where the
X-ray photon density will be high. So, the optical depth is
large. In this sense, the smaller the angle Φ, the smaller the
d, and the higher the X-ray photon density, which results
in a higher optical depth. On the other hand, when the
angle Φ is too large, the boosting factor δ is very small,
so that the optical depth is also large. Therefore, there
are angles that correspond to a small optical depth. From
Fig. 2, one can see that the minimum value depends on the
central black hole mass, so one can choose a mass so that
the minimum value of the optical depth is 1.0. The mass
estimated in this way is an upper limit. From our calcu-
lation, we found that Dopper factor δ < 1 in some cases;
the lower than unity Doppler factors do not conflict with
the beaming argument since we proposed that the emis-
sion is not isotropic in the present work. Other authors
assumed that the emission is isotropic. The different as-
sumptions will result in a ( 1−cosΦ
4+α times difference in
the Doppler factor. For the γ-ray emission regions, the ob-
)
1
2
tained results indicate that they are in the range of 17.2Rg
to 713Rg (λ = 1.0) or 18.8Rg to 640Rg (λ = 0.1).
In this paper, the optical depth of a γ-ray travelling
in the field of a two-temperature disk and beaming effects
have been used to determine the central mass, M , for 23
γ-ray-loud blazars with available short time-scales. The
masses obtained in the present paper are in the range of
107M⊙ to 109M⊙ for the whole sample. In the case of
black hole mass, there is no clear difference between BLs
and FSRQs, which suggests that the central black hole
masses do not play an important role in the evolutionary
sequence of blazars.
Acknowledgements. This work is partially supported by the
National 973 project (NKBRSF G19990754), the National
Science Fund for Distinguished Young Scholars (10125313),
and the Fund for Top Scholars of Guangdong Province (Q
02114). I thank the anonymous referee for the constructive
suggestions and comments, Prof Jiansheng Chen and Prof.
Youyuan Zhou for suggestions, Dr. Alok C. Gupta for review-
ing the language for me, Dr. Hongguang Wang for useful dis-
cussion. I also thank the Guangzhou City Education Bureau,
which supports our research in astrophysics and the Chinese
Academy of Sciences for the support for advanced visiting
scholars.
References
Barth, A.J., Ho, L.C., Sargent, W.L.W., 2002, ApJL, 566, L13
Becker P., Kafatos, M. 1995, ApJ, 453, 83
Bednarek W. 1993, A&A, 278, 307
J.H. Fan: The basic parameters of γ-ray-loud blazars
7
Courvoisier, T. J.-L.; Robson, E. I., Hughes, D. H., Blecha, A.,
Bouchet, P., et al. 1988, Nat, 335,683
Kranich, D., De Jager, O.C., Kestel, M., et al. 1999, Proc. of
26th ICRC(Slat Lake City), Vol. 3, p346
Dermer C.D., Schlickeiser R., Mastichiadis A. 1992, A&A
256,L27
D'Elia, V., & Cavaliere, A., 2000, PASP, 227, 252
Dibai, E. A., 1984, SvA, 28, 245
Dondi, L., Ghisellini, G., 1995, MNRAS, 273, 583
Fan J.H. 2002, PASJ, 54, L15
Fan J.H., Xie G.Z., Bacon, R. 1999, A&AS, 136, 13
Fan J.H., Adam G., Xie G.Z. et al. 1998, A&A 338, 27
Ferrarese, L. & Merritt, D. 2000, ApJ, 539L, L9
Ferrarese, L., Pogge, R. W., Peterson, B. M., et al. 2001, ApJL,
555, L79
Fossati, G., Maraschi, L., Celotti, A., Comastri, A., Ghisellini,
G. 1998, MNRAS, 299, 433
Franceschini, A., Vercellone, S., & Fabian, A.C., 1998,
MNRAS, 297, 817
Gebhardt, K.,Kormendy, J., Ho, L. C., et al. 2000, ApJL, 543,
L5
Genzel, R., Eckart, A., Ott, T., Eisenhauer, F., 1997, MNRAS,
291, 219
Ghisellini G., Celotti A., Fossati G., et al. 1998, MNRAS 301,
451
Hartman, R.C., Bertsch, D. L., Chen, A. W., et al. 1999, ApJS,
123, 79
Hartman, R.C., Webb J.R., Marscher A.P. et al. 1996, ApJ,
461, 698.
Hartman, R. C., 1996, ASP Conf. Ser. 110, p33
Hayashida, N., Hirasawa, H., Ishikawa, F., et al. 1998, ApJ,
504L, L71
Ho, L.C., 2002, ApJ, 564, 120
Kaspi, S., Smith, P. S., Netzer, H., Maoz, D., Jannuzi, B. T.,
Giveon, U., 2000, ApJ, 533, 631
Kataoka, J., Mattox, J. R., Quinn, J., et al. 1999, APh, 11, 149
Kniffen D.A., Bertsch D.L., Fichtel C.E. et al. 1993, ApJ 411,
133
Komossa, S., Burwitz, V., Hasinger, G., et al. 2003, ApJL, 582,
L15
Kormendy, J., & Richstone, D., 1995, ARA&A, 33, 581
Kormendy, J., & Gehbardt, K., 2001, Proceedings of the 20th
Texas Symposium on Relativistic Astrophysics, eds. H.
Martel & J.C. Wheeler.
Kranich, D., Mirzoyan, R., Petry, D., et al. 1999, APh, 12, 65
Laor, A., 1998, ApJ, 505L, 83
Lawson, A. J. & McHardy, I. M., 1998, MNRAS, 300, 1023
Madejski, G., Takahashi, T., Tashiro, M., et al. 1996, ApJ, 459,
156
Magorrian, J., Tremaine, S., Richstone, D., et al. 1998, AJ,
115, 2285
Mattox J.R., Bertsch D.L., Chiang J. et al. 1993, ApJ 410, 609.
Mattox J.R., Wagner S.J., Malkan M., et al. 1997, ApJ 476,
692
Fig. 3. Histogram of black hole mass for BLs(dotted lines)
and FSRQs (solid lines) for the case of λ = 0.1.
Bloom S.D., Bertsch D.L., Hartman R.C. et al. 1997, ApJ 490,
L145
Bottcher, M., & Dermer, C.D., 2002, ApJ, 564, 86
Cao, X.W. 2002, ApJL, 570, L13
Cheng K.S., Fan, J.H., Zhang, L., 1999, A&A, 352, 32
Chiappetti, L., Maraschi, L., Tavechio, F., et al. 1999, ApJ,
521, 552
Ciaramella, A., Bongardo, C., Aller, H. D., et al. 2004, A&A,
McLure, R. J. & Dunlop, J. S., 2001, MNRAS, 327, 199
Merritt, D., & Ferrarese, L., 2001, in: The Central Kiloparsec
of Starbursts and AGNs, eds. J.H. Knapen, J.E. Beckman
et al., ASP Conf. Proc. 249, p335.
Padovani, P. & Rafanelli, P., 1988, A&A, 205, 53
Perlman E.S., Stocke, J. T., Schachter, J. F., et al. 1996, ApJS
419, 485
104, 251
Comastri, A., Fossati, F., Ghisellini, G., et al. 1997, ApJ, 480,
534
Pica A.J., Smith, A. G., Webb, J. R., et al. 1988, AJ, 96, 1215
Quinn, J., Akerlof, C.W., Biller, S. et al. 1996, ApJ, 456, L83
8
J.H. Fan: The basic parameters of γ-ray-loud blazars
Rieger, F. M. & Mannheim, K., 2000, A&A, 359, 948
Rieger, F. M. & Mannheim, K., 2003, A&A, 397, 121
Sambruna, R. M., Maraschi, L., Urry, C. M., 1996, ApJ, 463,
444
Scarpa, R., Falomo, R., 1997, A&A, 325, 109
Sillanpaa, A., Haarala, S., Valtonen, M. J., wt al. 1988, ApJ,
325, 628
Stacy J.C., Vestrand W.T., Sreekumar P. et al. 1996, A&AS
120, 549
Urry C.M., Treves, A., Maraschi, L., et al. 1997, ApJ, 486, 799
Vestergaard, M., 2002, ApJ, 571, 733
Villata, M., Raiteri, C. M., 1999, A&A, 347, 30
von Montigny C., Bertsch D.L., Chiang J. et al. 1995, ApJ 440,
525
Wandel, A. & Yahil, A., 1985, ApJ, 295L, 1
Wandel, A., Peterson, B.M., & Malkan, M.A., 1999, ApJ, 526,
579
Wang, J.M., Xue, S.J., Wang, J.C., 2001, ApJ(submitted),
(astro-ph/0111209)
Wehrle A.E., Pian E., Urry C.M. et al. 1998, ApJ 497, 178
Wu, X.B., Liu, F.K., Zhang, T.Z., 2002, A&A, 389, 742
Zhang L., Cheng K.S. 1997, ApJ 475, 534
4
3
2
1
u
a
T
M=130
M=95
M
=60
0
0
5
10
Phi
15
6
N
4
2
0
6
7
8
Log M
9
|
0711.3212 | 2 | 0711 | 2007-11-29T02:51:49 | Limits on Hot Galactic Halo Gas from X-ray Absorption Lines | [
"astro-ph"
] | Although the existence of large-scale hot gaseous halos around massive disk galaxies have been theorized for a long time, there is yet very little observational evidence. We report the Chandra and XMM-Newton grating spectral detection of OVII and NeIX Kalpha absorption lines along the sight-line of 4U 1957+11. The line absorption is consistent with the interstellar medium in origin. Attributing these line absorptions to the hot gas associated with the Galactic disk, we search for the gaseous halo around the Milky Way by comparing this sight-line with more distant ones (toward X-ray binary LMC X-3 and the AGN Mrk 421). We find that all the line absorptions along the LMC X-3 and Mrk 421 sight-lines are attributable to the hot gas in a thick Galactic disk, as traced by the absorption lines in the spectra of 4U~1957+11 after a Galactic latitude dependent correction. We constrain the OVII column density through the halo to be N(OVII) < 5E15 cm^{-2} (95% confidence limit), and conclude that the hot gas contribution to the metal line absorptions, if existing, is negligible. | astro-ph | astro-ph |
Accepted for publication in the ApJ Letters
Preprint typeset using LATEX style emulateapj v. 08/22/09
LIMITS ON HOT GALACTIC HALO GAS FROM X-RAY ABSORPTION LINES
Y. YAO1, M. A. NOWAK1, Q. D. WANG2, N. S. SCHULZ1, AND C. R. CANIZARES1
Accepted for publication in the ApJ Letters
ABSTRACT
Although the existence of large-scale hot gaseous halos around massive disk galaxies have been theorized
for a long time, there is yet very little observational evidence. We report the Chandra and XMM-Newton
grating spectral detection of O VII and Ne IX Kα absorption lines along the sight-line of 4U 1957+11. The
line absorption is consistent with the interstellar medium in origin. Attributing these line absorptions to the
hot gas associated with the Galactic disk, we search for the gaseous halo around the Milky Way by comparing
this sight-line with more distant ones (toward LMC X -- 3 and the AGN Mrk 421). We find that all the line
absorptions along the LMC X-3 and Mrk 421 sight-lines are attributable to the hot gas in a thick Galactic disk,
as traced by the absorption lines in the spectra of 4U 1957+11 after a Galactic latitude dependent correction.
We constrain the O VII column density through the halo to be NOVII < 5 × 1015 cm- 2 (95% confidence limit),
and conclude that the hot gas contribution to the metal line absorptions, if existing, is negligible.
Subject headings: Galaxy: halo -- Galaxy: structure -- X-rays: individual (4U 1957+11, Mrk 421, LMC X -- 3)
1. INTRODUCTION
Many semi-analytic calculations and numerical simulations
for disk galaxy formation predict the existence of extended
hot gaseous halos around massive spirals due to the accretion
of the intergalactic medium. For the Milky Way, the gas tem-
perature can be shock-heated to ∼ 106 K at the virial radius
(∼ 250 kpc; e.g., Birnboim & Dekel 2003; Fukugita & Pee-
bles 2006), and the total mass contained in the large scale halo
can be comparable with or even greater than the total baryonic
mass of stars and the interstellar medium in the Galactic disk
(e.g., Sommer-Larsen 2006). Clearly, an observational mea-
surement of such an extended halo will provide an important
test of galaxy formation theories.
Searching for X-ray emission from large scale halos around
nearby disk galaxies has proven unsuccessful. Extraplanar X-
ray emissions have indeed been routinely detected around a
number of nearby spirals, but only on scales of several kpc
(except for the star burst galaxies), and most likely as the
result from on-going stellar feedback in galactic disks (e.g.,
Tüllmann et al. 2006; Li et al. 2006). The X-ray surface
brightness of a galactic halo must be very weak at large scales,
at least partly because of the density square dependence of the
emission and the possible low metallicity of the gas. So far,
the only claimed detection of an apparent X-ray-emitting halo
on a scale of ∼ 20 kpc is around the quiescent edge-on disk
galaxy NGC 5746 (Pedersen et al. 2006).
The large-scale hot gaseous halo around the Milky Way is
indirectly evidenced by the presence of high velocity clouds
(HVCs), in particular, the detection of their associated O VI
line absorption (e.g., Spitzer 1956; Sembach et al. 2003). Be-
cause the O VI-bearing gas is preferentially populated at in-
termediate temperatures of ∼ 3 × 105 K where thermally un-
stable cooling occurs, it is believed to be produced at the in-
terfaces between the cold/warm and the hot media. Without
knowing the temperature distribution and the metallicity of
the hot gas, however, it is very hard to reliably estimate its
1 Massachusetts Institute of Technology (MIT) Kavli Institute for Astro-
physics and Space Research, 70 Vassar Street, Cambridge, MA 02139; yaoys,
mnowak, nss, and [email protected]
2 Department of Astronomy, University of Massachusetts, Amherst, MA
01003; [email protected]
total mass.
The best way to directly measure the hot medium is through
its X-ray absorption line features.
Indeed, O VII, O VIII,
and/or Ne IX absorption lines consistent with zero velocity
shifts (cz ∼ 0) have been detected unambiguously toward sev-
eral bright extragalactic sources (e.g., LMC X -- 3, Mrk 421,
and PKS 2155-304; Nicastro et al. 2002; Fang et al. 2003;
Wang et al. 2005; Yao & Wang 2007a, YW07a hereafter).
But these sight-lines pass through the Galactic disk, the ex-
tended Galactic halo, and (for those AGNs) the possible warm
hot intergalactic medium (WHIM) in the Local Group. Re-
cently, it has been argued that the bulk of these absorp-
tions can not be produced in the WHIM (Fang et al. 2006;
Yao & Wang 2007a; Bregman & Lloyd-Davies 2007).
In-
deed, these highly ionized absorption lines have also been de-
tected in spectra of many Galactic sources (e.g., Yao & Wang
2005; Juett et al. 2006), clearly indicating the existence of the
hot gas in the Galactic disk. All existing X-ray absorption
and emission data are consistent with the hot gas located in
the Galactic disk with a vertical exponential scale height of
∼ 2 kpc (Yao & Wang 2005, 2007a), comparable to those in-
ferred from the angular distributions of column densities of
the O VI-bearing gas and pulsar dispersion measures (Savage
et al. 2003; Berkhuijsen et al. 2006). Clearly, the constraint
on the location of the hot gas depends on the modeling. For
instance, Bregman & Lloyd-Davies (2007) found that the ab-
sorbing gas toward AGNs are also consistent with a uniformly
spherical distribution within a radius of 20 kpc. The questions
here are, whether the extended Galactic halo (defined as the
hot gas in the region of >10 kpc beyond the Galactic plane but
within the Galactic virial radius) exists, and how much does it
contribute to the observed highly ionized X-ray absorptions?
In this Letter, we present a differential study searching for
such a large-scale hot gaseous halo around our Galaxy by
comparing observations between sight-lines of 4U 1957+11
and Mrk 421, and between those of LMC X -- 3 and Mrk 421.
Ne IX, O VII, and/or O VIII absorption lines at cz ∼ 0 have been
detected in spectra of all three sources observed with Chandra
and/or XMM-Newton X-ray Observatories. These compar-
isons enable us to probe the Galactic halo by examining its
contribution to the observed absorption lines, and then to es-
2
Yao et al.
timate its total mass and/or its metallicity.
Throughout the Letter, we adopt the solar abundances from
Wilms et al. (2000), quote statistical errors (or upper limits)
for single floating parameters at 90% (95%) confidence lev-
els, and assume that the hot gas in both the Galactic disk and
halo is in the collisional ionization equilibrium state and is ap-
proximately isothermal. We further assume that the disk gas
is distributed following an exponential decay law with a ver-
tical scale height of 2 kpc (e.g., Savage, et al. 2003; YW07a).
All the data analyses are performed with the software package
XSPEC (version 11.3.2).
2. OBSERVATIONS AND DATA REDUCTION
4U 1957+11 (V1408 Aquilae; l,b = 51.◦31,- 9.◦33) is a per-
sistent low mass X-ray binary (Nowak & Wilms 1999 and
references therein). Chandra observed this source for 67 ks
on 2004 September 7 (ObsID 4552) with the High Energy
Transmission Grating Spectrometer (HETGS; Canizares et
al. 2005) and XMM-Newton observed it for 45 ks on 2004
October 16 (ObsID 206320101).
We reprocessed both Chandra and XMM-Newton obser-
vations following the standard procedures. Using the soft-
ware package CIAO 3 and calibration database (CALDB) ver-
sion 3.4, we re-calibrated the Chandra observation, extracted
grating spectra, and calculated corresponding instrumental re-
sponse files (RSPs). In this study, we only use the first or-
der spectra from the medium energy grating because of its
large effective area at longer wavelengths (> 13 Å). To fur-
ther enhance the counting statistics, we co-added the positive-
and the negative-grating spectra and RSPs. For the XMM-
Newtonobservation, we used the software package SAS(ver-
sion 6.50) 4 to remove those events contaminated with back-
ground flares (by screening out the time intervals with an
event count rate > 0.4 counts/s on the CCD9), and then ex-
tracted spectra and calculated RSPs from the Reflection Grat-
ing Spectrometer (RGS) events by running the thread rgsproc.
Because of the failure of the CCD4 in the RGS2, we only use
the RGS1 spectrum.
In order to measure absorption line properties we only rely
on the nearby continuum levels. We therefore use several
parts of the spectra that are local to the lines of our inter-
est. For the Chandra spectrum, we use a wavelength range
of 12.0 -- 22.5 Å covering the Kα and/or Kβ transition lines of
O VII, O VIII Ne VIII, Ne IX, and Ne X. For the case of the XMM-
Newtonspectrum, the RGS1 has no effective area around the
Ne IX Kα line and there is a bad pixel near the O VIII Kα line
at 18.969 Å. Here we use ranges of 18.0 -- 18.9 and 19.2 -- 22.5
Å covering the O VII Kα and Kβ lines.
Mrk 421, a bright quasar at z = 0.03 (l,b = 179.◦83,65.◦03),
is a Chandra calibration target and has been observed with
the high resolution grating instruments multiple times; most
of the observations have been reported in YW07. In this Let-
ter, we use the same observations, and the same spectra and
RSPs as in YW07. To fit the spectra without worrying about
overlapping spectral orders, for each of the observations taken
with High Resolution Camera (HRC; Table 1 in YW07), we
derive the first order spectrum with minimal higher (> 1) or-
der confusion via the following steps. We first fit the broad-
band (1.5-35 Å) order-overlapped spectrum with the order
combined RSP, and then multiply the spectral counts chan-
3 http://cxc.harvard.edu/ciao/
4 http://xmm.vilspa.esa.es/sas/current/documentation/threads
nel by channel with the ratio between the model predicated
counts based on the first order RSP and that based on the
order-combined RSP. We then combine all the first order spec-
tra of different observations using the same procedure as in
YW07.
In YW07, we have reported the O VII Kα, Kβ, and O VIII
Kα absorption lines in the spectrum of Mrk 421 and measured
their equivalent widths (EWs; see also Williams et al. 2005;
Kaastra et al. 2006). In fact, the Ne IX Kα line has also been
detected in the spectrum (Williams et al. 2005). However, we
find that, due to an instrumental feature near the line, measur-
ing the amount of the Ne IX absorption is severely affected by
how the continuum is placed. To avoid such an uncertainty,
in our analysis we only use the spectral range of 18 -- 22.5 Å,
covering the oxygen lines as mentioned above.
LMC X -- 3 (l,b = 273.◦58,- 32.◦08) is located ∼ 50 kpc away
from the Sun and Chandra observed it with the Low Energy
Transmission Grating Spectrometer plus HRC for ∼ 100 ks
in total. The interstellar O VII and Ne IX Kα absorption lines
observed in this sight-line have been reported by Wang et al.
(2005). In this Letter, we use the same spectra and RSPs as
in Wang et al. . Following the same procedure as for Mrk 421
spectra (see above), we obtained the co-added first order spec-
trum and the RSP. We use the spectral range of 12.0 -- 22.5 Å
in the following analysis.
3. ANALYSIS AND RESULTS
We focus our efforts on searching for and analyzing the
highly ionized absorption lines at cz ∼ 0 that are likely pro-
duced in the hot ISM (§ 4). The Ne IX (13.447 Å) and O VII Kα
(21.602 Å) lines consistent with cz ∼ 0 appear in the Chandra
and XMM-Newton spectra at ∼ 3 and 4σ significance levels
(Fig. 1), respectively. Modeling these two lines with nega-
tive Gaussian models, we measure their EWs as 6.3(3.9, 8.8)
and 18.7(8.3, 30.6) mÅ. Note that the EWs measured here are
for reference only. The ionic column density and the disper-
sion velocity of the absorbing gas will be measured in below.
No other highly ionized O and Ne lines are detected at >
∼2σ
significance. Fixing the line centroid at the rest frame wave-
length (Verner et al. 1996; Behar & Netzer 2002), we obtain
the EW upper limits of Ne X(12.134 Å), Ne VIII (13.646 Å) and
O VIII Kα (18.969 Å), and O VII Kβ (18.629 Å) lines as 1.2,
2.8, 15.6, 10.1 mÅ, respectively. Since all these line are un-
resolved, we fix the line width σ to 100 km s- 1 (please refer
to vb in Table 1) during these measurements and we use ei-
ther the Chandra or XMM-Newton spectrum, whichever has
higher counting statistics near the line (Fig. 1).
A comparison of the line absorption between 4U 1957+11
and Mrk 421 sight-lines provides an effective probe of the
large scale Galactic hot gaseous halo. The distance of
4U 1957+11 is estimated to be D ∼ 10 - 25 kpc (M. Nowak
et al. in preparation), a location of 1.6 -- 4.1 kpc below the
Galactic disk plane where it samples 55-87% of the Galactic
disk gas in the vertical direction (§ 1) along the line of sight.
In comparison, the line of sight towards Mrk 421 samples the
hot gas not only in the Galactic disk but also in the putative
Galactic halo (§ 1). Therefore the differential absorptions be-
tween these two sight-lines allow us to directly constrain the
absorption contribution from the halo.
We first characterize the hot absorbing gas towards these
two sight-lines by using our absorption line model, absline.
This model adopts the the Voigt function to approximate an
individual line profile, and uses the line centroid El, veloc-
Limits on Hot Galactic Halo Gas
3
FIG. 1. -- The oxygen and neon absorption lines detected in the spectra
of 4U 1957+11. The upper plots in panels (b) and (c) are from the XMM-
Newton spectrum; the others are from the Chandra spectrum. The vertical
dotted lines mark zero velocity. The bin-size is 22.2 mÅ and 10.0 mÅ for
XMM-Newton and Chandra spectra, respectively.
TABLE 1
LINE ANALYSIS RESULTS
T
(106 K)
vb
(km s- 1)
64(48, 104)
155(70, 301)
79(62, 132)
70(50, 172)
NOVII
(1015 cm- 2)
10.0(6.6, 14.7)
3.1(1.8, 8.2)
11.7(7.4, 18.5)
7.0(3.2, 12.7)
NHAO
(1019 cm- 2)
1.4(1.0, 2.0)
1.1(0.6, 1.8)
2.3(1.5, 3.7)
1.4(0.7, 2.2)
· · ·
· · ·
1.8(1.6, 2.0)
1.4(1.3, 1.6)
2.2(1.6, 2.7)
1.3(0.8, 2.0)
1.8(1.7, 2.1)
Mrk 421a
4U 1957a
LMC X3a
4U 1957b
Haloc
LMC X3b
Haloc
NOTE. -- The uncertainty ranges (upper limits) are quoted at 90% (95%)
confidence levels. The dependence of the column density on the Galactic
latitude (sinb factor) has been corrected with respect to Mrk 421 direction.
AO is the oxygen abundance in unit of the solar value. a Results from fitting
lines in each direction separately. b Results from a joint analysis of lines in
the source and those in Mrk 421. c The absorption contribution from the halo
gas beyond the source. See text for detail.
8.8(5.8, 12.7)
1.7(1.2, 2.2)
64(49, 108)
< 0.9
< 0.7
< 4.8
< 3.7
· · ·
· · ·
ity dispersion vb, absorbing gas temperature T , and reference
ionic column density NX (e.g., X=O VII or H) as fitting pa-
rameters.
It therefore can be used to conduct a joint anal-
ysis of multiple absorption lines (including non-detections).
5 For the sight-line toward Mrk 421, we use O VII Kα, Kβ,
and O VIII Kα absorption lines from the Chandra spectrum
(YW07). For sight-line toward 4U 1957+11, we use O VII,
O VIII, Ne VIII, Ne IX, Ne X Kα, and O VII Kβ lines in the
Chandraand XMM-Newtonspectra where applicable, and as-
sume the neon to oxygen abundance ratio to be the solar value
(Yao & Wang 2006). From the joint analysis, we obtain the
T , vb, and NOVII (or its equivalent NH for a given metallicity),
as reported in Table 1. For ease of comparison, we take into
account the dependence of the column density on the Galactic
latitude (sinb factor; § 1) for the 4U 1957+11 sight-line with
respect to the Mrk 421 direction (Table 1).
Please note that in the above characterization we assume
that the intervening hot gas in both sight-lines is isothermal.
In reality, the temperature of both the disk and the halo gas
could be a function of the vertical off-plane distance, the ra-
dial off-Galactic center distance, or a combination of both
(e.g., Toft et al. 2002; YW07). This simplified description
aims to provide a direct comparison of the amount of absorp-
tion between the two sight-lines.
As demonstrated in Table 1 and Figure 2, the absorption
5 Please refer to Yao & Wang (2005, 2006) for the model and the joint
analysis procedure.
FIG. 2. -- The observed Mrk 421 spectrum (cross) around the oxygen ab-
sorption lines and the best fit model (green histogram) convolved with the
instrumental response. The red histograms mark the amount of absorption
predicted for the Mrk 421 direction from the best fit of the lines observed in
the spectra of 4U 1957+11 (leftpanels) and LMC X -- 3 (rightpanels).
towards 4U 1957+11 can account for the bulk of the absorp-
tion towards Mrk 421. To quantify the absorption contribution
from the putative Galactic halo gas, we jointly analyze the ab-
sorption lines in these two directions, "subtract" the absorp-
tion spectrum toward 4U 1957+11 direction from that toward
Mrk 421 direction, and attribute the residual absorption to the
halo gas. We find that the halo gas is consistent with no con-
tribution to the observed line absorptions. Assuming that the
halo gas has the same thermal (e.g., T ) and dynamic (e.g.,
vb) properties as the disk gas, we estimate the column den-
sity upper limit of the halo gas as NOVII < 4.8 × 1015 cm- 2 (or
NHAO < 9 × 1018 cm- 2, where AO is the oxygen abundance in
unit of the solar value; Table 1). The halo gas and the disk gas
may have different T ; for the halo gas over a broad tempera-
ture range of 0.4 - 2.5 × 106 K, we find NHAO < 1019 cm- 2.
Following the same procedure as above, we characterize the
line absorptions toward LMC X -- 3 (Table 1; see also Wang et
al. 2005), and then search for the absorption feature of the
halo gas beyond LMC X -- 3 by comparing it with Mrk 421.
Again, we find that the line absorptions toward LMC X -- 3 can
account for all the absorptions observed toward the Mrk 421
sight-line (Fig. 2), and obtain the upper limits of the col-
umn density of the halo by jointly analyzing the sight-lines
of LMC X -- 3 and Mrk 421 (Table 1).
4. DISCUSSION
In this study, we have assumed that the absorption lines
observed in the spectra of 4U 1957+11 are produced in the
ISM. However, the photo-ionized in- or out-flow material in-
trinsic to 4U 1957+11 could contaminate these lines. From
the best fit to the Chandra spectrum in the range of 2 -- 23
Å, we estimate the source luminosity over 0.5 -- 10 keV as
10kpc × 1037 ergs s- 1, where D10kpc is the source dis-
L = 1.19D2
tance in units of 10 kpc. A curve of growth analysis with a
dispersion velocity (vb) of 0 km s- 1 gives a firm upper limit
of the Ne IX column density to be NNeIX < 6.6 × 1017 cm- 2
(assuming the solar abundance of Ne). This is equivalent
to a hydrogen column density NH < 1.1 × 1022 cm- 2, as-
suming an ionization fraction of 0.5 for Ne IX. For the ab-
sorber, the ionization parameter U = Lx/n(r)r2 must satisfy
log(U) ≤ 2.5, since the Ne X Kα absorption line has not been
detected (Kallman & McCray 1982). Further assuming a ra-
dial gas density distribution n(r) ∝ (rw/r)2, where rw is the lo-
4
Yao et al.
cation where wind launches and r > rw, we obtain rw > 50R⊙,
which is much larger than the star separation ( <
∼6R⊙) of the
system (estimated from its orbital period 9.33 hours and tak-
ing an upper limit of the X-ray compact object MX < 16M⊙).
These estimates make the in-flow scenario very unlikely. On
the other hand, if the absorber is out-flow material, the lines
are expected to be blue-shifted at velocities larger than the
escape velocity of the system (e.g., GRO J1655 -- 40; Miller
et al. 2006), which is 260 km s- 1 for 4U 1957+11 (taking
MX = 10M⊙ and rw = 6R⊙). In contrast, the detected Ne IX
Kα velocity is 0 ± 114 km s- 1. The above arguments suggest
that the absorption lines are more likely produced in the ISM
(see also Juett et al. 2006), although the intrinsic scenario can
not be completely ruled out.
We have compared the highly ionized line absorptions of
O VII, O VIII, and Ne IX observed in the spectra of 4U 1957+11
and Mrk 421, and in the spectra of LMC X -- 3 and Mrk 421 to
search for the large scale Galactic halo. We found that there is
no significant X-ray absorption due to hot gas beyond either
4U 1957+11 or LMC X -- 3. These results are consistent with
our previous conclusion that bulk of the absorptions observed
toward the Mrk 421 sight-line are due to the hot gas around the
Galactic disk with a scale height of ∼ 2 kpc based on a joint
analysis of the absorption and emission data (YW07). Note
that the upper limits obtained toward LMC X -- 3 are slightly
lower than those toward 4U 1957+11 (Table 1). Some of
the difference are due to the partial sampling of the Galac-
tic disk by the latter sight-line, and the better spectral quality
of LMC X -- 3 also matters.
Large scale hot gaseous halos are commonly expected in
modern disk formation models for massive spirals (e.g. White
& Frenk 1991). For the Milky Way in particular, the hot
gaseous halo can extend up to rvirial ∼ 250 kpc and the to-
tal mass contained is expected to be ∼ 6 × 1010 M⊙, which is
currently missing in the baryon mass inventory in the Lambda
Cold Dark Matter (ΛCDM) cosmology (Dehnen & Binney
1998; Klypin et al. 2002; Sommer-Larsen 2006). Recently,
Maller & Bullock (2004) assumed that
the fragmentated
warm clouds could be formed out of the hot gas halo and
proposed a multiphase cooling scenario for the galaxy for-
mation. This model alleviates the so-called "over-cooling"
problem faced by the standard cooling model (e.g., Klypin et
al. 2002). They then obtained a density profile for the resid-
ual hot halo gas as a function of radius (Eq. 21 and Fig. 4
in Maller & Bullock 2004). Normalizing their density profile
by requiring the total column density integrated from 10 to
250 kpc to be NHAO < 1019 cm- 2 (Tabel 1), we obtain the to-
tal mass contained in the halo as MH < 1.2/AO × 1010 M⊙.
Clearly,
to match this upper limit with the total missing
baryon mass in the MW, AO should be < 20% of the solar
value. More recently, Hansen & Sommer-Larsen (2006) as-
sumed that gaseous halos of galaxies are in a hydrostatic equi-
librium state and that the gas density and the total mass den-
sity profiles are power laws, and then derived a hot halo gas
density profile for the MW-like galaxy (Fig. 2 in Hansen &
Sommer-Larsen 2006). Again, normalizing their profile we
obtain the total hot gas mass as < 2.2/AO × 109 M⊙, requir-
ing AO to be < 3.7% of the solar value.
Low metallicity of the hot halo gas is not surprising; the
halo gas is believed to be primarily accreted from the inter-
galactic medium that is expected to be metal poor. This would
also suggest that the star formation and the associated stellar
feedback via Type II supernovae (SNeII) that occurred in the
Galactic disk and bulge are not important in regulating the
metal content of the halo. Hot gas resulting from SNeII, if
leaking into the halo (e.g., Mac Low & Ferrara 1999), may
cool quickly enough (via adiabatic expansion and atomic line
radiation) to form metal-rich cold clouds in the local environ-
ment before homogenizing with the halo gas (Wang 2007). As
these clouds return back to the disk plane as Galactic foun-
tains, their interaction with the embedding hot gas could be
another mechanism of warm cloud formation as required in
multiphase cooling scenarios (Maller & Bullock 2004). The
interfaces between these two media may harbor the bulk of
HVCs as observed via O VI absorptions at large distance (e.g.,
Sembach et al. 2003).
In this Letter, we derive the mass and metallicity limits on
the putative Galactic halo by comparing the observed absorp-
tion lines among the sight-lines of 4U 1957+11, LMC X -- 3,
and Mrk 421. However, the amount of absorptions observed
along different AGN sight-lines could vary substantially (e.g.,
Bregman & Lloyd-Davies 2007), which may be due to ad-
ditional absorption components (e.g., 3C 273; Yao & Wang
2007b) and/or different absorbing gas properties. More com-
parisons among other sight-lines are therefore needed to con-
firm the presented results.
We are grateful to Li Ji and the anonymous referee for
their insightful comments, which helped to improve the pre-
sentation of the paper. This work is supported by NASA
through the Smithsonian Astrophysical Observatory contract
SV3-73016 to MIT for support of the Chandra X-Ray Cen-
ter under contract NAS 08-03060. Support from SAO/CXC
grants AR6-7023 and AR7-8014 are also acknowledged.
Behar, E., & Netzer, H. 2002, ApJ, 570, 165
Berkhuijsen, E., et al. 2006, AN, 327, 82
Birnboim, Y., & Dekel, A. 2003, MNRAS, 345, 349
Bregman, J. N., & Lloyd-Davies, E. J. 2007, ApJ in press,
astro-ph/07071699
Canizares, C. R., et al. 2005, PASP, 117, 1144
Dehnen, W., & Binney, J., 1998, MNRAS, 294, 429
Fang, T., et al. 2003, ApJ, 586, 49
Fang, T., et al. 2006, ApJ, 644, 174
Fukugita, M., & Peebles, P. 2006, ApJ, 639, 590
Hansen, S., & Sommer-Larsen, J. 2006, ApJ, 653, L17
Juett, A., et al. , 2006, ApJ, 648, 1066
Kaastra, J, S., Werner, N., den Herder, J. W. A., et al. 2006, 652, 189
Kallman, T. R., & McCray, R. 1982, ApJS, 50, 263
Klypin, A., Zhao, H., & Sommerville, R. S. 2002, ApJ, 573, 597
REFERENCES
Li, Z., et al. 2006, MNRAS, 371, 147
Maller, A. H., & Bullock, J. 2004, MNRAS, 355, 694
Mac Low, M., & Ferrara, A. 1999, ApJ, 513, 142
Miller, J., et al. 2006, Nature, 441, 953
Nicastro, F., et al. 2002, ApJ, 573, 157
Nowak, M. A., & Wilms, J. 1999, ApJ, 522, 476
Pedersen, K., et al. 2006, New A., 11, 465
Savage, B., et al. 2003, ApJS, 146, 125
Sembach, K., et al. 2003, ApJS, 146, 165
Sommer-Larsen, J. 2006, 644, L1
Spitzer, L. 1956, ApJ, 124, 20
Toft, S., et al. 2002, MNRAS, 335, 799
Tüllmann, R, et al. 2006, AA, 448, 43
Verner, D, A., et al. 1996, Atomic Data & Nuclear Data Tables, 64, 1-180
Wang, Q. D., et al. 2005, ApJ, 635, 386
Limits on Hot Galactic Halo Gas
5
Wang, Q. D. 2007, EAS, 24, 59
Williams, R., Mathur, S., & Nicastro, F., et al. 2005, ApJ, 631, 856
Wilms, J., Allen, A., & McCray, R. 2000, ApJ, 542, 914
White, S. D. M., & Frenk, C. S. 1991, ApJ, 379, 52
Yao, Y., & Wang, Q. D., 2005, ApJ, 624, 751
Yao, Y., & Wang, Q. D., 2006, ApJ, 641, 930
Yao, Y., & Wang, Q. D., 2007a, ApJ, 658, 1088
Yao, Y., & Wang, Q. D., 2007b, ApJ, 666, 242
|
astro-ph/9510010 | 2 | 9510 | 1996-04-22T12:40:49 | Chaos, Fractals and Inflation | [
"astro-ph",
"gr-qc",
"hep-th"
] | In order to draw out the essential behavior of the universe, investigations of early universe cosmology often reduce the complex system to a simple integrable system. Inflationary models are of this kind as they focus on simple scalar field scenarios with correspondingly simple dynamics. However, we can be assured that the universe is crowded with many interacting fields of which the inflaton is but one. As we describe, the nonlinear nature of these interactions can result in a complex, chaotic evolution of the universe. Here we illustrate how chaotic effects can arise even in basic models such as homogeneous, isotropic universes with two scalar fields. We find inflating universes which act as attractors in the space of initial conditions. These universes display chaotic transients in their early evolution. The chaotic character is reflected by the fractal border to the basin of attraction. The broader implications are likely to be felt in the process of reheating as well as in the nature of the cosmic background radiation. | astro-ph | astro-ph | Chaos, Fractals and Inflation
CWRU-P12-95, CFPA-95-TH-26
Neil J. Cornish† and Janna J. Levin⋆
†Department of Physics, Case Western Reserve University
Cleveland, OH 44106-7079
⋆Center for Particle Astrophysics, UC Berkeley
301 Le Conte Hal l, Berkeley, CA 94720-7304
Abstract
In order to draw out the essential behavior of the universe, investigations of
early universe cosmology often reduce the complex system to a simple inte-
grable system. Inflationary models are of this kind as they focus on simple
scalar field scenarios with correspondingly simple dynamics. However, we can
be assured that the universe is crowded with many interacting fields of which
the inflaton is but one. As we describe, the nonlinear nature of these inter-
actions can result in a complex, chaotic evolution of the universe. Here we
illustrate how chaotic effects can arise even in basic models such as homoge-
neous, isotropic universes with two scalar fields. We find inflating universes
which act as attractors in the space of initial conditions. These universes
display chaotic transients in their early evolution. The chaotic character is
reflected by the fractal border to the basin of attraction. The broader impli-
cations are likely to be felt in the process of reheating as well as in the nature
of the cosmic background radiation.
05.45.+b, 95.10.E, 98.80.Cq, 98.80.Hw
6
9
9
1
r
p
A
2
2
2
v
0
1
0
0
1
5
9
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
Typeset using REVTEX
1
I. INTRODUCTION
The inflationary paradigm strives to deliver a smooth universe from random initial con-
ditions.
If inflation is a robust attractor in the space of initial conditions, then it earns
its claim to naturalness and genericity [1,2]. A universe which hosts many different fields,
including an inflaton candidate, can develop an extreme sensitivity to initial values. This
sensitivity marks the onset of chaos. Chaotic dynamics does not in itself destroy the robust-
ness of an inflationary phase. Rather, it can lead to some powerful and perhaps observable
implications for a realistic universe. For instance, a fractal pattern in the spectrum of den-
sity fluctuations could be generated. Also, the final phase of inflation marked by reheating
would unavoidably be a setting for chaos.
In simple cosmologies, the ultimate fate of the universe can be predicted once a set of
initial conditions is prescribed.
In a closed cosmology for instance, it can be determined
from the initial prescription if the universe inflates or collapses. A plot in phase space will
show regions or basins within which all of the initial conditions lead to the same outcome.
There will be basins of inflation and basins of collapse. If the dynamics is not chaotic, these
basins of attraction1 are distinctly separated by smooth, regular boundaries. If the dynamics
becomes chaotic, then the smooth boundaries begin to break up, ultimately becoming fractal.
The models described in the following sections are chosen on the grounds of simplicity,
and do not necessarily conform to any standard inflationary scenario. In our current models
the primary inflaton is weakly coupled and essentially not dynamical, leaving the chaotic
dynamics to the other scalar fields. Multiply coupled fields are quite natural in any particle
theory. The Higgs field for instance must couple to the standard model fermions in order to
induce fermion masses. In supersymmetric theories, a glut of coupled particles is expected.
The fields behave effectively like nonlinearly coupled harmonic oscillators and so naturally
bring on chaos. In future studies we intend to connect more concretely to specific hybrid
inflationary models [3] and draw out implications for the spectrum of fluctuations or the
end of a realistic model. The main aim of this paper is to illustrate the complex dynamics
that can arise in relatively simple cosmologies.
In addition to highlighting the appearance of chaos in inflationary cosmology, we aim to
demonstrate the power of fractals as a quantitative measure of chaos in relativity. In general
relativity coordinate independent measures of chaos are of vital importance. One of the most
valuable measures of a chaotic system in flat space, the Lyapunov exponents, can be removed
by a simple coordinate transformation in curved space. Thus the usual coordinate dependent
measures of chaos become ambiguous in a relativistic context. Topological signals such as
fractal basins, cantori or stochastic layers in phase space are needed for conclusive evidence
of chaos. In order to search for cantori or stochastic layers it is necessary to construct slices
through phase space known as Poincar´e sections. Since Poincar´e sections rely on quasi-
periodic behaviour, the system must complete many cycles for a useful picture to emerge.
1 Throughout the paper we loosely refer to attractors in phase space. In more formal terminology,
the word attractor is reserved for dissipative systems. Strictly speaking, our attractors are just
asymptotic regions of phase space.
2
Oftentimes relativistic systems are not so obliging as the evolution may end at singularities,
such as the big crunch or inside a black hole. In these cases the dynamics is better suited
to an outcomes based approach such as the study of fractal basin boundaries [4].
In a chaotic system the different possible outcomes will each have a basin of attraction
in the space of initial conditions, with the basins separated by a fractal border. A specific
examination of phase space will require a coordinate system to be chosen. One might
worry that in a different time slicing, the fractal would disappear. This is not possible. A
coordinate transformation must be smooth and differentiable. No smooth map can undo a
truly fractal pattern as fractals are non-differentiable. While the features of a fractal may
be altered by a coordinate transformation, the existence of a the fractal is unambiguous.
The cosmological context we employ allows us to demonstrate the technique of fractal
basin boundaries. What emerges is a definitive manifestation of chaos in cosmology. We have
already commented that we focus on multi-field systems. If only one scalar field is present
and the universe is closed, then there is still the possibility of chaotic regions in phase
space [5–7]. However, in order to generate chaotic dynamics with just one matter field, the
universe must oscillate between expansion and contraction many times. The requirement of
many bounces makes these otherwise interesting solutions unlikely if not truly unphysical.
We consider single field scenarios in §III. Since these bouncing candidates do not represent
viable cosmologies, we turn our attention to many field systems and the demonstration of
fractal basin boundaries in §IV-V. Also note, that while we study closed cosmologies, the
chaotic transients can be seen in a universe which never collapses or bounces. When fields
interact, the chaotic nature is therefore not limited to closed cosmologies. It is thus possible
that there was a transient chaotic epoch in the history of our own universe.
II. THE COSMOLOGICAL MODEL
In the coming examples we consider closed Friedman-Robertson-Walker (FRW) uni-
verses. For potential driven inflation to be successful the inflationary potential needs to
be fairly constant. For our purposes the inflaton can be modeled by a simple cosmological
constant. If this were the complete system there would be of course no chaotic dynamics.
However, the universe is created bursting with matter fields. We model the matter content
by a variety of conformally and minimally coupled fields. These matter fields interact and
can incite chaos. For the inflationary cases at hand then, the chaotic behaviour is principally
matter driven, i.e. chaos in Tµν causing chaotic evolution of gµν .
The physical picture is that of a closed, pre-inflationary universe just exiting the Planck
era. The space of initial conditions is probed by assigning three possible outcomes. Either
the universe inflates forever, inflates for short spurts but then collapses, or collapses without
any inflationary event. The outcome depends on the relative sizes of the various kinetic
and potential energies in the matter fields. As the interaction between the matter fields is
turned up the boundary which separates inflating from noninflating initial conditions blurs,
eventually becoming fractal.
While we only consider closed models in our outcomes based approach, the appearance
of chaotic transients will be generic, regardless of curvature. This becomes clear since chaos
is also nested within a given outcome basin. For instance, a universe will often go through
3
rocky beginnings, enduring many fits of inflation before taking off smoothly or collapsing.
Within a collapse basin, the sensitivity to the initial conditions shows up as a random scatter
in the maximum radius of the universe or in the final value of the fields. If the universe has
managed to inflate by several e-folds there is no turning back as the kinetic energy which
might interrupt inflation quickly redshifts away. This feature of the de Sitter attractor is
often referred to as cosmic baldness. We shall see that while the de Sitter attractor might
end up bald, it can have very hairy beginnings.
We shall consider FRW universes described by the metric
ds2 = dt2 − a2 dr2
1 − kr2 − r2dΩ2! ,
dr2
= a2 dτ 2 −
1 − kr2 − r2dΩ2! ,
where t is cosmic time and τ is conformal time. Throughout, dots will denote derivatives
with respect to cosmic time and dashes will denote derivatives with respect to conformal
time. The matter Lagrangian will contain various combinations of conformally, Ψ, and
minimally, Φ, coupled scalar fields with a variety of interaction terms described by the
potential V (Ψ, Φ):
(1)
1
∂µΦ∂ µΦ −
12 RΨ2 .
LM = V −
For comparison with the conformally coupled term, i.e. the last term in eqn (2), the gravi-
tational Lagrangian is LG = − 1
12 R. We have chosen units where 4πG/3 = c = 1. In terms
of cosmic time t the field equations read
∂µΨ∂ µΨ +
(2)
1
2
1
2
Φ + 3H Φ + ∂ΦV = 0 ,
(3)
Ψ = 0 ,
Ψ + 3H Ψ + ∂ΨV + R
6
a + a 2 Φ2 + ( Ψ + H Ψ)2 +
a2 Ψ2 + Ψ∂ΨV − 2V ! = 0 .
k
The Ricci scalar can be related to the scale factor through R/6 = a/a + H 2 + k/a2 . The
Hubble expansion factor is given by H = a/a. The total energy of the system is
(5)
(4)
k
k
a2 − Φ2 − ( Ψ + H Ψ)2 −
a2 Ψ2 − 2V = 0 .
H = H 2 +
The constraint eqn (6) can be obtained directly from Einstein’s field equations and represents
the first integral of eqn (5).
Since we are dealing with the entire universe, the system is necessarily conservative. We
use eqn (6) to ensure that energy is in fact conserved. It is amusing to notice that if we
isolate the matter sector, this subsystem looks dissipative. Energy is lost to the gravitational
field through the friction terms ∼ 3H Φ. We can see the effects of dissipation within the
4
(6)
larger context of the Hamiltonian system. For instance we can watch the matter tra jectories
shrink down onto an attractor as the volume in phase space is dissipated (cf §V).
In some cases it is profitable to recast the field equations in terms of conformal time τ ,
and the rescaled variables ψ = aΨ, φ = aΦ and U = a4V :
φ′′ −
a′′
a
φ + ∂φU = 0
ψ ′′ + kψ + ∂ψ U = 0
(7)
(8)
a′
a
(9)
(10)
(11)
a′
a
1
a
−
a′′ + ka +
(4U − ψ∂ψ U ) = 0 .
φ!2
a φ′ −
1
The total energy of the system can be expressed as
φ!2
a4H = (a′ )2 + ka2 − φ′ −
− (cid:16)(ψ ′ )2 + kψ 2(cid:17) − 2U = 0 .
A universe is said to inflate if the scale factor a accelerates in terms of cosmic time, i.e.
a > 0. The cosmic time acceleration is given by
a3
φ!2
2U − ψ∂ψ U − (ψ ′ )2 − kψ 2 − 2 φ′ −
1
.
Notice that conformally coupled fields tend not to contribute to a positive acceleration. In
other words, even with a potential ∝ Ψn , conformally coupled fields do not drive inflation
(unless n < 2). The generalization of the above equations to describe two or more scalar
fields of either type is direct.
The field equations are invariant under the combined rescaling
aa′′ − (a′ )2
a3
a =
=
a′
a
V
V →
t → xt,
a → xa,
x2 ,
where x is a constant. This freedom is removed when we set our length scale by choosing
dimensionfull values for quantities such as masses and coupling constants.
As we describe below, the asymptotic solutions are of two kinds. The universe eventually
approaches a smooth de Sitter phase or it ultimately collapses into a big crunch. In highly
simplified models the division between these two outcomes can be expressed as a simple
partition in the space of initial conditions. However, we shall see that even in simple models
with two interacting fields the division is no longer clean, as the boundaries separating the
different outcomes are no longer smooth curves but fractals.
(12)
A. From integrability to chaos
When the various scalar fields are massless and do not interact the equations of motion
can be integrated exactly and there is no chaos. Before launching into the chaotic dynamics,
5
we take a look in this subsection at the two asymptotic possibilities, the big crunch and
de Sitter expansion, which will be the basis of our outcomes approach in the following
sections. We also show the phase space portraits for the non-interacting, closed system.
In Ref. [1], a detailed analysis was given of the phase space portraits for a single, massive,
minimally coupled field in a universe with arbitrary curvature. According to their portraits,
the tra jectories drawn are untangled and therefore are not chaotic. As the authors noted,
there do exist a set of measure zero oscillatory solutions which do show chaotic behavior [5].
At the close of §III we return to discuss this special set of perpetually bouncing solutions.
For now we demonstrate the non-chaotic phase space for a universe full of noninteracting,
garden variety fields (both minimally and conformally coupled). Since it is impossible to tell
one scalar field from another in the absence of interactions, the general case of N minimally
coupled and M conformally coupled scalar fields reduces to a universe with one field of
each type. For a universe with a scalar field of each type and cosmological constant Λ the
equations simplify to
ψ = A cos τ ,
(13)
′
=
′
a
(14)
(15)
φ
a !
B
a2 ,
= ±sA2 + 2a4Λ +
B 2
a2 − a2 .
The phase space is divided by a separatrix into two classes of tra jectories, those that termi-
nate at the big crunch and those that inflate. The form of the solutions can easily be found
in the neighborhood of these two geometrically distinct attractors. When the dynamics is
dominated by potential terms, such as a cosmological constant Λ, the universe undergoes
exponential expansion and matter fields are redshifted away:
√2Λ τ ! , φ ∼ a−2 , ψ ∼ cos τ .
√2Λ t ∼ − 1
a ∼ e
Conversely, when the dynamics is dominated by the kinetic energy of the matter fields or
spatial curvature, the universe collapses to the big crunch at time τc :
a ∼ τ − τc 1/2 , φ ∼ a ln a , ψ ∼ cos τ
for B 6= 0 and for B = 0:
a ∼ τ − τc , φ ∼ constant , ψ ∼ cos τ .
The separatrix that partitions these possibilities is defined by the tra jectory with
A2 + 2√A4 + 3B 2
6(A2 + √A4 + 3B 2)2 .
For these simple, integrable cosmologies the basins of attraction for the big crunch and de
Sitter attractors are separated by a smooth curve. This smooth curve is a portion of the
(16)
Λ =
(17)
(18)
(19)
6
separatrix. In Fig. 1 we display phase space portraits in the (a, a′ ) and (ψ , ψ ′ ) planes for a
universe with B = 0 and Λ = 1/8. The cross-hatched region is the basin of the big crunch
attractor and the solid line is the separatrix.
When interactions are included the separatrix breaks up and is replaced by a fractal curve.
The gaps in the broken separatrix have the structure of a Cantor set. The broken separatrix
no longer partitions phase space and tra jectories may diffuse through it. For example, a
universe that was destined to collapse in the integrable case might diffuse through the broken
barrier and inflate. The breaking of the separatrix is reflected in the fractal nature of the
basin boundaries for chaotic universes. The smooth basin boundaries shown in Fig. 1 should
be compared to the fractal boundaries seen in Figs. 4 and 7. The break up of the separatrix
is further described in §IV.
Even when interactions are included, the asymptotic behaviour of tra jectories on either
attractor is completely regular and non-chaotic. Examples of this fact are given in §V. In
the parlance of dynamical systems theory, the attractors are neither strange nor chaotic.
The chaotic behaviour is transient [8], and occurs when tra jectories approach the broken
separatrix. Physically this corresponds to an epoch in which the universe coasts with a
fairly constant radius but with wildly varying acceleration. During this epoch the kinetic
and potential energies in the system fight for supremacy and the universe teeters between
collapse and violent expansion. If the kinetic energy wins the day, the universe collapses
and the asymptotic solution can be found by neglecting all potential terms in the equations
of motion. Conversely, if the potential energy wins the day the asymptotic solution can be
found by neglecting kinetic energy terms.
The transient nature of the chaos is similar to that found in the Mixmaster universe [9,10],
where it has been shown that the underlying attractors are neither strange nor chaotic [11].
We remark that transient chaos appears to be the hallmark of relativistic systems.
III. COSMOLOGIES WITH A SINGLE SCALAR FIELD
Since we work with a closed FRW cosmology, there is only one parameter describing
the gravitational sector – namely, the scale factor. The necessary elements for chaos are
present if the scale factor interacts even with just one matter field. However, the dynamical
timescale for the onset of chaos is longer than the life of one universe. The chaotic dynamics
results as the two oscillators interact. Typically, at least a few oscillations are needed for
the effects to surface. We discuss such an example in this section. On the other hand, if
there are many interacting matter fields in the universe, then their chaotic evolution will
make an impact during the lifetime of one universe. The examples of the following section
reveal chaos on such short time scales.
We begin with an example that is chaotic, but only on a timescale longer than the life
of one universe. The model describes a single, conformally coupled scalar field in a closed
(k = 1) universe. We choose the potential to have both a mass term and a cosmological
constant Λ:
U =
1
2
m2a2ψ 2 + a4Λ .
7
(20)
This example has previously been considered by Calzetta and El Hasi [6,7]. The Hamiltonian
takes the form
(21)
a4H = (a′ )2 + a2 − ((ψ ′ )2 + ψ 2 + m2a2ψ 2 ) − 2a4Λ = 0 ,
which, aside from the wrong sign for the gravitational contributions, is the Hamiltonian for
two coupled harmonic oscillators.
For a metric with only one dynamical degree of freedom it is always possible to perform
a combined field redefinition and conformal transformation to a coordinate system in which
the dynamics appears to be nonsingular. By using conformal, rather than cosmic time to
describe the evolution of this system, the dynamical equations can be smoothly integrated
past the big bang and big crunch singularities at a = 0. This allows many cosmic cycles to
be considered if we continue the scale factor into negative values. When evolved through a
series of cosmic cycles the system is clearly chaotic [6,7], as we might expect for nonlinearly
coupled oscillators. It should be noted that the cosmic cycles are physically meaningless as
all memory of the previous cycle is erased at each big crunch singularity.
By introducing the fiction of cosmic cycles, the dynamics can be surveyed using the
standard tools of Poincar´e sections (return maps) and Lyapunov exponents. Lyapunov ex-
ponents measure the rate of separation of tra jectories in phase space. Only if tra jectories
separate exponentially fast do they have positive exponents. Systems with positive Lya-
punov exponents are said to exhibit sensitive dependence on initial conditions - one of the
two ingredients of chaos (the other being the mixing and folding of tra jectories). The in-
verse of the positive Lyapunov exponents is referred to as the Lyapunov timescale. This
timescale sets the dynamical timescale over which chaotic effects make themselves felt. In
general relativity, Lyapunov exponents must be used with extreme care, if at all, as they are
coordinate dependent. Indeed, a simple coordinate transformation can give a non-chaotic
system positive exponents and a chaotic system vanishing exponents.
Putting these reservations aside, we may compare the Lyapunov time to the time taken
to complete a cosmic cycle, and infer whether or not chaotic effects can make themselves
felt in the lifetime of a single universe. Typically, the Lyapunov timescale was found to
be in the range 10 → 1000 cosmic cycles. Even when the mass is taken to be very large,
the Lyapunov timescale is always found to be greater than half a cosmic cycle, or in other
words, the timescale for chaos to become important always exceeds the life of one universe.
This result is easily understood. The chaotic behaviour is due to resonances between the
two oscillating fields a and ψ . In order for the resonance to take effect, both fields typically
need to oscillate several times. However, a can only complete half an oscillation before the
big crunch, making it exceedingly difficult for a chaotic resonance to occur.
In Ref. [6] it was argued that chaos had been viewed within the span of one life cycle.
Their conclusion was based on what appeared to be a scatter between initial values of the
matter fields and the final values. The correlation between initial and final values of the
scalar field ψ was found to be 0.01. However, this low value for the correlation actually
stems from a coarse sampling of a high frequency function. By regenerating Fig. 6. of Ref.
[6] with a sampling rate that is ten times higher we see from Fig. 2. that the true correlation
coefficient is 1.00. This confirms that the system shows no meaningful chaotic effects in the
life of one universe.
8
,
(22)
V =
m2Φ2 +
Similar conclusions hold for universes inhabited by a single minimally coupled scalar
field. Again, the equations of motion can lead to chaotic behaviour as they are nonlinear
and have phase space dimension greater than 2. However, meaningful chaotic effects can
only occur if the universe itself oscillates. The dynamics of an inflationary model driven by
a minimally coupled scalar field with potential
1
λ
0(cid:17)2
4 (cid:16)Φ2 − Φ2
2
was studied by Belinskii et al. [1], and with λ = 0 by Hawking [12] and Page [5]. Typical
tra jectories were not chaotic. Rather, they see the universe smoothly evolve from the big
bang to the big crunch with various amounts of inflation [1]. However, the inflationary
potential allows some atypical tra jectories for which the universe undergoes a number of
nonsingular bounces [12]. Page [5] suggested that there exist an uncountably infinite but
discrete set of perpetually bouncing universes with vanishing Lebesgue measure but non-
vanishing fractal dimension. If Page’s suggestion is correct, it would prove that the dynamics
is chaotic as his “fractal set of perpetually bouncing universes” corresponds to what is now
known as a strange repeller [13]. In contrast to the fictional cosmic cycles used to describe a
conformally coupled scalar field, Page’s bouncing universes are true, nonsingular solutions.
However, these solutions have obvious drawbacks as plausible cosmologies. As remarked in
Ref. [1], the fine tuning required to arrive at these chaotic tra jectories rules them out as a
robust physical model displaying chaotic behaviour. Perhaps in a model of the early universe
that generically displays nonsingular bounces we can hope to see interesting chaotic effects
caused by an oscillating scale factor. In the absence of such a model we have to look to
additional matter fields to provide the nonlinear resonances needed to incite chaos.
IV. COSMOLOGIES WITH TWO CONFORMALLY COUPLED FIELDS
If additional fields occupy the universe, then the scale factor will not be the principle
source of chaos. Two scalar fields can oscillate many times in the lifetime of one uni-
verse, leading to truly chaotic behaviour. To demonstrate the chaos we show the fractal
basin boundaries for a universe which contains two conformally coupled fields which inter-
act through the potential
U =
1a2ψ 2
m2
1 +
1 ψ 2
2 + λ2ψ 2
2a2ψ 2
m2
2 + a4Λ .
1
1
2
2
The period of oscillation for each field is governed by its effective mass. We define the
reduced effective mass for each field as the derivative with respect to the field of the field
eqn (9). In other words, M 2
ψ has the form of ∂ 2W/∂ψ2 were W is anything which acts as a
potential in the equations of motion:
2 − 4Λa2)1/2 ,
2ψ 2
1 − m2
1ψ 2
Ma = (1 − m2
(24)
(23)
2 )1/2 ,
1a2 + 2λ2ψ 2
M1 = (1 + m2
1 )1/2 .
2a2 + 2λ2ψ 2
M2 = (1 + m2
9
(25)
(26)
Increasing m1 , m2 , Λ and λ slows the recollapse of a and speeds the oscillation of ψ1 and ψ2 ,
thus increasing the probability of chaotic resonances. However, if m1 or m2 greatly exceed
λ, the resonances will be washed out and no chaos will be seen. Conversely, if m1 or m2
are both zero, the oscillations tend to freeze when ψ1 and ψ2 hit small values, again making
chaotic resonances unlikely.
To gain some intuition we can find a simple analytic approximation which corresponds to
a familiar chaotic system. During the ma jority of the universe’s evolution, the scale factor
varies much more slowly than the scalar fields so that a′/a ≪ ψ ′i/ψi . The scalar fields behave
like coupled nonlinear oscillators, adiabatically pumped by the slowly varying scale factor.
To leading order we can ignore the adiabatic pumping all together and study the scalar
field dynamics in a fixed background (a′ ≈ 0). This approximation is particularly good
for describing universes that are vacillating between collapse and inflationary expansion.
Importantly, this is just the region where the chaotic transients occur that destroy the
smooth separatrix of the integrable model described in §II A. When a′ ≈ 0 the dynamics
simplifies to that of two coupled oscillators:
1 ψ1 + 2λ2ψ 2
ψ ′′1 + ω 2
2 ψ1 = 0 ,
2 ψ2 + 2λ2ψ 2
ψ ′′2 + ω 2
1 ψ2 = 0 ,
(27)
(28)
i a2 is the fixed frequency of the uncoupled (λ = 0) oscillators. The above
i = 1 + m2
where ω 2
system of equations describes a known chaotic system [14], and the transition to chaos as
λ is increased can be studied using the Chirikov resonance overlap condition [15]. Having
established that the fast variables ψ1 and ψ2 behave chaotically, we can then consider how
they backreact on the slow variable a. When looked at on timescales long compared to the
periods of the scalar fields, the evolution of the scale factor is similar to Brownian motion,
and can be described in terms of chaotic diffusion equations [8]. It is this buffeting of the
scale factor by the matter fields that breaks the separatrix in the (a, a′ ) plane and causes
the universe to evolve in a chaotic manner.
Returning to the full, unapproximated equations we numerically investigate the phase
space of initial conditions. For a given set of initial conditions we can identify three main
outcomes. The first possibility sees the universe expand and collapse without any inflationary
burst. The second possibility sees the universe undergo one or many short bursts of inflation,
but failing to become a macroscopic universe. The third possibility sees the universe sustain
a prolonged and violent period of inflation resulting in the formation of a macroscopic
universe. The first and second possibilities (coloured Black and Grey respectively) combine
to form the big crunch basin of attraction. This artificial division of the big crunch basin is
mostly for visual effect. There is a fourth possible outcome that should be mentioned. There
are a set of tra jectories with zero Lebesgue measure that oscillate eternally, never entirely
collapsing or reaching the de Sitter attractor. These tra jectories form the border between
the big crunch and de Sitter basins of attraction. We will see that these tra jectories belong
to a fractal set of perpetually bouncing universes.
In the parlance of dynamical systems
theory, this set forms the stable manifold of a strange repeller [8].
The three possibilities are displayed graphically in Fig. 3 for the choice of parameters
(Λ = 0.0001, m1 = 0, m2 = 0.05, λ = 1) and initial conditions {a(0) = 2, ψ1 = 0.4, ψ2 =
10
6, ψ ′2 = 20}. The initial values of ψ ′1 are {−23.31, −23.32, −23.33}, and a′ (0) is fixed by
the Hamiltonian constraint.
The fact that minute changes in the initial conditions can lead to such dramatic changes
in the outcome suggest that the fate of our model universe is indeed chaotic. This suspicion
can be confirmed by studying the boundary between the basins of attraction of the three
outcomes. Since the basins are embedded in a six dimensional phase space, we are forced
to consider lower dimensional slices through the boundary.
In Fig. 4 we display a two-
dimensional slice in the ψ1 – ψ ′1 plane for universes with parameters and conditions identical
to those used in Fig. 3. The three basins of attraction (Black, Grey, White) are dramatically
intermixed strange basins, as at least a portion of the boundaries are fractal. The boundaries
near the origin are regular and smooth while the outer boundaries appear fragmented. A
detail of the outer region is shown in Fig. 5, visually confirming the fractal nature of the
boundary. Repeated magnification reveals similar striated pictures on all scales.
Rather than rely on these qualitative features, we may quantify the fractal nature of the
boundary in terms of the fractal dimension. There are many definitions of fractal dimension
that we may choose from, but the one best suited to our situation is the box counting
dimension. On a two-dimensional slice through phase space we cover the fractal with a grid
of squares with side length ε. We then count the number, N (ε), of squares needed to cover
the fractal, i.e. the number of squares containing more than one colour. The box dimension
dB is defined by
.
(29)
ln N
dB = − lim
ln ε
ε→0
For self-similar structures the formal limit ε → 0 need not be taken, and in all practical
situations we are only interested in the existence of such scaling laws over a large, but
not necessarily infinite, range of scales. Since the fractal dimension is not invariant under
homeomorphisms, it is not a true topological invariant. However, it is invariant under
diffeomorphisms, so it does provide a topological measure in general relativity. The existence
of fractal structures in phase space provides a coordinate independent signal of chaos in
relativity.
The importance of the fractal dimension of the basin boundaries can be described in
terms of final state sensitivity [16]. Consider an initial configuration near the basin boundary,
where the uncertainty in the initial conditions describes an N-dimensional ball of radius δ in
the N-dimensional phase space. The final state sensitivity fδ is the fraction of phase space
volume which has an uncertain outcome due to the uncertainty in the initial conditions, and
is given by
fδ = δα , α = N − dB .
For a non-chaotic system α = 1 and the final state sensitivity is directly proportional to
the initial uncertainty. For chaotic systems however, 0 < α < 1, and the uncertainty in the
outcome is greater than the uncertainty in the initial conditions. For example, if α = 0.47, a
50% reduction in the initial uncertainty only reduces the final state uncertainty by 28%. In
this way, the dimension of the basin boundary is a direct measure of “sensitive dependence
on initial conditions”.
(30)
11
In Fig. 6 we display the plot used to determine the fractal dimension of Fig. 5. Because
the three boundaries are densely interwoven in this case, we chose only to calculate the
dimension of the boundary between the big crunch and de Sitter attractors, i.e. counting
Grey and Black as one basin. Using an 840 × 840 grid we found the dimension to be
1.58 ± 0.02. The grid size of 840 = 23 × 3 × 5 × 7 was chosen as it has the most factors
of any number below 1000. The curvature of the data points at small and large ε is to be
expected. For large ε the covering is very inefficient, while for small ε the squares saturate
the resolution used to generate the fractal. These effects cause dB to tilt toward 2 for
large ǫ and toward 1 for small ǫ. Despite these limitations, accurate fractal dimensions can
be obtained very quickly and easily. For different choices of parameters we found fractal
dimensions ranging from 1 to 1.96, essentially filling the allowed range dB = [1, 2].
The boundary was found to be fractal on all possible two-dimensional slices. For example,
in Fig. 7, the boundary is shown in the a – a′ plane for a slice which intersects Fig. 3 along
the line ψ1 = 1.0. The fractal dimension of this slice was found to be dB = 1.37 ± 0.02.
While the previous chaotic pictures were typical of those found, the dynamics of the
system is not always chaotic. For small values of λ, m1 , m2 (at fixed scaling x) the dynamics
is near integrable and the basins are not strange, but regular. In Fig. 8 we increment λ
while keeping all other parameters and initial conditions fixed. The mixing of the basins
is reminiscent of the blending of viscous fluids. The dimension of the basin boundary for
λ = 0.5 was found to be dB = 0.99 ± 0.02, which is consistent with a dimension of 1.
So, within errors, this boundary is smooth and non-chaotic. To compare, the dimensions
of the boundaries for λ = 2.0 was dB = 1.16 ± 0.05 (Grey-White) and dB = 1.26 ± 0.05
(Grey-Black).
An important property of dynamical systems with strange attractor basins is that the
chaotic dynamics is not restricted to phase space tra jectories near the fractal boundaries.
One way to see this might be to use the fiction of cosmic cycles to follow the evolution of
tra jectories starting in the big crunch basin. The Lyapunov exponents and Poincar´e sections
for these tra jectories would reveal chaotic behaviour across the basin. However, we are not
really interested in effects which take longer than one universe’s lifetime to make themselves
felt. Instead we plot in Fig. 9 the correlation between the initial value of ψ1i and the value
at the point of maximum expansion, ψ1m . The graphs are for a ψ ′1 = 0 slice through the
big crunch basin of Fig. 8 with λ = 2.0. The big crunch basin stretches from ψ1 = 0 to
ψ1 ∼ 2.9 (and similarly for negative ψ1 ). A general increase in frequency with increasing
ψ1i requires that we use several plots, each covering half the region of the last, to cover the
basin. Unlike the regular plot seen in Fig. 2, the relationship between initial and final values
of ψ1 is highly erratic, with apparently random changes in frequency and amplitude.
V. COSMOLOGIES WITH MINIMALLY AND CONFORMALLY COUPLED
FIELDS
The chaotic behaviour seen in the previous system is not restricted to conformally coupled
fields. Similar behaviour is found for minimally coupled fields with the same choice of
potential. The main difference in this case comes from the scalar fields themselves being
a source of inflation, in addition to the cosmological constant. The acceleration in this
12
1
2
1
2
(31)
1Φ2
m2
1 +
example is given by
a = 2a (cid:18)Λ +
2(cid:19) ,
2 − Φ2
1 − Φ2
1Φ2
2 + λ2Φ2
2Φ2
m2
where we have reverted to the unscaled field variables. These models are able to successfully
inflate even when there is no cosmological constant, in a manner similar to Linde’s “chaotic
inflation” [17]. However, we did not see any strange basins when Λ = 0 as successful
inflation generally required the fields to become stuck high up in their potentials after just
a few oscillations. Otherwise, their ability to climb high enough was lost due to friction and
redshifting of kinetic energy. It may be that chaotic behaviour does occur when Λ = 0, but
it is difficult to search for as the inflationary bursts must be followed for ∼ 60 e-folds in
comparison to the ∼ 5 − 10 e-folds required to ensure we have reached the de Sitter attractor
when Λ 6= 0.
In Fig. 10 we display the basins of attraction in the Φ1 - Φ1 plane for universes with
(m1 = 0, m2 = 0.04, Λ = 0.00005, λ = 1.0) and fixed initial conditions {a(0) = 10.0, Φ2 =
Φ2 = 0.16}. A detail of the outer boundary is shown in Fig. 11, where the dimension was
0.4,
found to be 1.54 ± 0.02. Because the scalar fields themselves contribute to the inflationary
bursts there is typically far more of the Grey basin than we saw for conformally coupled
fields.
We can compare an analysis of the chaotic tra jectories with the non-chaotic tra jectories
of §II A. While the universe is expanding (H > 0), we see from eqn (3) that the scalar field
dynamics is effectively that of a damped harmonic oscillator. Conversely, as the universe
contracts the dynamics is that of a pumped harmonic oscillator. This behaviour is apparent
in Fig. 12, where we have displayed typical tra jectories leading to the de Sitter and big
crunch attractors. For the de Sitter attractor the scalar fields spiral into a fixed point as
cosmic baldness asserts itself, while for the big crunch attractor the scalar fields first spiral in
and then spiral out again as the universe collapses. The de Sitter attractor has an interesting
structure when two fields are present as one scalar field gets locked at a constant value. The
attractor is of the form
a = ac exp(√2Λt) ,
3√2Λ
t! cos ω t ,
Φ2 = Φ2c exp −
2
2c exp(−3√2Λt) t
Φ1 = Φ1c "1 + λ2Φ2
3√2Λ
sub ject to the restriction
2ω cos2 ω t + 3√2Λ cos ω t sin ω t
2ω (9Λ + 2ω 2)
!# ,
(32)
(33)
(34)
+
2 + 2λ2Φ2
ω 2 = m2
1c −
For reference, the tra jectory in Fig. 12a has Φ1c = 0.00573, Φ2c = 105.9 and ω = 0.03796.
The big crunch attractor is unchanged from the one-field case, and takes the form
(35)
Λ .
9
2
13
a = ac (tc − t)1/3 ,
Φ1 ∼ Φ1c ln(tc − t) ,
(36)
(37)
(38)
Φ2 ∼ Φ2c ln(tc − t) ,
1c + Φ2
with Φ2
2c = 1/9. The tra jectory shown in Fig. 12b has tc = 200.004, ac = 13.68,
Φ1c = 0.294 and Φ2c = 0.167. Since we are able to write down analytic solutions for
tra jectories on the attractors, it is clear that the de Sitter and big crunch attractors are
non-chaotic. The chaotic behaviour seen during the evolution of the universe is restricted
to the 5 → 10 transient orbits seen in Fig. 12. The fractal nature of the attractor basin
boundaries is due entirely to these brief chaotic transients.
We close with a word on a mixed cosmology which contains one minimally coupled and
one conformally coupled scalar field. The conformally coupled scalar field Ψ is taken to be
massless. When Ψ’s coupling to the minimally coupled scalar field Φ is small it behaves like
radiation. The interaction potential is taken to be
V =
1
2
m2Φ2 + λ2Ψ2Φ2 + Λ .
(39)
As before, the minimally coupled scalar field is able to contribute a negative pressure to that
of the inflaton, thereby increasing the likelihood of inflationary bursts. Again, this increases
the proportion of Grey over what we saw for two conformally coupled fields.
By viewing the basins in the a - a plane we see some rather striking ink-blot and crystal
boundaries. An example of this is shown in Fig. 13 for the choice of parameters (m =
Ψ = 0.08}. A detail
0.05, Λ = 0.0001, λ = 2), and fixed initial conditions {Φ = 0.2, Ψ = 0.1,
of the Grey-White crystal boundary is shown in Fig. 14. The high degree of self-similarity
of this fractal allowed a particularly accurate determination of the fractal dimension using a
standard 840 × 840 grid. The dimension was found to be dB = 1.484 ± 0.005. The visually
less fractal Grey-Black ink-blot boundary was found to have a smaller fractal dimension of
dB = 1.11 ± 0.05, with the larger error due to a lower degree of self-similarity.
In addition to studying different combinations of minimally and conformally coupled
scalar fields, we also considered a variety of polynomial potentials V (Φ, Ψ). The qualitative
results were the same for all cases, showing that chaotic evolution was a generic feature of
all multi-field models.
VI. DISCUSSION
The early universe is likely to be home to many interacting fields. We have show that if
these interactions are sufficiently strong, the evolution of the universe will be chaotic. The
fractal basin boundaries reveal the chaos in a coordinate independent manner. Additionally,
the method does not require one universe to pass through many cycles. We now have to
ask how prevalent chaotic behaviour will be in particular models of inflation, and what
implications it might have for processes such as reheating or galaxy formation.
14
One model of inflation where chaotic dynamics is bound to be important is hybrid infla-
tion. Hybrid models employ several interacting scalar fields, and arise naturally in various
supersymmetric theories where the breaking of large gauge groups employs many Higgs par-
ticles [18]. If in the early stages the universe inflates in jolts, an exciting possibility exists for
the spectrum of primordial density fluctuations. Chaotic resonances could lead to a fractal
power distribution, perhaps helping to explain the hierarchical clustering seen in the current
universe. Moreover, chaotic evolution of the scale factor would leave a unique imprint on
the gravitational waves produced during inflation [19]. Any chaotic behaviour would have
to occur within the last ∼ 60 e-folds of inflation to be observable today. This would require
some artificial fine tuning in a single field model but may be more natural in a hybrid model.
Even in inflationary models where chaotic evolution is unimportant at early stages, chaos
is sure to play an important role at the end stages. To illustrate, consider again a hybrid
model. The fields to which the inflaton couples dictate the occurrence of the true vacuum and
so control the nature of the exit from inflation. The setting is prime for chaotic interactions
which would certainly impact on the exit style. More generically, at the end of any inflation
model, the universe reheats as the inflaton oscillates about the minimum of its potential.
Particles are thereby produced through the inflaton’s coupling to other matter fields. If the
matter fields are dynamical and chaos reigns, then the process of entropy production would
deserve rethinking. The importance of parametric resonances, which are closely related to
chaotic behaviour, has already been stressed in this context [20,21].
It seems appropriate to consider how chaotic dynamics might impact on “chaotic” infla-
tion [17]. In chaotic inflation different patches of the universe are taken to have different
values of the inflaton and matter fields. In some patch, it is argued, the inflaton is suffi-
ciently high up in the potential and the matter fields are sufficiently small so as to permit
a long-lived inflationary epoch [2]. For initial field values deep within a basin, away from
the fractal borders, the usual arguments hold and the chaotic inflation paradigm is largely
unaffected. However, if in a given patch the field values are near a fractal basin boundary,
it can become difficult to find a patch of any size across which the conditions are regular
enough to allow this thinking. Even the slightest variation in the initial conditions across
the patch will lead to an entirely different outcome. Due to the self-similar nature of the
fractal, no matter how small you try to make the patch, there will still be slight variations
in the conditions and hence the outcome. For cosmological conditions in the vicinity of the
fractal basin boundary then, the simple FRW thinking must be abandoned.
For similar reasons, caution would be needed for slow-roll initial conditions as well.
In fact the slow-roll scenario will likely be more fragile as the inflaton is more easily kicked
around. For chaotic initial conditions by contrast the field high up in its potential is resiliant
against the influence of kicks and bumps.
Aside from direct observational effects there are also some important theoretical im-
plications raised by chaotic evolution. For example, chaotic systems are characterised by
an entropy, the Kolmogorov-Sinai entropy, which is related to the spectrum of Lyapunov
exponents. This introduces a chaotic arrow of time in addition to the cosmological and
thermodynamic arrows of time. As well, it raises the question of a possible connection be-
tween the Kolmogorov-Sinai entropy and the thermodynamic entropy released at the end of
inflation. Another issue raised by chaotic dynamics concerns the recovery of a semiclassical
limit in quantum cosmology due to the breakdown of the WKB approximation in chaotic
15
systems [22].
We have suggested a few implications for chaotic dynamics in largely unexplored terrain.
Chaos theory grew out of Poincar´e’s study of the solar system, and with its development
came insights into the intricate structures of our neighborhood such as the asteroid belt and
saturn’s rings. It seems fitting for chaos to have an impact not only on the evolution of the
solar system but also on the birth of the universe. We are left to ask if chaotic fingerprints
have been left on the large-scale landscape as they were on the landscape of our own solar
system.
ACKNOWLEDGEMENTS
We have enjoyed discussions with Matt Holman and Jihad Touma. We thank John
Barrow, Norm Frankel, Andrei Linde and Don Page for their helpful comments. We are
indebted to Carl Dettmann for making his original programs available to us, and to Jacques
Legare for his help in making modifications.
16
REFERENCES
[1] V. A. Belinskii, L. P. Grishchuk, I. M. Khalatnikov and Ya. B. Zel’dovich, Phys. Lett.
B155, 232 (1985); V. A. Belinskii, L. P. Grishchuk, I. M. Khalatnikov and Ya. B.
Zel’dovich, Sov. Phys. JETP 62, 195 (1985).
[2] D. S. Goldwirth and T. Piran, Phy. Rep. 214, No. 4, 223 (1992); D. S. Goldwirth and
T. Piran, Phys. Rev. D40, 3263 (1989); A. D. Linde, Phys. Lett. B162, 281 (1985); R.
H. Brandenberger and J. H. Kung, Phys. Rev. D42, 1008 (1990).
[3] A. D. Linde, Phys. Lett. B259, 38 (1991); A. D. Linde, Phys. Rev. D49, 748 (1994).
[4] C. P. Dettmann, N. E. Frankel and N. J. Cornish, Phys. Rev. D50, R618 (1994); Frac-
tals, 3, 161 (1995).
[5] D. N. Page, Class. Quantum Grav. 1, 417 (1984).
[6] E. Calzetta and C. El Hasi, Class. Quant. Grav. 10, 1825 (1993).
[7] E. Calzetta and C. El Hasi, Phys. Rev. D51, 2713 (1995).
[8] E. Ott, Chaos in dynamical systems, (Cambridge University Press, Cambridge, 1993).
[9] C. Misner, Astrophys. J. 151, 431 (1968).
[10] J. D. Barrow, Phys. Rep. 85, 1 (1982).
[11] P. Ma and J. Wainwright, Deterministic Chaos in General Relativity, eds. A. Burd, A.
Coley and D. Hobill (Plenum, New York, 1994) p.449.
[12] S. W. Hawking, Lectures at the NATO Summer School on Relativity, Groups, and Topol-
ogy in Les Houches, France. (1983).
[13] L. P. Kadanoff and C. Tang, Natl. Acad. Sci. USA 81, 1276 (1984); H. Kantz and P.
Grassberger, Physica D17, 75 (1985).
[14] B. V. Chirkov and D. L. Shepelyansky, JETP Lett. 34, 163 (1981); G. K. Savvidy, Phys.
Lett. B130, 303 (1983); G. Contopoulos, H. E. Kandrup and D. Kaufmann, Physica
D64, 310 (1993).
[15] B. V. Chirikov, Phys. Rep. 52, 265 (1979).
[16] C. Grebogi, S. McDonald, E. Ott, and J. Yorke, Phys. Lett. A99, 415 (1983).
[17] A. D. Linde, Phys. Lett. B129, 177 (1983).
[18] A. Guth, L. Randall and Soljacic in preparation (1995).
[19] L. P. Grishchuk and M. Solokhin, Phys. Rev. D43, 2566 (1991).
[20] L. Kofman, A. D. Linde and A. A. Starobinsky, Phys. Rev. Lett. 73, 3195 (1994).
[21] Y. Shtanov, J. Traschen and R. H. Brandenberger, Phys. Rev. D51, 5438 (1995).
[22] M. V. Berry, Chaotic Behaviour of Deterministic Systems, Les Houches, Session 36,
eds. G. Ioos, R. Helleman and R. Stora, (North Holland, Amsterdam, 1983).
17
FIGURES
a′
>
<
ψ ′
a
ψ
Phase space tra jectories in the (a, a′ ) and (ψ , ψ ′ ) planes for universes with
FIG. 1.
non-interacting scalar fields. The solid line is the separatrix and the cross hatched regions mark
the big crunch basin of attraction.
18
ψf
ψi
FIG. 2. The correlation between initial and final values of the scalar field (a) is compared to
the apparently chaotic behaviour seen in (b) where the sampling rate is ten times lower.
FIG. 3. The three possible outcomes for the universe. In each case the solid line is the scale
factor a and the dashed line is the scaled acceleration aa3 . The initial values for ψ ′1 are −23.31,
−23.32 and −23.33 respectively.
19
ψ ′1
FIG. 4. The basins of attraction for universes similar to those shown in Fig. 3.
ψ1
ψ ′1
ψ1
FIG. 5. A detail of Fig. 4.
20
FIG. 6. Finding the fractal dimension for Fig. 5. The solid line is a least-squares fit to the
box counting data. The dimension was found to be dB = 1.58 ± 0.02.
a′
FIG. 7. A slice in the a - a′ plane which intersects the ψ1 - ψ ′1 plane of Fig. 4. along the line
ψ1 = 1.0.
a
21
ψ ′1
ψ1
FIG. 8. The road to chaos: As λ is incremented from 0.5 to 2.0, the nonlinear distortion of the
attractor basin boundaries mounts. Once λ exceeds 1.0 the mixing is so strong that the boundaries
become fractured, and eventually, fractal. The graphs were generated for the choice of parameters
and initial conditions Λ = 0.00002, a0 = 10, ψ2 = 5.0, ψ ′2 = 10.0, m1 = 0.2, m2 = 0.1.
FIG. 9. The correlation between ψ1i (vertical axis) and ψ1m (horizontal axis) on a ψ ′1 = 0 slice
through the big crunch basin of attraction of Fig. 8. (λ = 2.0).
22
Φ1
Φ1
FIG. 10. Basins of attraction in the Φ1 - Φ1 plane for universes containing two minimally
coupled scalar fields.
Φ1
Φ1
FIG. 11. A detail of Fig. 10. where the dimension is dB = 1.54 ± 0.02.
23
Φ2
FIG. 12. (a) A tra jectory spiraling into the de Sitter attractor. (b) A nearby tra jectory which
flows out to the big crunch attractor.
Φ2
a
a
FIG. 13. The basins of attraction in the a - a plane for universes containing a minimally
coupled scalar field and a massless, conformally coupled scalar field.
24
a
a
FIG. 14. A detail of Fig. 13. where the dimension is dB = 1.484 ± 0.005.
25
|
astro-ph/0701139 | 1 | 0701 | 2007-01-05T13:47:23 | Temperature fluctuations in H II regions: t2 for the two-phase model | [
"astro-ph"
] | Aims: We investigate temperature fluctuations in H II regions in terms of a two-phase model, which assumes that the nebular gas consists of a hot and a cold phase.
Methods: We derive general formulae for T([O III), the [O III] forbidden line temperature, and T(H I), the hydrogen Balmer jump temperature, in terms of the temperatures of the hot and cold phases, T_h and T_c.
Results: For large temperature differences, the values of t2 required to account for the observed difference between T([O III]) and T(H I) are much lower than those deduced using the classical formulae that assume random and small amplitude temperature fluctuations. One should therefore be cautious when using a two-phase model to account for empirically derived $t^2$ values. We present a correction of a recent work by Giammanco & Beckman, who use a two-phase model to estimate the ionization rate of H II regions by cosmicrays. We show that a very small amount of cold gas is sufficient to account for t2 values typically inferred for H II regions. | astro-ph | astro-ph | Astronomy&Astrophysicsmanuscript no. 6564
October 22, 2018
c(cid:13) ESO 2018
Temperature fluctuations in H ii regions: t2 for the two-phase model
Y. Zhang1,2, B. Ercolano3, and X.-W. Liu1
1 Department of Astronomy, Peking University, Beijing 100871, China
e-mail: [email protected]
2 Department of Physics, University of Hong Kong, Pokfulam Road, Hong Kong, China
3 Harvard-Smithsonian Centre for Astrophysics, 60 Garden Street, Cambridge, MA 02138, USA
7
0
0
2
n
a
J
5
1
v
9
3
1
1
0
7
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
Received; accepted
ABSTRACT
Aims. We investigate temperature fluctuations in H ii regions in terms of a two-phase model, which assumes that the nebular gas
consists of a hot and a cold phase.
Methods. We derive general formulae for T ([O iii]), the [O iii] forbidden line temperature, and T (H i), the hydrogen Balmer jump
temperature, in terms of the temperatures of the hot and cold phases, Th and Tc.
Results. For large temperature differences, the values of t2 required to account for the observed difference between T ([O iii]) and
T (H i) are much lower than those deduced using the classical formulae that assume random and small amplitude temperature fluctu-
ations. One should therefore be cautious when using a two-phase model to account for empirically derived t2 values. We present a
correction of a recent work by Giammanco & Beckman, who use a two-phase model to estimate the ionization rate of H ii regions by
cosmic rays. We show that a very small amount of cold gas is sufficient to account for t2 values typically inferred for H ii regions.
Key words. ISM: general -- ISM: H ii regions
1. Introduction
Temperature fluctuations in H ii regions are a much-discussed problem. Peimbert (1967) investigated for the first time the effects
of such fluctuations on temperatures empirically derived from spectroscopic observations and found that they may lead to higher
electron temperatures being derived from the collisionally excited [O iii] nebular-to-auroral forbidden-line ratio, T ([O iii]), than
those derived from the Balmer jump of the H i recombination spectrum, T (H i). If significant temperature fluctuations exist in H ii
regions and yet are ignored in the analysis, they may lead to underestimating ionic abundances calculated from collisionally excited
lines (CELs) (e.g. Esteban et al. 2002). The parameter t2 (Peimbert 1967) was introduced to quantitatively characterize temperature
fluctuations. Since then the parameter has been extensively used in nebular studies. Under the conditions that [T (r) − T0]2/T 2
0 ≪ 1,
where T (r) is the local electron temperature and T0 the average value weighted by the square of density. The value of t2 can be
determined either by comparing T ([O iii]) and T (H i) or by comparing ionic abundances derived from CELs and from recombination
lines (RLs) (see Peimbert et al. 2004, for further details).
Two-phase models, which approximate a nebula by two components of different physical conditions, represent an over-
simplified, yet frequently used method of studying nebular physics (e.g. Viegas & Clegg 1994; Zhang et al. 2005). Using a two-phase
model, Stasi´nska (2002 ) points out that the classical picture of temperature fluctuations may be misleading under certain conditions,
and three parameters are needed to characterize temperature inhomogeneities. One of the open questions in the study of H ii regions
is that values of t2 derived from observations are consistently higher than those predicted by photoionization models (Stasi´nska
2000). Recently, Giammanco & Beckman (2005; GB05 thereafter) constructed a two-temperature-phase model capable of explain-
ing t2 values deduced for a number of H ii regions by means of incorporating a component of cool ionic gas ionized by cosmic rays.
However, t2 values deduced from observations cannot be applied directly to two-phase models. This is because empirical values
of t2 deduced from observations were calculated from formulae derived by assuming random and small-amplitude temperature
fluctuations, assumptions that are apparently broken for a two-phase model.
The purpose of the current work is to quantitatively study the relationship between the t2 values predicted by two-phase models
and those measured by observations. We show that, in the case of a very cold ionic gas component embedded in a 'normal' H ii
region, t2 deduced from observations using the empirical method may have significantly overestimated the real values.
2. Analysis
According to Peimbert (1967), for a given ionic species of number density Ni, the thermal structure of an H ii region can be
characterized by an average temperature T0 and a mean square temperature fluctuation parameter t2, defined as
T0 = R TeNeNidV
R NeNidV
(1)
Send offprint requests to: Y. Zhang
2
and
Y. Zhang et al.: Temperature fluctuations in H ii regions: t2 for the two-phase model
t2 = R (Te − T0)2NeNidV
,
T 2
0 R NeNidV
respectively. Assuming t2 ≪ 1, T0 and t2 can be determined from measured T ([O iii]) and T (H i) using relations,
T ([O iii]) = T0"1 +
2 9.13 × 104
− 3! t2#
1
T0
(2)
(3)
and
T (H i) = T0(1 − 1.67t2)
(Peimbert 1967; Garnett 1992). Esteban et al. (2002) report t2 values for a sample of H ii regions. Their values are reproduced in
the first row of Table 1. Although they were obtained by comparing ionic abundances derived from CELs and from RLs, we assume
that they are the same as would be deduced from from T ([O iii]) and T (H i) using Eqs. (3) and (4). The assumption is supported by
studies of Torres-Peimbert & Peimbert (2003), who find general agreement between t2 values inferred from the differences between
T ([O iii]) and T (H i) and those inferred from the apparent discrepancies between CEL and ORL abundances. Esteban et al. (2002)
did not provide values of T0. For the purpose of comparison, we follow GB05 and assume that T0 = T ([O iii]), as given in the second
row of Table 1. However, as pointed out previously by Stasi´nska (2002), the empirical method of estimating t2 using Eqs. (3) and
(4) may be invalid under certain conditions. In the following, we show that in the two-phase model, values of t2 that are required
to account for the measured differences between T ([O iii]) and T (H i) (or differences between ORL and CEL abundances) may be
much lower than that derived from the empirical method.
(4)
In the framework of the two-phase model, which assumes that the electron temperature structure of an H ii region consists of a
hot and a cold phase, we follow GB05 and assume equal densities of the two phases and an ionization fraction of unity for the hot
gas. Electron temperatures are designated as Th and Tc for the hot and cold phases, respectively. The intensity of an [O iii] forbidden
line transition of wavelength λ is given by
I([O iii])λ ∼ Z N(O2+)NeT −1/2
e
exp(−∆E/kTe)dV ,
where ∆E is the excitation energy of the upper level. It follows that
I([O iii])4959,5007
I([O iii])4363
≡ C × exp( 33000
T ([O iii]))
e VhT −1/2
e VhT −1/2
N(O2+)hNh
N(O2+)hNh
= C ×
h
h
exp(−29200/Th) + N(O2+)cNc
exp(−62200/Th) + N(O2+)cNc
c
e VcT −1/2
e VcT −1/2
c
(5)
(6)
exp(−29200/Tc)
exp(−62200/Tc)
,
where C is a constant depending only on atomic data, and super- or subscript 'h' and 'c' refer to quantities of the hot and cold
phases, respectively. In the cold phase, the ionization fraction of oxygen is expected to be lower than that of hydrogen. As a
reasonable approximation, we assume that N(O2+)c/N(O)c = 0.1N(H+)c/N(H)c. Our analysis is insensitive to this assumption since
the cold phase essentially contributes no [O iii] forbidden line fluxes given its very low temperature and the fact that emissivities of
forbidden lines decline exponentially with decreasing temperature (c.f Eq. 5). We thus obtain
T ([O iii]) = 33000 ln−1
T −1/2
h
T −1/2
h
exp(−29200/Th) + 0.1θx2T −1/2
exp(−62200/Th) + 0.1θx2T −1/2
c
c
exp(−29200/Tc)
exp(−62200/Tc)
,
(7)
where, using the notation of GB05, θ is the mass ratio of the cold to hot gas, and x is the ionization fraction of hydrogen in the cold
gas. It can be easily seen that for Th ≫ Tc, T ([O iii]) ∼ Th.
Similarly, from the flux of the Balmer jump,
I(Bal, 3646) ∼ Z N(H+)NeT −3/2
e
dV ,
and the flux of Hβ,
I(Hβ) ∼ Z N(H+)NeT −5/6
e
dV ,
we have
T −3/2
h
T −5/6
h
T (H i) =
+ θx2T −3/2
+ θx2T −5/6
c
c
,
−3/2
(8)
(9)
(10)
Y. Zhang et al.: Temperature fluctuations in H ii regions: t2 for the two-phase model
3
Fig. 1. t2 versus [T ([O iii]) − T (H i)], deduced from the empirical method [Eqs. (3) and (4), dotted lines] and for two-phase model
[Eqs. (7), (10), (12), and (13), solid lines]. Left panels: θx2 = 0.01; Right panels: θx2 = 0.1; Upper panels: Th = 10000 K; Lower
panels: Th = 15000 K.
or
θx2 = Th
Tc!−3/2 T (H i)−2/3T 2/3
h − 1
1 − T (H i)−2/3T 2/3
.
c
Equation (10) shows that T (H i) weights towards Tc. Following GB05, for a two-phase model, we obtain
T0 =
Th + θx2Tc
1 + θx2
,
t2 = θx2 Th − Tc
Th + θx2Tc!2
,
θx2 =
(Th − T0)2
(T0 − Tc)2
=
T 2
0 t2
(T0 − Tc)2 ,
and
Th = T0 +
T 2
0 t2
T0 − Tc
.
(11)
(12)
(13)
(14)
(15)
Therefore, for the two-phase scenario, T0 and t2 should be determined from Eqs. (7), (10), (12), and (13) instead of from Eqs. (3)
and (4). The latter are only valid for random and small amplitude fluctuations.
In Fig. 1 for given values of θx2 and Th, we compare t2 as a function of T ([O iii]) − T (H i) derived from the empirical method
using Eqs. (3) and (4), and that derived in the scenario of two-phase model using Eqs. (7), (10), and (13). The plots show that,
depending on θx2 above a critical value of T ([O iii]) − T (H i), the empirical method significantly overestimates t2, particularly for
the case of small θx2. The amount of deviation is insensitive to the value adopted for Th. In addition, we find that as the temperature
difference between the two phases is larger than a critical value (typically ∼ 6000 K), empirical t2 deduced from observations can
no longer be used at their face values to constrain two-phase models, a point overlooked by GB05 as discussed in the following
section.
4
Y. Zhang et al.: Temperature fluctuations in H ii regions: t2 for the two-phase model
Table 1. Estimated values of t2, T0, and θx2 for Tc = 100, 1000, and 4000 K for a sample of H ii regions.
T0,E and t2
E are taken from Esteban et al. (2002). The numbers in parentheses are values derived by GB05
and included here for comparison.
E,T0,E)
E,T0,E)
Object
t2
E
T0,E
T ([O iii])(t2
T (H i)(t2
(t2)100
(t2)1000
(t2)4000
(T0)100
(T0)1000
(T0)4000
(θx2)100
(θx2)1000
(θx2)4000
NGC 604
NGC 5461
NGC 5471
NGC 2363
0.027
8150
9052
7783
0.00013
0.00413
0.02511
9051
9011
8518
0.041
8600
9943
8011
0.00016
0.00053
0.03503
9941
9884
9370
0.074
14100
15913
12358
0.00009
0.00312
0.02401
15911
15860
15413
0.128
15700
18528
12344
0.00013
0.00429
0.03540
18526
18445
17719
0.00013(0.028)
0.00522(0.035)
0.08690(0.10)
0.00017(0.042)
0.00659(0.052)
0.10664(0.14)
0.00010(0.075)
0.00356(0.086)
0.04378(0.14)
0.00013(0.130)
0.00480(0.146)
0.05905(0.23)
3. New estimates of θx2 for GB05 model
GB05 showed that the ionization of cold neutral gas by cosmic rays may significantly contribute to temperature fluctuations. They
used a two-phase model to explain t2 values obtained by Esteban et al. (2002) for a number of H ii regions. However, the high
temperature difference between the two phases in GB05 model (see their Table 1) suggests that t2 values obtained by Esteban et al.
cannot be applied directly to two-phase models. Values of θx2 derived by GB05 need to be re-considered.
We re-estimate θx2 values for the sample of H ii regions of Esteban et al. (2002). Following GB05, three values of Tc are
assumed, 100, 1000, and 4000 K. Under these conditions, temperature in the cold gas is too low to collisionally excite the [O iii]
lines, and consequently Eq. (7) can be simplified to
T ([O iii]) = Th .
(16)
Substituting values of T0 and t2 given by Esteban et al. (2002) [hereafter referred as T0,E and t2
E, in order to distinguish them from
those deduced from Eqs. (12) and (13)] into Eqs. (3) and (4), we obtain T ([O iii])(t2
E,T0,E), which are tabulated in
Rows 3 and 4 of Table 1. Note that here we take the same assumption by GB05 that T0,E = T ([O iii]). Then θx2 can be determined
from Eq. (11), where T (H i) = T (H i)(t2
E,T0,E) thus obtained are
slightly higher than the actual values deduced from observations (see Table 1), the resulting θx2 are hardly affected.
E, T0,E). Although the values of T ([O iii])(t2
E, T0,E), Th = T ([O iii])(t2
E,T0,E) and T (H i)(t2
In Table 1 we compare our θx2 values to those of GB05 (given in parentheses); for low values of θx2 differences of up to a
factor of a hundred are found. It can easily be seen that the discrepancies increase with decreasing temperature of the cold gas, as
suggested by Fig. 1. For Tc = 100 K, our derived values of θx2 are very low, suggesting that the values of t2 reported by Esteban et
al. (2002) can be explained by the existence of a very small amount of cold gas.
Table 1 also gives t2 and T0 values derived from the Eqs. (12) and (13). As the Table shows, the real t2 values are lower when
Tc ≪ T0 than those derived from the empirical method. As a result, values of the cosmic ray ionization rate, ζ, derived by GB05
have been grossly overestimated [c.f. their Eq. (17)]. Our conclusion is consistent with the range of values inferred for the Orion
nebula from Gamma ray observations.
4. Conclusion
We have studied the relationship between values of t2 predicted by a two-phase model and those derived empirically from observa-
tions (empirical method). Our results show that the existence of extremely cold gas within H ii regions may lead to overestimated t2
calculated from empirically determined T ([O iii]) and T (H i). We stress that care should be taken when using the two-phase model
to study large temperature fluctuations of H ii regions. In this model, CELs are hardly produced by the cold gas, which on the other
hand makes a large contribution to the flux at the Balmer jump, due to the T −3/2
dependence of I(Bal, 3646). Accordingly, the
existence of a very small amount of cold material may lead to a large discrepancy between T ([O iii]) and T (H i). In other words,
in spite of its small mass, the existence of extremely cold material can reproduce apparently large t2 (as derived from the empirical
method), much larger than the actual value [as defined by Eq. (2)].
Finally, we revisited the GB05 study of cosmic ray ionization as a mechanism for creating temperature fluctuations in H ii
regions. While this provides a potential mechanism for creating cold ionized plasma in H ii regions, we show that the values of
ζ required to produce the ionization have been overestimated in their treatment, due to the t2 discrepancy discussed above. Based
on the formulae presented here, we re-estimated their model parameters. The corrections are apparent, particularly in cases where
temperature of the cold gas component is low, resulting in lower values of ζ that agree better with the estimates for the Orion nebula
published in the literature.
e
Acknowledgements. We thank the referee, Dr. C. Morisset, for helpful comments that improved clarity of the paper. We would also like to thank Dr. Morisset for
computing new values of t2 and T0, based on our estimates of θx2. Those values are now tabulated in Table 1. YZ and XWL acknowledge support by NSFC grant
#10325312.
Y. Zhang et al.: Temperature fluctuations in H ii regions: t2 for the two-phase model
5
References
Esteban, C., Peimbert, M., Torres-Peimbert, S., & Rodr´ıguez, M. 2002, ApJ, 581, 241
Garnett, D. R. 1992, AJ, 103, 1330
Giammanco, C., & Beckman, J. E. 2005, A&A, 437, L11 (GB05)
Peimbert, M. 1967, ApJ, 150, 825
Peimbert, M. , Peimbert, A., Ruiz, M. T., & Esteban, C. 2004, ApJ, 150, 431
Stasi´nska, G. 2000, RevMaxAA, 9, 158
Stasi´nska, G. 2002, RevMaxAA, 12, 62
Torres-Peimbert, S., & Peimbert, M. 2003, in IAU Symp. 209, Planetary Nebulae: Their Evolution and Role in the Universe, ed. P. R. Wood & M. Dopita (San
Francisco: ASP), p363
Viegas, S. M., & Clegg, R. E. S. 1994, MNRAS, 271, 993
Zhang, Y., Liu, X.-W., Liu, Y., & Rubin, R. H. 2005, MNRAS, 358, 457
|
astro-ph/0308291 | 1 | 0308 | 2003-08-18T06:37:38 | On the Quintessence with Abelian and Non-abelian Symmetry | [
"astro-ph",
"hep-th"
] | We study the perturbations on both "radial" and "angular" components of the quintessence with an internal abelian and non-abelian symmetry. The properties of the perturbation on the "radial" component depend on the specific potential of the model and is similiar for both abelian and non-abelian case. We show that the consine-type potential is very interesting for the O(\textit{N}) quintessence model and also give a critical condition of instability for the potential. While the properties of perturbations on "angular" components depend on whether the internal symmetry is abelian or non-abelian, which we have discussed respectively. In the non-abelian case, the fluctuation of the "angular" component will increase rapidly with time while in the abelian case it will not. | astro-ph | astro-ph | On the Quintessence with Abelian and Non-abelian Symmetry
Xin-zhou Li,∗ Jian-gang Hao, Dao-jun Liu, and Xiang-hua Zhai
Shanghai United Center for Astrophysics,
Shanghai Normal University, Shanghai 200234 ,China
(Dated: November 15, 2018)
Abstract
We study the perturbations on both "radial" and "angular" components of the quintessence with
an internal abelian and non-abelian symmetry. The properties of the perturbation on the "radial"
component depend on the specific potential of the model and is similiar for both abelian and non-
abelian case. We show that the consine-type potential is very interesting for the O(N ) quintessence
model and also give a critical condition of instability for the potential. While the properties of
perturbations on "angular" components depend on whether the internal symmetry is abelian or
non-abelian, which we have discussed respectively. In the non-abelian case, the fluctuation of the
"angular" component will increase rapidly with time while in the abelian case it will not.
3
0
0
2
g
u
A
8
1
1
v
1
9
2
8
0
3
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
∗Electronic address: [email protected]
1
1. Introduction
Recent observations[1, 2, 3] show that our Universe is flat and 73 percent of its total energy
density is resulted from "dark energy", which has a negative pressure and can accelerate the
expansion of the Universe[4, 5]. using recent announcement by the WMAP satellite[3], we
still do not judge whether dark energy is due to an unchanging, uniform and inert "vacuum
energy" (also known as a cosmological constant) or a dynamic cosmic field that changes with
time and varies across space (known as quintessence). Quintessence can be considered as a
spatially homogeneous scalar field that evolves with time and many models[6, 7, 8, 9, 10, 11]
have been constructed so far. The possibility of quintessence as rolling tachyon has also
been investigated [12].
Boyle et al[13] and Gu and Hwang[14] have discussed quintessence with complex scalar
field.
In a previous paper [15], we have further generalized their ideas by replacing the
complex scalar field with a N -plet scalar field which is spinning in a O(N )-symmetric po-
tential. When one of the angular components is fixed, this O(N ) quintessence model will
reduce to the O(N -1) quintessence model. If the N -2 angular components are fixed, O(N )
quintessence model can reduce to the complex scalar model mentioned above. If all angular
components are fixed, O(N ) quintessence model will reduce to the quintessence.
It is worth noting that this generalization does not hold its importance for the widely
studied tracker-type potentials because the amplitude of the scalar field will increase steadily
and make the angular contribution negligible. But for another type widely investigated
potential, the cosine-type potential[9, 15], it will be very interesting. It has been pointed
out that a natural introduction of quintessence is an ultralight axion with an almost massless
quark[9] and the cosine-type potential can be derived from such a model. Furthermore, This
potential requires that the amplitude of the field should not goes to be very large and
therefore, unlike the tracker-type potential, make the angular contribution significant.
It would be very interesting to study the behaviours of the field when it is perturbed.
We firstly investigate the fluctuation of the "radial" component and find that its stability
depends on the specific potential of the model while has nothing to do with whether the
the internal symmetry is abelian and non-abelian. We give a critical condition of the insta-
bility for the "radial" perturbation, under which the "radial" fluctuation will grow rapidly
with time. Fortunately, we find that this condition is not satisfed by most potentials for
quintessence models. While the fluctuation of the "angular" components depend heavily
2
on whether the internal symmetry is abelian or non-abelian. In the abelian case (N = 2),
the "angular" fluctuation will always damp, but in the non-abelian case(N > 2), there are
possibilities for the "angular" perturbations to grow rapidly with time and become strongly
space-dependent. It is worth noting that when we take the metric fluctuations(which are
generally small compared with the fluctuations of the quintessence field) into account, the
above conclusions are still held. In order to make the discussion more clear, in this paper
we will restrict ourselves to the N = 2 and N = 3 cases which represent the abelian and
nonabelian symmetry respectively.
2. The Model
We start from the flat Robertson-Walker metric
ds2 = dt2 − a2(t)(dx2 + dy2 + dz2)
The Lagrangian density for the quintessence with O(3) symmetry is
LΦ =
1
2
gµν(∂µΦa)(∂νΦa) − V (Φa)
(0.1)
(0.2)
where Φa is the component of the scalar field, a = 1, 2, 3. To make it possess a O(3)
symmetry, we write it in the following form
Φ1 = R(t) cos ϕ1(t)
Φ2 = R(t) sin ϕ1(t) cos ϕ2(t)
Φ3 = R(t) sin ϕ1(t) sin ϕ2(t)
(0.3)
Therefore, we have Φa = R and assume that the potential of the O(3) quintessence depends
only on R. It is clear that when we set the component ϕ2 to zero, the above O(3) system
will reduce to the O(2) abelian case.
The Einstein equations and equations of motion for the scalar fields can be written as
H 2 = (
a
a
)2 =
8πG
3
[
1
2
( R2 +
Ω2
a6R2 ) + V (R)]
(
a
a
) = −
8πG
3
[ R2 +
Ω2
a6R2 − V (R)]
3
(0.4)
(0.5)
R + 3H R −
Ω2
a6R3 +
∂V (R)
∂R
= 0
ϕ1 + (3H + 2
R
R
) ϕ1 − sin ϕ1 cos ϕ1 ϕ2
2 = 0
ϕ2 + (3H + 2
R
R
) ϕ2 + 2 cot ϕ1 ϕ1 ϕ2 = 0
(0.6)
(0.7)
(0.8)
where H is Hubble parameter and the term Ω2
a6R2 comes from the first integrals of the equations
of motion for the angular components(detailed discussion see[15]).
3. General Equations of Motion for "Radial" and "Angular" Perturbation
In order to eliminate the ambiguity from gauge freedom, one has to identify gauge in-
variant quantities or choose a given gauge and perform the calculations of perturbation in
that gauge. In this paper, we will carry out our investigation in synchronous gauge just as
Ratra and Peebles have done in Ref.[4]. As we shall see, in synchronous gauge the evolution
of the modes that we consider here depend on the gravitational field only through a gauge
invariant quantity
h
2 , so we can consider a single perturbation mode conveniently in the
following sections.
The line element of perturbed metric of a spatially flat FRW spacetime is taken as
ds2 = dt2 − a2(t)(δij − hij)dxidxj
(0.9)
where hij are the metric fluctuations and hij << 1. For simplicity, we write the first-order
equations of perturbations for the N = 3 case. Decomposing the components of the field as
follow
R(t, x) = R(t) + δR(t, x)
ϕ1(t, x) = ϕ1(t) + δϕ1(t, x)
ϕ2(t, x) = ϕ2(t) + δϕ2(t, x)
(0.10)
(0.11)
(0.12)
We can obtain the equations of motion for the fluctuations as
4
(δR) −
1
a2 ∇2(δR) − [ ϕ1
2 + sin2 ϕ1](δR) + 3
(δR)
a
a
′′
+V
(R)(δR) −
−R sin(2ϕ1) ϕ2
h R − 2 ϕ1R( δϕ1)
1
2
2(δϕ1) − 2R sin2(ϕ1) ϕ2( δϕ2) = 0
(0.13)
(δϕ1) −
1
a
a2 ∇2(δϕ1) + 3
a
1
h ϕ1 − sin(2ϕ1) ϕ2( δϕ2)
2 cos(2ϕ1)(δϕ1) −
2
(δϕ1) + 2
( δϕ1)
R
R
ϕ1( δR) −
R ϕ1(δR) = 0
(0.14)
− ϕ2
+
2
R
2
R2
(δϕ2) −
1
a2 ∇2(δϕ2) + 3
a
a
(δϕ2) + 2
R
R
( δϕ2) −
h ϕ2
1
2
+2 cot(ϕ1)[ ϕ1(δϕ2) + ϕ2( δϕ1)] − 2 csc2(ϕ1) ϕ1 ϕ2(δϕ1)
+
2
R
ϕ2( δR) −
2
R2
R ϕ2(δR) = 0
(0.15)
h + 2H h = 2 R (δR) + 2R2 ϕ1(δ ϕ1) + 2R2 sin2(ϕ1) ϕ2( δϕ2)
+ 2
Ω2
a6R3 (δR) + R2 sin(2ϕ1) ϕ2
2(δϕ1) − V
′
(R)(δR)
(0.16)
h,i − hij,j =
R∂i(δR) + R2 ϕ1∂i(δϕ1)
+ R2 sin2(ϕ1) ϕ2∂i(δϕ2)
(0.17)
1
a2 (hij,kk + h,ij − hik,jk − hjk,ik) − 3H hij
(R)(δR)
−H hδij − hij = δijV
′
(0.18)
where δR, δϕ1 and δϕ2 are fluctuations of the " radial" and "angular" components re-
spectively. The above equations of motion for fluctuations are the most general case for
5
quintessence with O(3) internal symmetry. When we consider only the "radial" perturba-
tion, we set ϕ1, ϕ2, δϕ1 and δϕ1 to zero. If we deal with the perturbation on the "angular"
component in the abelian case, what we need to do is to set another "angular" component
ϕ2 together with its fluctuation δϕ2 to zero. In the subsequent sections, we will discuss them
respectively.
4. Perturbation on the "Radial" Components
As we have pointed out in the introduction, the property of the perturbation on the
"radial" component depend only on the potential of the model and is independent of whether
the internal symmetry is abelian or non-abelian. In this section, we will show this in detail.
As we all know that scalar fields with an internal symmetry are likely to produce Q balls
or other non-topological solitons, which have been studied by many authors[16, 17, 18].
In their work, they generally put the gravitational effects aside because the magnitude of
gravitational fluctuations induced by the fluctuations of the scalar fields are far smaller
than the self-interaction of the scalar fields. Firstly, we, following the previous study in this
field[16, 17, 18], also investigate the perturbation without considering the metric fluctuation.
After this, we take the gravitational effects(metric fluctuations)into account and carry out
a similar study on the fluctuations. The equations of motion for fluctuations of the "radial"
component are
(δR) + 3H (δR) −
1
a2 ∇2(δR) +
Ω2
a6R4 (δR)
(R)(δR) = 0
′′
+V
(0.19)
which is obtained by setting δϕ1, δϕ2 and h in Eq.(0.13) to zero.
If we choose for the
fluctuation the following form:
δR(t, x) = δR0 exp[α(t) + ikx]
(0.20)
then for nontrivial δR0 , we have
α + α2 + 3H α −
Ω2
a6R4 +
k2
a2 +
∂2V
∂R2 = 0
(0.21)
Following the authors in Ref.[16, 17], we assume that α(t) is a slow-varying function, i.e.
α(t) << α2 and α ≈ Const. In the following sections, we shall always hold this assumption.
Therefore, neglecting the α in the above equation, we have
6
α =
1
2(cid:20) − 3H ±r(3H)2 − 4(
k2
a2 −
Ω2
a6R4 +
∂2V
∂R2 )(cid:21)
(0.22)
If
α is real and positive, the fluctuation will grow rapidly with time. Therefore the
instability band for this fluctuation is
From Eq.(0.23), one can find that the instability band depends on the specific potential of
0 < k2 <
Ω2
a4R4 −
a2∂2V
∂R2
(0.23)
the quintessence.
When considering the metric fluctuations, one can obtain the following equations of
motion for the fluctuations
(δR) + 3H (δR) −
Ω2
a6R4 (δR)
1
h R = 0
2
1
a2 ∇2(δR) +
(R)(δR) −
+V
′′
h + 2H h = 2 R (δR) + 2
Ω2
a6R3 (δR) − V
′
(R)(δR)
h,i − hij,j = R∂i(δR)
1
a2 (hij,kk + h,ij − hik,jk − hjk,ik) − 3H hij
(R)(δR)
−H hδij − hij = δijV
′
(0.24)
(0.25)
(0.26)
(0.27)
If we choose Ω = 0, i.e.
in the case of N = 1 the Eqs.(0.24)-(0.27) will reduce to Ratra-
Peebles's results in the absence of baryonic term[4]. Clearly, since the equations of motion
for the metric and scalar fluctuations(Eq.(0.25) and Eq.(0.24)) are linear equations, the
fluctuations could be taken as the following form
δR(t, x) = δR0 exp[α(t) + ikx]
(0.28)
7
h(t, x) = h0 exp[α(t) + ikx]
(0.29)
Since there are no ∇2h term in Eq.(0.25), h can not oscillate rapidly and in fact, h will
not be able to react, in lowest order, to the rapidly oscillating source terms in Eq.(0.25)[4].
Then for nontrivial δR0 and h0, we have
α2 +(cid:18)3H −
1
2
h0
δR0
R(cid:19) α −
Ω2
a6R4 +
k2
a2 +
∂2V
∂R2 = 0
So
α =
1
2
h0
δR0
R)
1
2(cid:20) − (3H −
±s(3H −
1
2
h0
δR0
R)2 − 4(
k2
a2 −
Ω2
a6R4 +
∂2V
∂R2 )(cid:21)
(0.30)
(0.31)
From Eq.(0.31), it is clear that the instability band is the same as that when we did not
consider the metric fluctuations. This is reasonable in physics because the metric fluctuations
induced by the fluctuations of the field are far smaller than the fluctuations of the field and
its back-reaction to the field would be even smaller and thus negligible. So, when considering
the metric fluctuations, there should not be a substantial change on the properties of the
fluctuation of the quintessence field.
5. Perturbation on the "Angular" Components in the Abelian Case
In this section, we investigate the perturbation on "angular" component in the abelian
case, that is, the case in which the quintessence fields possess a O(2) internal symmetry.
By setting δR, h, δϕ2 and ϕ2 in Eq.(0.14) to zero, we can obtain the equation of motion
for the fluctuation of "angular" component up to the first order as:
δϕ1 + (3H + 2
R
R
) δϕ1 −
1
a2 ∇2δϕ1 = 0
where δϕ1 is the fluctuation of the "angular" component. If we choose
δϕ1(t, x) = δφ10 exp[α(t) + ikx]
then for nontrivial δϕ10, from Eq.(0.32) we have
α2 + (3H + 2
R
R
) α +
k2
a2 = 0
8
(0.32)
(0.33)
(0.34)
α =
1
2(cid:20) − (3H + 2
) ±s(3H + 2
R
R
R
R
)2 − 4
k2
a2(cid:21)
(0.35)
It is clear that
From Eq.(0.35),
α is always negative and thus the fluctuation will damp quickly. That is,
the angular perturbation will not produce a significant "angular" inhomogeneity and the
global symmetry will not likely to become space-dependent.
In the following, we will introduce the metric fluctuations into our analysis. By a similar
procedure as that in last section, we can obtain the equations of motion for the fluctuations
of metric and "angular" component up to the first order as following:
h + H h = R2 ϕ1 δϕ1
1
2
h,i − hij,j = R2 ϕ1∂i(δϕ1)
1
a2 (hij,kk + h,ij − hik,jk − hjk,ik) − 3H hij
−H hδij − hij = 0
δϕ1 + (3H + 2
R
R
) δϕ1 −
1
a2 ∇2δϕ1 −
1
2
h ϕ1 = 0
(0.36)
(0.37)
(0.38)
(0.39)
Since Eq.(0.36)and Eq.(0.39) are linear equations, the fluctuations could be chosen in the
following form
δϕ1(t, x) = δϕ10 exp[α(t) + ikx]
h(t, x) = h0 exp[α(t) + ikx]
Then for nontrivial δϕ10, we have (from Eq.(0.39))
α2 + (3H + 2
R
R
−
1
2
h0
δϕ10
ϕ1) α +
k2
a2 = 0
9
(0.40)
(0.41)
(0.42)
Therefore, one can obtain
α =
1
R
R
−
1
2
h0
δϕ10
ϕ1)
(0.43)
2(cid:20) − (3H + 2
±s(3H + 2
R
R
−
1
2
h0
δϕ10
ϕ1)2 − 4
k2
a2(cid:21)
From Eq.(0.43), it is clear that α will always be negative even if the metric fluctuations are
considered.
6. Perturbation on the "Angular" Components in the Non-abelian Case
In this section, we generalize the discussions in last section to the case in which the
internal symmetry is non-abelian, that is the symmetry group is O(3). We restrict ourselves
to the case that only one "angular" component is perturbed. This will not lose its generality
but greatly facilitate the discussion because we can always choose a coordinate system in
which the perturbation appears in one angular direction. By setting δR, δϕ2 and h in
Eq.(0.14) to zero, we can obtain the equation of motion for the angular fluctuation up to
the first order as following:
δϕ1 + (3H + 2
R
R
) δϕ1 −
1
a2 ∇2δϕ1 − cos 2ϕ1 ϕ2
2δϕ1 = 0
(0.44)
where ϕ1 and ϕ2 are the homogeneous parts of the "angular" components. If we choose for
δϕ1 the same form as in Eq.(0.40), then for nontrivial δϕ10, from Eq.(0.44)we have
α2 + (3H + 2
R
R
) α +
k2
a2 − cos 2ϕ1 ϕ2
2 = 0
It is clear that
α =
1
2(cid:20) − (3H + 2
R
R
) ±s(3H + 2
R
R
)2 − 4(cid:18) k2
a2 − cos 2ϕ1 ϕ2
2(cid:19)(cid:21)
From Eq.(0.46), one can find that α could be positive if
k2
a2 − cos 2ϕ1 ϕ2
2 < 0
(0.45)
(0.46)
(0.47)
That is, under the above condition(Eq.(0.47)) the "angular" inhomogeneity might grow
rapidly with time and make the symmetry group become space-dependent.
10
When taking into account the metric fluctuation and following a similar process as that in
section 5, we can obtain the equations of motion for the fluctuations of "angular" component
as follow
δϕ1 + (3H + 2
R
R
) δϕ1 −
1
a2 ∇2δϕ1
1
2δϕ1 −
2
− cos 2ϕ1 ϕ2
h ϕ1 = 0
(0.48)
The equations of motion for metric fluctuations are
h + 2H h = 2R2 ϕ1(δ ϕ1) + R2 sin(2ϕ1) ϕ2
2(δϕ1)
(0.49)
h,i − hij,j = R2 ϕ1∂i(δϕ1)
1
a2 (hij,kk + h,ij − hik,jk − hjk,ik) − 3H hij
−H hδij − hij = 0
(0.50)
(0.51)
Similarly, we choose the fluctuations to be the form of Eq.(0.40) and Eq.(0.41) and for
nontrivial δϕ10, from Eq.(0.48) we have
α2 + (3H + 2
R
R
−
1
2
h0
δϕ10
ϕ1) α − cos 2ϕ1 ϕ2
2 +
k2
a2 = 0
and therefore
α =
1
2(cid:20) − (3H + 2
R
R
−
1
2
h0
δϕ10
ϕ1) ±
s(3H + 2
R
R
−
1
2
h0
δϕ10
ϕ1)2 − 4(cid:18) k2
a2 − cos 2ϕ1 ϕ2
2(cid:19)(cid:21)
(0.52)
(0.53)
From Eq.(0.53), one can easily identify that the condition for positive α is the same as that
in Eq.(0.47). This shows that in the O(3) case, the instability condition for the "angular"
fluctuations won't be changed when considering the metric fluctuation, just as that we have
proved in the O(2) case.
11
7. Conclusion And Discussion
In this paper, we investigate the perturbation on both "radial" and "angular" components
of the quintessence fields with an internal abelian and non-abelian symmetry. We find
that the fluctuation of the "radial" component depends on the specific potential of the
quintessence model. Under certain condition, this fluctuation could grow rapidly and thus
make the "radial" component space-dependent. For the "angular" perturbation, we find
that the properties of the fluctuations depend on whether the internal symmetry group is
abelian or non-abelian. In the abelian case, the fluctuation of the "angular" component will
damp rapidly with time and therefore the symmetry group will not become space-dependent.
In the case that the internal symmetry is non-abelian, we find that under certain condition,
the "angular" inhomogeneities might increase rapidly with time and make the symmetry
group space-dependent, or local. Here we briefly interpret the physical meaning of this
instability: δϕ1 represents a small internal angular inhomogenity; This inhomogenity will
increase rapidly under the condition(Eq.(0.47)). Note that this angular inhomogenity is not
that of space-time. It is surely interesting to study the quintessence with a "local" internal
symmetry, which we will investigate in a preparing work. When taking into account the
metric fluctuations induced by the fluctuations of the quintessence field and assuming that
the back-reaction of the metric fluctuations on the quintessence field are negligible as shown
in Ref.[4], the above conclusion still hold true.
It is worth noting that we choose the non-abelian symmetry group as O(3) in this paper.
It is not difficult to generalize the O(3) case to O(N ) case and one may find that in the
O(N ) case, the above conclusion for the non-abelian symmetry group still hold true even if
the specific condition under which the inhomogeneity increase will change.
ACKNOWLEDGMENTS
This work was partially supported by National Nature Science Foundation of China under
Grant No. 19875016, and Foundation of Shanghai Development for Science and Technology
No. 01JC14035.
[1] P. de Bernardis et al., Nature 404, 955 (2000);S. Hanany et al. Astrophys. J. 545, 1 (2000).
12
[2] N. Bahcall, J. P. Ostriker, S. Perlmutter and P. J. Steinhardt, Science 284, 1481 (1999);S.
Perlmutter et al., Astrophys. J. 517, 565 (1999); A. G. Riess et al., Astron. J. 116, 1009
(1998), astro-ph/9805201.
[3] C. L. Bennett et al., [astro=ph/0302207]; D. N. Spergel et al., [astro-ph/0302209]; G. Hinshaw
et al.,[astro-ph/0302217].
[4] B. Ratra and P. J. Peebles, Phys. Rev. D37, 3406 (1988).
[5] R. R. Caldwell, R. Dave and P. J. Steinhardt, Phys. Rev. Lett. 80, 1582 (1998).
[6] P. J. Steinhardt, L . Wang and I . Zlatev, Phys. Rev. D59, 123504 (1999), astro-ph/9812313.
[7] I. Zlatev, L. Wang and P. J. Steinhardt, Phys. Rev. Lett. 82,896 (1999), astro-ph/9807002.
[8] K. Coble, S. Dodelson, J. Frieman, Phys. Rev. D55, 1851 (1997).
[9] J. E. Kim, JHEP 9905, 022 (1999);Y. Nomura, T. Watari and T. Yanagida, Phys. Lett. B484,
103(2000); Phys. Rev. D61, 105007 (2000).
[10] T. Chiba, Phys. Rev. D64, 103503 (2001).
[11] A. Masiero, M. Pietroni and F. Rosati, Phys. Rev. D61, 023504 (2000); T. Barreiro, E. J.
Copeland and N. J. Nunes, Phys. Rev. D61, 127301 (2000); E. J. Copeland, N. J. Nunes and
F. Rosati, Phys. Rev. D62, 123503 (2000).
[12] X. Z. Li, J. G. Hao and D. J. Liu, Chin. Phys. Lett. 19, 1584(2002); J. G. Hao
and X. Z. Li, Phys. Rev. D66, 087301(2002); J. S. Bagla, H. K. Jassal and T.
Padmanabhan,[astro-ph/0212198]; T. Padmanabhan, [hep-th/0212290]; X. Z. Li and X. H.
Zhai, Phys. Rev. D67 067501(2002).
[13] L. A. Boyle, R. R. Caldwell and M. Kamionkowski, Phys. Lett. B545, 17(2002).
[14] Je-An Gu and W-Y. P. Hwang, Phys. Lett. B517, 1 (2001).
[15] X. Z. Li, J. G. Hao and D. J. Liu, Class. Quantum Grav. 19, 6049(2002); X. Z. Li, D. J. Liu
and J. G. Hao, Chin. Phys. Lett. 19, 295(2002); X. Z. Li and J. G. Hao,[hep-th/0303093].
[16] A. Kusenko, M. Shaposhnikov, Phys. Lett. B418, 46 (1998); S. Kasuya and M. Kawasaki,
Phys. Rev. D61, 041301 (2000); Phys. Rev. D62, 023510 (2000).
[17] S. Kasuya, Phys. Lett. B515, 121 (2001), astro-ph/0105408; X. Z. Li and X. H. Zhai, Phys.
Lett. B364, 212 (1995); X. Z. Li, X. H. Zhai and G. Chen, Astropart. Phys. 13, 245 (2000):
X. Z. Li, J. G. Hao, D. J. Liu and G. Chen , J. Phys. A34, 1459 (2001); X. Z. Li and J. G.
Hao, Phys. Rev. D66, 107701(2002); J. G. Hao and X. Z. Li, Class. Quantum Grav. 20, 1703
(2003).
13
[18] S. Coleman, Nucl. Phys. 262, 263 (1985).
14
|
astro-ph/0206018 | 1 | 0206 | 2002-06-03T05:01:11 | Chandra Snapshot Observations of LINERs with a Compact Radio Core | [
"astro-ph"
] | The results of Chandra snapshot observations of 11 LINERs (Low-Ionization Nuclear Emission-line Regions), three low-luminosity Seyfert galaxies, and one HII-LINER transition object are presented. Our sample consists of all the objects with a flat or inverted spectrum compact radio core in the VLA survey of 48 low-luminosity AGN (LLAGN) by Nagar et al. (2000). An X-ray nucleus is detected in all galaxies except one and their X-ray luminosities are in the range 5x10^38 to 8x10^41 ergs s^-1. The X-ray to Halpha luminosity ratios for 11 out of 14 objects are in good agreement with the value characteristic of LLAGNs and more luminous AGNs, and indicate that their optical emission lines are predominantly powered by a LLAGN. For three objects, this ratio is less than expected. Comparing with multi-wavelength results, we find that these three galaxies are most likely to be heavily obscured AGN. We compare the radio to X-ray luminosity ratio of LLAGNs with those of more-luminous AGNs, and confirm the suggestion that a large fraction of LLAGNs are radio loud. | astro-ph | astro-ph | Proceedings of the Workshop: "X-ray spectroscopy of AGN with Chandra and XMM-Newton"
held in Garching, December 3-6, 2001 eds. Th. Boller, S. Komossa, S. Kahn, H. Kunieda, MPE Report, pp.
2
0
0
2
n
u
J
3
1
v
8
1
0
6
0
2
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
Chandra Snapshot Observations of LINERs with a Compact
Radio Core
Y. Terashima1,2 and A.S. Wilson2,3
1 Institute of Space and Astronautical Science, 3-1-1 Yoshinodai, Sagamihara, Kanagawa 229-8510, Japan
2 Astronomy Department, University of Maryland, College Park, MD 20742, USA
3 Space Telescope Science Institute, 3700 San Martin Drive, Baltimore, MD 21218, USA
Abstract. The results of Chandra snapshot observations
of 11 LINERs (Low-Ionization Nuclear Emission-line Re-
gions), three low-luminosity Seyfert galaxies, and one HII-
LINER transition object are presented. Our sample con-
sists of all the objects with a flat or inverted spectrum
compact radio core in the VLA survey of 48 low-luminosity
AGN (LLAGN) by Nagar et al. (2000). An X-ray nucleus
is detected in all galaxies except one and their X-ray lu-
minosities are in the range 5 × 1038 to 8 × 1041 ergs s−1.
The X-ray to Hα luminosity ratios for 11 out of 14 ob-
jects are in good agreement with the value characteristic
of LLAGNs and more luminous AGNs, and indicate that
their optical emission lines are predominantly powered by
a LLAGN. For three objects, this ratio is less than ex-
pected. Comparing with multi-wavelength results, we find
that these three galaxies are most likely to be heavily ob-
scured AGN. We compare the radio to X-ray luminosity
ratio of LLAGNs with those of more-luminous AGNs, and
confirm the suggestion that a large fraction of LLAGNs
are radio loud.
paucity of obscured AGN in LINERs may indicate that
LINER 2s are not simply a low-luminosity extension of
luminous Seyfert 2s, which often show heavy obscuration
with a column density averaging NH ∼ 1023 cm−2 (e.g.,
Turner et al. 1997). Alternatively, biases against finding
heavily obscured LLAGNs may be important. For exam-
ple, objects selected through optical emission lines or X-
ray fluxes are probably biased in favor of less absorbed
ones, even if one uses the X-ray band above 2 keV.
In contrast, radio observations, particularly at high fre-
quency, are much less affected by absorption. Nagar et al.
(2002) have reported a VLA 2 cm radio survey of all 96
LLAGNs within a distance of 19 Mpc. These LLAGNs
come from the Palomar spectroscopic survey of bright
galaxies (Ho et al. 1997a). As a pilot study of the X-ray
properties of LLAGNs, we report here a Chandra survey
of a subset, comprising 14 galaxies, of Nagar et al's (2002)
sample. We have detected 13 of the galactic nuclei with
Chandra. We also examine the "radio loudness" of our
sample and compare it with other classes of AGN.
1. Introduction
Low-Ionization nuclear emission-line regions (LINERs)
are found in many nearby bright galaxies (e.g., Ho, Fil-
ippenko, & Sargent 1997a). Extensive studies in various
wavelengths have shown that type 1 LINERs (LINER 1s,
i.e., those galaxies having broad Hα and possibly other
broad Balmer lines in their nuclear optical spectra) are
powered by a low-luminosity AGN (LLAGN) with a bolo-
metric luminosity less than ∼ 1042 ergs s−1 (Ho et al.
2001; Terashima, Ho, & Ptak 2000a; Ho et al. 1997b). On
the other hand, the energy source of LINER 2s is likely to
be heterogeneous. Some LINER 2s show clear signatures
of the presence of an AGN, while others are most proba-
bly powered by stellar processes, and the luminosity ratio
LX/LHα can be used to discriminate between these power
sources (e.g., Terashima et al. 2000b). It is interesting to
note that currently there are only a few LINER 2s known
to host an obscured AGN (e.g., Turner et al. 2001). This
2. The Sample and Observations
Our sample is based on the 15 GHz VLA observations by
Nagar et al. (2000). Their sample of 48 objects consists
of 22 LINERs, 18 transition objects between LINERs and
HII nuclei, and eight low-luminosity Seyferts selected from
the optical spectroscopic survey of Ho et al. (1997a).
We selected 14 objects showing a flat to inverted spec-
trum radio core (α ≥ −0.3, Sν ∝ να) according to Nagar
et al.'s (2000) comparison with longer wavelength radio
data published in literature. The targets are summarized
in Table 1. The sample consists of seven LINER 1s, three
LINER 2s, two Seyfert 1s, one Seyfert 2, and one tran-
sition 2 object. 12 out of these 14 objects have been ob-
served with the VLBA and high brightness temperature
(Tb > 107 K) radio cores were detected in all of them (Fal-
cke et al. 2000; Ulvestad & Ho 2001; Nagar et al. 2002).
The exposure time was typically two ksec each. All the
objects were observed with the ACIS-S3 back-illuminated
CCD chip. Eight objects were observed in 1/8 sub-frame
mode (frame time 0.4 s) to minimize effects of pileup. 1/2
2
Y. Terashima & A.S. Wilson: Chandra Snapshot Observations of LINERs with a Compact Radio Core
Table 1. The Sample.
Class
Name
NGC 266, 2787, 3226, 4143, 4203, 4278, 4579
LINER 1
NGC 3169, 4548, 6500
LINER 2
NGC 4565, 5033
Seyfert 1
Seyfert 2
NGC 3147
Transition 2 NGC 5866
sub-frame modes were used for three objects. Detailed re-
sults are given in Terashima & Wilson (2002c).
3. Results
An X-ray nucleus is seen in all the galaxies except for NGC
5866. The X-ray luminosities corrected for absorption are
in the range 5 × 1038 to 8 × 1041 ergs s−1. The positions
of the X-ray nuclei coincide with the radio core positions
to within the positional accuracy of Chandra.
Spectral fits were performed for relatively bright ob-
jects. The pileup effect for the three objects with the
largest count rate per frame (NGC 4203, NGC 4579, and
NGC 5033) is serious and we did not attempt detailed
spectral fits. Instead, we use the spectra and fluxes mea-
sured with ASCA for these three objects (Terashima et al.
2002b and references therein) in the following discussions.
We confirmed that the nuclear X-ray source dominates the
hard X-ray emission within the beam size of ASCA.
A power-law model modified by absorption was applied
and acceptable fits were obtained in all cases. The photon
indices of the nuclear sources are generally consistent with
the typical values observed in LLAGNs (photon index Γ =
1.6 − 2.0, e.g., Terashima et al. 2002a, 2002b), although
errors are quite large due to the limited photon statistics.
The two objects (LINER 2s NGC 3169 and NGC 4548)
show large absorption column density NH=1.1×1023 cm−2
and 1.6 × 1022 cm−2 , respectively, while NGC 3226 is
less absorbed (NH=9.3 × 1021 cm−2). Others have small
column densities which are consistent with 'type 1' AGNs.
NGC 2787 has only 8 detected photons in the 0.5 -- 8 keV
band and is too faint to obtain spectral information.
Seyferts, and QSOs presented in Terashima et al. (2000a)
and Ho et al. (2001). This indicates that their optical emis-
sion lines are predominantly powered by a LLAGN.
The three objects NGC 2787, NGC 5866, and NGC
6500, however, have much lower LX/LHα ratios (log
<
∼ 0) than expected from the correlation, and
LX/LHα
their X-ray luminosities are not enough to power the Hα
luminosities. This X-ray faintness could indicate one or
more of several possibilities such as (1) an AGN is the
power source, but is heavily absorbed at energies above
2 keV, (2) an AGN is the power source, but is currently
switched-off or in a faint state, and (3) the optical narrow
emission lines are powered by some other source(s) than
an AGN.
If an AGN is present in these X-ray faint objects and
absorbed in the hard energy band above 2 keV, only scat-
tered and/or highly absorbed X-rays can be observed,
and then the intrinsic luminosity would be much higher
than that observed. This can account for the low LX/LHα
ratios and high radio to X-ray luminosity ratios (νLν(5
GHz)/LX). If the intrinsic X-ray luminosities are about
one or two orders of magnitude higher than those ob-
served, as is often inferred for Seyfert 2 galaxies, LX/LHα
and νLν(5 GHz)/LX become typical of LLAGNs.
Additional lines of evidence which support the pres-
ence of an AGN include the fact that all three of these
galaxies (NGC 2787, NGC 5866, and NGC 6500) have
VLBI-detected, sub-pc scale, nuclear radio core sources
(Falcke et al. 2000), a broad Hα component (in NGC
2787, and an ambiguous detection in NGC 5866; Ho et al.
1997b), a variable radio core in NGC 2787, and a jet-like
linear structure in a high-resolution radio map at 5 GHz
with the VLBA (NGC 6500; Falcke et al. 2000). Only an
upper limit to the X-ray flux is obtained for NGC 5866.
If an X-ray nucleus is present in this galaxy and its lumi-
nosity is only slightly below the upper limit, this source
could be an AGN obscured by a column density NH∼ 1023
cm−2 or larger. If the intrinsic luminosity of the nucleus is
much lower than the observed upper limit, an AGN would
have to be almost completely obscured and/or the optical
emission lines powered by some other source(s). The opti-
cal classification (transition object) suggests the presence
of an ionizing source other than an AGN.
4. Discussion
4.1. Power Source of LINERs
4.2. Obscured LLAGNs
We test whether the detected X-ray sources are the power
source of their optical emission lines by examining the lu-
minosity ratio LX/LHα. The Hα luminosities (LHα) were
taken from Ho et al. (1997a) and corrected for the redden-
ing estimated from the Balmer decrement for the narrow
lines. The X-ray luminosities (LX) in the 2 − 10 keV band,
and corrected for absorption, were used. The resulting
LX/LHα ratios of most objects are in the range of AGNs
>
∼ 1) and in good agreement with the strong
(log LX/LHα
correlation between LX and LHα for LLAGNs, luminous
In our sample, we found at least two highly absorbed
LLAGNs (NGC 3169 and NGC 4548). In addition,
if
the X-ray faint objects discussed in the previous subsec-
tion are indeed AGNs, they are most probably highly
absorbed with NH> 1023 cm−2. Among these absorbed
objects, NGC 2787 is classified as a LINER 1.9, NGC
3169, NGC 4548, and NGC 6500 as LINER 2s, and NGC
5866 as a transition 2 object. Thus, heavily absorbed
LINER 1.9s/2s, of which few are known, are found in the
present observations demonstrating that radio selection is
Y. Terashima & A.S. Wilson: Chandra Snapshot Observations of LINERs with a Compact Radio Core
3
show high absorption columns in their X-ray spectra. In
this subsection, we study radio loudness by comparing ra-
dio and hard X-ray luminosities. Since the unabsorbed
∼ 1023 cm−2 (equivalent
>
luminosity for objects with NH
>
∼ 50 mag for a normal gas to dust ratio) can be
to AV
reliably measured in the 2 -- 10 keV band, it is clear that re-
placement of optical by hard X-ray luminosity potentially
yields considerable advantages.
In the following analysis, radio data at 5 GHz taken
from the literature are used since fluxes at this frequency
are widely available for various classes of objects. We used
the radio luminosities primarily obtained with the VLA at
<
∼ 1′′ resolution for the present sample. High resolution
VLA data at 5 GHz are not available for several objects.
For four objects among such cases, VLBA observations at
5 GHz with 150 mas resolution are published in the lit-
erature (Falcke et al. 2000) and are used here. For two
objects, we estimated 5 GHz fluxes from 15 GHz data by
assuming a spectral slope of α = 0 (cf. Nagar et al. 2001).
Since our sample is selected based on the presence of a
compact radio core, the sample could be biased to more
radio loud objects. Therefore, we constructed a larger sam-
ple by adding objects taken from the literature for which
5 GHz radio, 2 -- 10 keV X-ray, and RO measurements are
available.
First, we introduce the ratio RX = νLν(5 GHz)/LX as
a measure of radio loudness and compare the ratio with
the conventional RO parameter. The X-ray luminosity LX
in the 2 -- 10 keV band (source rest frame), corrected for
absorption, is used. We examine the behavior of RX using
samples of AGN over a wide range of luminosity, including
LLAGN, the Seyfert sample of Ho & Peng (2001) and PG
quasars which are also used in their analysis. The X-ray
luminosities (mostly measured with ASCA) are compiled
from the literature.
Fig. 2 compares the parameters RO and RX for the
Seyferts and PG sample. These two parameters correlate
well for most Seyferts. Some Seyferts have higher RO val-
ues than indicated by most Seyferts. This could be a result
of extinction. Seyferts showing X-ray spectra absorbed by
a column greater than 1022 cm−2 (NGC 2639, 4151, 4258,
4388, 4395, 5252, and 5674) are shown as open circles in
Fig. 2. At least four of them have larger RO than indi-
cated by the correlation. The correlation between log RO
and log RX for the less absorbed Seyferts can be described
as log RO = 0.88 log RX + 5.0. According to this relation,
the boundary between radio loud and radio quiet object
(log RO = 1) corresponds to log RX = −4.5.
The PG quasars show systematically lower RO values
than those of Seyferts at a given log RX. For the former
objects, log RO = 1 corresponds to log RX = −3.5. This
apparently reflects a luminosity dependence of the shape
of the SED: luminous objects have steeper optical-X-ray
slopes αox = 1.4 − 1.7 (S ∝ ν−α), where αox is often
measured as the spectral index between 2200 A and 2
keV, while less luminous AGNs have αox = 1.0 − 1.2 (Ho
Fig. 1. Examples of Chandra spectra. (a) NGC 3169 and
(b) NGC 4548
a valuable technique for finding obscured AGNs. Along
with heavily obscured LLAGNs known in low-luminosity
Seyfert 2s (e.g., NGC 2273, NGC 2655, NGC 3079, NGC
4941, and NGC 5194; Terashima et al. 2002a), our obser-
vations show that at least some type 2 LLAGNs are sim-
ply low-luminosity counterparts of luminous Seyferts in
which heavy absorption is often observed. Some LINER
2s (e.g., NGC 4594, Terashima et al. 2002a; NGC 4374,
Finoguenov & Jones 2001; NGC 4486, Wilson & Yang
2002) and low-luminosity Seyfert 2s (NGC 3147) show no
strong absorption. Therefore, the orientation dependent
unification scheme does not always apply to AGNs in the
low-luminosity regime.
4.3. Radio Loudness of LLAGNs
Earlier studies have suggested that LLAGNs tend to be
radio loud compared to more luminous AGNs based on the
spectral energy distributions of seven LLAGNs (Ho 1999)
and, for a larger sample, on the conventional definition
of radio loudness RO = Lν(5 GHz)/Lν(B) (the subscript
"O" stands for optical), with RO > 10 being radio loud
(Ho & Peng 2001). Ho & Peng (2001) measured the lumi-
nosities of the nuclei by spatial analysis of optical images
obtained with HST to reduce the contribution from stellar
light. A caveat in the use of optical measurements for the
definition of radio loudness is extinction, which will lead
to an overestimate of RO. Although Ho & Peng (2001)
used only type 1 -- 1.9 objects, some objects of these types
4
Y. Terashima & A.S. Wilson: Chandra Snapshot Observations of LINERs with a Compact Radio Core
Fig. 3. X-ray luminosity dependence of RX = νLν(5
GHz)/LX for the present LLAGN sample, Seyfert galax-
ies, and PG quasars. The boundary between "radio loud"
and "radio quiet" objects (log Rx = −4.5) is shown as a
horizontal dashed line.
References
Falcke, H., Nagar, N. M., Wilson, A. S., & Ulvestad, J., S.
2000, ApJ, 542, 197
Finoguenov, A. & Jones, C. 2001, ApJ, 547, L107
Ho, L. C. 1999, ApJ, 516, 672
Ho, L. C., et al. 2001, ApJ, 549, L51
Ho, L. C., Filippenko, A. V., & Sargent, W. L. W. 1997a, ApJS,
112, 315
Ho, L. C., Filippenko, A. V., Sargent, W. L. W., & Peng, C. Y.
1997b, ApJS, 112, 391
Ho, L. C., & Peng, C. Y. 2001, ApJ, 555, 650
Nagar, N. M., Falcke, H., Wilson, A. S., & Ho, L. C. 2000, ApJ,
542, 186
Nagar, N. M., Falcke, H., Wilson, A. S., & Ulvestad, J. S. 2002,
A&A, submitted
Nagar, N. M., Wilson, A. S., & Falcke, H. 2001, ApJ, 559, L87
Terashima, Y., Ho, L. C., Iyomoto, N., & Ptak, A. F. 2002a,
ApJ, in preparation
Terashima, Y., Ho, L .C., & Ptak, A. F. 2000a, ApJ, 539, 161
Terashima, Y., Ho, L .C., Ptak, A. F., et al. 2000b, ApJ, 533,
729
Terashima, Y., Iyomoto, N., Ho, L. C., & Ptak, A. F. 2002b,
ApJS, 139, 1
Terashima, Y. & Wilson, A. S. 2002c, ApJ, submitted
Turner, M. J., et al. 2001, A&A, 365, L110
Turner, T. J., George, I. M., Nandra, K., & Mushotzky, R. M.
1997, ApJS, 113, 23
Ulvestad, J. S., & Ho, L. C. 2001, ApJ, 562, L133
Wilson, A. S. & Yang, Y 2002, ApJ, 568, 133
Fig. 2. Comparison between RO = Lν(5 GHz)/Lν(B)
and RX = νLν(5 GHz)/LX for Seyferts and PG quasars.
The conventional boundary between "radio loud" and "ra-
dio quiet" objects (log RO = 1) is shown as a horizontal
dashed line.
1999). This is related to the fact that luminous objects
show a more prominent "big blue bump" in their spectra.
Figure 8 of Ho (1999) demonstrates that low-luminosity
objects are typically 1 -- 1.5 orders of magnitude fainter in
the optical band than luminous quasars for an given X-ray
luminosity.
The definition of radio loudness using the hard X-ray
flux (RX) appears to be more robust because X-rays are
less affected by both extinction at optical wavelengths and
the detailed shape of the blue bump. Further, measure-
ments of nuclear X-ray fluxes are much easier than mea-
surements of nuclear optical fluxes, since in the latter case
the nuclear light must be separated from the surrounding
starlight.
Fig. 3 shows the X-ray luminosity dependence of RX.
In this plot, the LLAGN sample discussed in the present
paper is shown in addition to the Seyfert and PG samples
used above. This is an "X-ray version" of the log RO-M nuc
plot (Fig. 4 in Ho & Peng 2001). Our plot shows that a
large fraction of LLAGNs (LX< 1042 ergs s−1) are radio
loud. This is a confirmation of Ho & Peng's (2001) finding.
Since radio emission in LLAGNs is likely to be dominated
by emission from jets (Nagar et al. 2001; Ulvestad & Ho
2001), these results suggest that, in LLAGN, the fraction
of the accretion energy that powers a jet, as opposed to
electromagnetic radiation, is larger than in more luminous
Seyfert galaxies and quasars.
B
Acknowledgements. Y.T. is supported by the Japan Society for
a Promotion of Science Postdoctoral Fellowship for Young Sci-
entists. This research was supported by NASA through grants
NAG81027 and NAG81755 to the University of Maryland.
|
astro-ph/0305547 | 3 | 0305 | 2004-08-19T20:20:03 | Extracting the Dark Matter Profile of a Relaxed Galaxy Cluster | [
"astro-ph"
] | Knowledge of the structure of galaxy clusters is essential for an understanding of large scale structure in the universe, and may provide important clues to the nature of dark matter. Moreover, the shape of the dark matter distribution in the cluster core may offer insight into the structure formation process. Unfortunately, cluster cores also tend to be the site of complicated astrophysics. X-ray imaging spectroscopy of relaxed clusters, a standard technique for mapping their dark matter distributions, is often complicated by the presence of cool components in cluster cores, and the dark matter profile one derives for a cluster is sensitive to assumptions made about the distribution of this component. In addition, fluctuations in the temperature measurements resulting from normal statistical variance can produce results which are unphysical. We present here a procedure for extracting the dark matter profile of a spherically symmetric, relaxed galaxy cluster which deals with both of these complications. We apply this technique to a sample of galaxy clusters observed with the Chandra X-ray Observatory, and comment on the resulting mass profiles. For some of the clusters we compare their masses with those derived from weak and strong gravitational measurements. | astro-ph | astro-ph |
Draft version October 21, 2018
Preprint typeset using LATEX style emulateapj v. 11/12/01
EXTRACTING THE DARK MATTER PROFILE OF A RELAXED GALAXY CLUSTER
J.S. Arabadjis1, M.W. Bautz1, and G. Arabadjis2
Draft version October 21, 2018
ABSTRACT
Knowledge of the structure of galaxy clusters is essential for an understanding of large scale structure
in the universe, and may provide important clues to the nature of dark matter. Moreover, the shape of
the dark matter distribution in the cluster core may offer insight into the structure formation process.
Unfortunately, cluster cores also tend to be the site of complicated astrophysics. X-ray imaging spec-
troscopy of relaxed clusters, a standard technique for mapping their dark matter distributions, is often
complicated by the presence of cool components in cluster cores, and the dark matter profile one derives
for a cluster is sensitive to assumptions made about the distribution of this component. In addition,
fluctuations in the temperature measurements resulting from normal statistical variance can produce
results which are unphysical. We present here a procedure for extracting the dark matter profile of a
spherically symmetric, relaxed galaxy cluster which deals with both of these complications. We apply
this technique to a sample of galaxy clusters observed with the Chandra X-ray Observatory, and com-
ment on the resulting mass profiles. For some of the clusters we compare their masses with those derived
from weak and strong gravitational measurements.
Subject headings: X-rays : galaxies: clusters -- cosmology : dark matter
1.
introduction
The cold dark matter (CDM) paradigm of modern cos-
mology has enjoyed spectacular success in describing the
formation of large-scale structure in the universe (Navarro,
Frenk & White 1997; Moore et al. 1999b; Lahav et al.
2001; Peacock et al. 2001). There are, however, several
nagging inconsistencies between the results of numerical
CDM experiments and observations. On small scales, the
dark matter halos in dwarf and low surface brightness
galaxies are much less cuspy than in CDM simulations
(Burkert 1995; McGaugh & de Blok 1998; Moore et al.
1999b). Disk galaxies produced in simulations tend to
have inadequate masses and angular momenta (Navarro &
Steinmetz 2000). The number of Milky Way satellites ap-
pears to be at least an order of magnitude lower than CDM
predictions (Kauffman, White, & Guideroni 1993; Moore
et al. 1999a; Klypin et al. 1999). On larger scales, some
studies (Tyson, Kochanski & Dell'Antonio 1998; Smail
et al. 2000) report galaxy clusters with central density
profiles that are flatter than CDM predictions, although
these are somewhat controversial (Broadhurst et al. 2000;
Shapiro & Iliev 2000).
The density profile of bound structures which form
through the hierarchical assembly of smaller structures is
usually parameterized as a power law at small scales and
a separate power law on large scales (e.g. Jing & Suto
(2002)):
ρ(r) =
(r/rs)α(1 + r/rs)γ−α
ρ0
(1)
1.0 (Navarro, Frenk & White 1996, 1997) and 1.5 (Moore
et al. 1999b; Fukushige & Makino 2001), roughly indepen-
dent of halo mass and formation epoch. In nature, how-
ever, α shows a larger variation, and is likely a function
of halo mass. Hα rotation curves of low-surface bright-
ness galaxies indicate density profiles which are signifi-
cantly flatter -- α ∼ 0.5 -- than CDM predictions (Swaters,
Madore & Trewhella 2001; Dalcanton & Bernstein 2000;
Borriello et al. 2003; Swaters et al. 2003). X-ray obser-
vations of galaxy clusters generally show steeper profiles,
however, with α ∼ 1.2 (Lewis, Buote & Stocke 2002) to
1.9 (Arabadjis, Bautz & Garmire 2002).
These discrepancies are often ascribed to limitations of
the astrophysics or the physics included in the simulations.
Baryon physics, if included, may be tacked on at the con-
clusion of a simulation according to a set of semi-analytic
and/or empirical prescriptions.
It is likely that baryon
physics will play a significant role in the evolution of the
central halo. Reports of a halo "entropy floor" (Ponman,
Cannon & Navarro 1999; Lloyd-Davies, Ponman, & Can-
non 2000) suggest non-gravitational sources of heating and
feedback either prior to or during halo formation (Balogh,
Babul & Patton 1999; Loewenstein 2000; Wu, Fabian &
Nulsen 2000) which are probably baryonic in origin (see
Mushotzky et al. (2003), however). The question then be-
comes one of determining where baryon physics ceases to
be important. While the inclusion of baryon astrophysics
in sufficient detail may remedy these problems, its effects
will require a great deal of effort to disentangle (Frenk
2002).
The four parameters in this description are the density ρ0
at some fiducial radius, the inner power law index α, the
outer power law index γ, and the scale radius rs setting
the break between the two power laws. While it is gen-
erally agreed that γ = 3, the value of α has generated
considerable debate. Simulations predict a value between
1 Center for Space Research, Massachusetts Institute of Technology, Cambridge, MA 02139; [email protected], [email protected]
2 Mitre Corporation, 202 Burlington Road, Bedford, MA 01730; [email protected]
It could be, however, that the missing ingredients in
the simulations are not all astrophysical. One possibility
is that the initial power spectrum of the primordial fluc-
tuations is not scale invariant. If the primordial spectral
index of density perturbations is not precisely 1 (as is nor-
mally assumed by appealing to standard inflationary cos-
1
2
mology), the formation epoch of halos may be delayed suf-
ficiently to ameliorate the central density problem (Alam,
Bullock & Weinberg 2002; Zentner & Bullock 2002). An-
other possibility is that important dark matter particle
physics is being overlooked, and that the assumption of
no non-gravitational self-interactions is faulty. Proposed
modifications of CDM include, though are not limited to,
self-interacting dark matter (Spergel & Steinhardt 2000;
Firmani et al. 2000), warm dark matter (Hogan & Dalcan-
ton 2000), annihilating dark matter (Kaplinghat, Knox &
Turner 2000), scalar field dark matter (Hu & Peebles 2000;
Goodman 2000), and mirror matter (Mohapatra, Nussinov
& Teplitz 2002), each of which is invoked to soften the core
density profile. Many of these modifications will soften the
core profile of galaxy clusters as well, in conflict with many
X-ray determinations of mass profiles, although other as-
trophysical processes such as the adiabatic contraction of
core baryons (Hennawi & Ostriker 2002) may mitigate this
effect.
In an effort to discriminate between CDM modifica-
tions and other astrophysical influences we are mapping
the dark matter profiles of a large sample of galaxy clus-
ters. Specifically, we use imaging spectroscopy from the
Chandra X-ray Observatory (Weisskopf et al. 2002) to de-
termine the deprojected temperature and density profiles
of the baryonic content of each galaxy cluster, which we
then use to derive its dark matter profile. In this paper
we describe our method, and apply it to a sample of low-
and moderate-redshift clusters. We describe our spectral
deprojection technique in §2; we discuss the problems in-
volved in converting these results to a mass profile and our
solution in §3; we examine the effects that cooling flow
model assumptions have on our profiles, and we present
a statistical analysis of the models and a prescription for
choosing among them using Markov Chain Monte Carlo
sampling in §4. Finally, we summarize our findings in §6.
In a subsequent paper we will examine these profiles for
their implications for large scale structure formation and
dark matter particle properties (Arabadjis & Bautz, in
preparation [AB]).
2. spectral deprojection
We begin with a Chandra imaging spectroscopic obser-
vation of a galaxy cluster using the ACIS detector, either
with the S3 chip or the I array. We start with a level 2 data
set that has been processed in the usual way, filtered for
periods of high background using the procedures described
in the CIAO Science Threads3, with point sources removed
(for details see Arabadjis, Bautz & Garmire (2002)). Af-
ter locating the center of the projected emissivity, we lay
down a series of adjacent, concentric annuli centered on the
emission peak. The annular dimensions are set to include
enough source photons (1000-2000+) to reliably determine
the plasma temperature. RMF and ARF response matri-
ces are constructed, as is a background spectrum from a
region external to the outermost annulus. The spectra are
recorded in PI format, and grouped such that there is a
minimum of 20 counts per channel.
Our spectral deprojection has been presented elsewhere
(Arabadjis, Bautz & Garmire 2002), and so we just briefly
summarize here. To derive spherical radial profiles we con-
3 See http://asc.harvard.edu/ciao/threads/all.html
struct a model consisting of N concentric spherical shells
whose inner and outer radii correspond to the inner and
outer cylindrical radii of the projected annuli in the data
set. The volume intersection matrix V, whose elements
Vij contain the volume of spherical shell j intersected by a
cylindrical shell formed by the projection of annulus i, is
used to set the linear relations between each of the normal-
izations as specified by the binning geometry. Each shell
i on [1, N ] (1 is the innermost shell -- a sphere -- and N
is the outermost shell) contains an optically thin thermal
plasma whose emission characteristics are determined by
the MEKAL model (Mewe, Gronenschild & van den Oord
1986; Mewe, Lemen & van den Oord 1986; Kaastra 1992;
Liehdal, Osterheld & Goldstein 1995) within XSPEC (Ar-
naud 1996) using two free parameters, the temperature T
and the normalization K.
In some models we will allow the innermost Nc shells to
contain a second emission component at a (lower) temper-
ature Tc as a first-order treatment of the cooler plasma.
This cool component is assumed to be in pressure equi-
librium with the hot component; this means that the cool
and hot components cannot both be in hydrostatic equi-
librium. We assume that the hot component is in hydro-
static equilibrium. We adopt this form for the cooling flow
model for two reasons: (1) there is no evidence that the
plasma in cooling flow cluster cores cools below about 1
keV (Peterson et al. 2003) as in the isobaric cooling flow
model of Mushotzky & Szymkowiak (2003), and (2) it is
arguably the simplest adjustment that can be made to the
uniphase model. The cool component could be arranged in
droplets or filaments which are replenished as they migrate
to the unspecified sink at r = 0, effecting a hydrodynam-
ical equilibrium. The details of the geometry of the cool
component are unimportant since they are below our spa-
tial resolution limit; it is only our assumption that it is
in pressure equilbrium with the hot component which has
observational consequences. In this study we use the pres-
sure gradient in the core to measure the radial dependence
of the enclosed gravitating mass.
The number of parameters in this model is rather large.
Each MEKAL component contains six parameters, of
which two (T and K) are allowed to vary. Thus the N
annuli in the data set are modelled using N + Nc emis-
sion components. Including a Galactic absorption column
yields a model with 6(N + Nc) + 1 parameters (although
only a subset of these actually vary). Each of the N annuli
independently constrain between 1 and N of the model
shells in the fitting, and so the complete XSPEC model
contains N [6(N + Nc) + 1] components, 2(N + Nc) + 1
of them variable. The current version of XSPEC admits
models with 1000 parameters, any 100 of which can vary.
This limits our model to N = 12 annuli for Nc = 0 (876
parameters, 25 variable) and N = 10 annuli for Nc = 2
(730 parameters, 25 variable). We have written a program
which reads in the data annuli dimensions and writes out
an XSPEC script that handles all of the data manipula-
tion. The script initializes the model parameters, performs
the χ2 minimization, and calculates parameter uncertain-
ties. This software is available to the public through re-
quests to the authors.
Most spectral deprojection schemes rely on an "onion
peeling" approach (Fabian et al. 1981; Allen & Fabian
1997; David et al. 2001; Lewis, Stocke & Buote 2002; Sun
et al. 2003): the outermost annulus is modeled using the
outermost spherical shell; its model parameters are then
frozen and its emission is subtracted from all annuli in-
terior to it. The next most outer shell is then modelled,
the resulting model again frozen and subtracted from the
interior, and so forth, until the entire cluster has been
modelled. The virtue of this technique is that the number
of model components scales as N instead of N 2, allow-
ing for greater spatial detail. However, because the pa-
rameters of each model shell are frozen and the model is
subtracted from interior shells as if it contained zero un-
certainty, the technique does not find the global "best fit"
of the model parameters. In many cases, the errors quoted
in onion-peeling analyses are the uncertainties associated
with a single layer of the onion (David et al. 2001). Error
estimates derived from Monte Carlo simulations are more
reliable, limited primarily by the number of simulations
used to estimate the uncertainties (Lewis, Stocke & Buote
2002).
In our study all of the model parameters are fit
simultaneously; the subsequent determination of error in
subsets of interesting variables is a true expression of the
parameter uncertainties in the model. We note here that
although our model appears to have many more parame-
ters than the other techniques, in actuality the numbers
are equivalent because of the web of linear dependences
among the component normalizations.
3. mass profiles
Many deprojection methods rely on analytic formulae
for the radial run of temperature, surface brightness or
mass, either during or after the fitting process (Allen 2001;
David et al. 2001; Hicks et al. 2002; Pizzolato et al. 2003).
The greatest advantage of a parametric treatment is nu-
merical stability.
In addition, it is common practice to
smooth noisy profiles before using them in subsequent cal-
culations. This latter technique is especially useful when
deriving gravitating mass profiles since a numerical deriva-
tive must be computed. A serious drawback of these ap-
proaches is that it is difficult to quantify the effect of the
parameterization, or the smoothing, on the results. Ad-
ditionally, it is often difficult or impossible to propagate
errors through to the results.
Our non-parametric deprojection technique does not
guarantee smooth temperature and density profiles, so
we have devised a method whereby the mass profile is
computed from within the error envelope of (ρ(r), T (r)).
By choosing a statistically reasonable realization of the
model, we are able to compute mass profiles which are
not only smoother than those obtained from the uncon-
strained (ρ, T ) set, but which avoid the unphysical results
that arise due to the statistical fluctuations inherent in
measurements. Essentially, our procedure imposes phys-
ically motivated constraints to reduce the uncertainty in
the temperature and mass profiles, and provides a statistic
which characterizes the reliability of the mass reconstruc-
tion.
The standard procedure for extracting the gravititating
mass profile of a galaxy cluster is to insert its deprojected
temperature and density profiles into the hydrostatic equa-
tion (Sarazin 1988):
Mr = −
kT
Gµmp/r (cid:18) d log T
d log r
+
d log ρ
d log r(cid:19) ,
Here T and ρ are the local (baryonic) plasma temperature
and density, r is the spherical radius, Mr is the total mass
enclosed within r (i.e. baryons plus dark matter), and mp
and µ are the proton mass and mean particle weight, re-
spectively. Implicit here is the assumption that the cluster
is supported solely by an isotropic thermal pressure gra-
dient, i.e. that random motions greatly exceed the bulk
rotational motion, and that the magnetic field energy den-
sity is negligible in comparison with the thermal energy
content of the plasma. We further assume that no recent
merger event has caused a disruption in the pressure and
density. Given the run of density and temperature in our
binning scheme (ri, ρi, and Ti, i = 1, 2, ..., N ), we can
calculate Mi using a simple difference equation version of
Equation 2:
Mi = −A ri Ti(cid:18) ti+1 − ti−1 + di+1 − di−1
xi+1 − xi−1
(cid:19)
(3)
where A = k/Gµmp, xi = log10 (ri), di = log10 (ρi), and
ti = log10 (Ti). Since the goal of this technique is to derive
enclosed mass profiles, we define r as the outer radius of
each shell.
Because of measurement error a mass profile calculated
in this way is not guaranteed to be physically reasonable.
Even if the assumptions of spherical symmetry and hy-
drostatic equilibrium were valid, statistical fluctuations in
the temperature measurements could result in unphysical
points in our derived mass profile, for example dMr/dr < 0
or even Mr < 0. To deal with the non-physical fluctua-
tions, we proceed under the assumptions that all unphys-
ical values in the mass profile are due to measurement
uncertainty in either ρ or T . Specifically, to estimate the
(ρ, T ) profiles we impose the condition that the computed
total gravitating mass profile is consistent with
ρ(total) = ρ(baryons) + ρ(dark matter) ≥ 0 for
3
(2)
r ≥ 0
(4)
dMr
dr and 1/4πr2 is positive definite,
Since ρ(total) = 1
this is equivalent to the constraint
dMr/dr ≥ 0
4πr2
(5)
Combining Equation 5 with the boundary condition
Mr(0) ≥ 0 (e.g., allowing for the presence of an unresolved
central object) we obtain a second constraint:
Mr(r) ≥ 0
(6)
Perhaps the most natural way to impose these condi-
tions would be to invoke them as a Bayesian prior in the
spectral fitting procedure. Bayes' Theorem (Bayes 1763;
Papoulis 1984) states that the probability of model M
given data set D (the posterior distribution P (MD)) is
proportional to the product of the probability of that data
set given the model (the likelihood function P (DM )) and
the probability of the model itself (the prior knowledge
function P (M)).
P (MD) ∝ P (DM) · P (M)
(7)
The model prior could be chosen to enforce the constraints
in Equations 5 and 6. Thus for certain combinations of
model parameters Ti, ρi, we could choose a prior such that
P (M(Ti, ρi)) = 0.
terms in a cost function, and henceforth refer to the opti-
mized solution for M as the constrained profile, with the
understanding that it is an estimate of the fully Bayesian
solution, and that the constraints do not rigorously for-
bid excursion into disfavored regions of parameter space.
Our formulation of the cost function will reflect our subjec-
tive assessment of the relative importance of fidelity versus
physicality, and use standard optimization algorithms to
minimize it.
We use the likelihood to characterize the fidelity term
in the cost function. The probability density P at the
vector (ρ, T ) = ( ρ1, ρ2, ..., ρN , T1, T2, ..., TN ), which is near
the unconstrained profiles (ρ, T ), is given by
N
Yi=1
P = (2π)N/2
(σρi σTi )−1 e−( ρi−ρi)2/2σ2
ρi e−( Ti−Ti)2/2σ2
Ti
(10)
where we have assumed that the errors in ρ and T are un-
correlated and normally distributed, characterized by σρi
and σTi , respectively.4 In maximum likelihood methods,
cost functions are conveniently defined as the negative log
of the likelihood (e.g. von Mises (1964)), so we define the
fidelity penalty function Q as
Q = −2 log P
= N log(2π) +
2
N
Xi=1 " log σρi + log σTi +
( ρi − ρi)2
2σ2
ρi
+
( Ti − Ti)2
2σ2
Ti
#
= Q0 + χ2
(11)
Here Q0 is a constant that depends only on the measure-
ment errors and χ2 is the 2N -dimensional variance:
χ2 =
N
Xi=1 " ( ρi − ρi)2
σ2
ρi
+
( Ti − Ti)2
σ2
Ti
#
= χ2
ρ + χ2
T
(12)
We will ignore Q0 since it will not affect the minimization.
We follow a similar procedure to derive the physical-
ity terms in the cost function. We wish to incorporate
the physicality contraints by penalizing negative values of
Yi and Zi in Equations 8 and 9 without attempting to
maximize them if they are positive. Consider the penalty
function
g(x) = (x/x0 − x/x0)η + (x/x0 − 1 + x/x0 − 1)η
4
Unfortunately this approach is computationally imprac-
tical. The difficulty lies in the fact that the relatively sim-
ple constraints on M (r) and dM/dr lead to complicated
(nonlinear) constraints of the coupled parameters of the
spectral models (Ki, Ti). The problem is more tractable
if we instead apply the Bayesian constraints after uncon-
strained density and temperature profiles have been deter-
mined using the spectral deprojection method described
in §2. Thus we compute an unconstrained mass profile
from Equation 3, and check it for consistency with the
constraints in Equations 5 and 6. These three equations
yield constraints on the solutions to the difference equa-
tions which we define using Yi and Zi:
Yi = −(ti+1 − ti−1 + di+1 − di−1) ≥ 0
(8)
Zi = −ri+1Ti+1(cid:18) ti+2 − ti + di+2 − di
+ riTi(cid:18) ti+1 − ti−1 + di+1 − di−1
xi+1 − xi−1
xi+2 − xi
(cid:19)
(cid:19) ≥ 0
(9)
Physically these constraints prohibit the pressure of the
X-ray emitting plasma from rising with radius. Using the
model parameter error estimates as a guide, we allow the
temperature and normalization of each model component
to vary if either of these constraints is violated by the pro-
file. Our goal is thus to find a new set of density and
temperature values ρ and T which are as close as possible
to the the original profiles ("fidelity") but which obey the
constraints of Equations 8 and 9 ("physicality").
At first glance this appears to be a standard problem
in constrained optimization. Two features of the prob-
lem suggest than an alternate route is preferable, how-
ever. First, as mentioned above, the constraints are non-
linear functions of the independent variables, so the con-
strained optimization approach is very complex. Second,
if we exercise control over the competing interests of fi-
delity and physicality, as incorporated in a penalty func-
tion, rather than simply eliminating solutions which vio-
late Equations 8 and 9, we retain the ability to modulate
departures from the unconstrained profiles. For example,
it may be the case that the temperature profile would need
substantial alteration in order to satisfy the rigid imposi-
tion of constraints 8 and 9, but that a minor violation of
Equation 9 (say, at only one point in the mass profile)
would allow a temperature profile of much greater fidelity.
In this instance it is to our advantage to have the abil-
ity to set the relative weighting between the fidelity and
physicality terms in the penalty function.
Implicit in our choice of a penalty function to represent
the constraints is the assumption that our estimate of the
location of the peak of the model probability distribution
function Pest(M(Ti, ρi)), obtained by enforcing fidelity, is
close to the actual peak of the distribution P0(M(Ti, ρi)),
which would obtain if the Bayesian priors were utilized
during the spectral deprojection. This idea is shown
schematically in Figure 1. As the figure illustrates, in the
absence of significant small-scale structure in P (M(Ti, ρi)),
Pest(M(Ti, ρi)) will ideed lie near P0(M(Ti, ρi)) in parame-
ter space. We therefore cast our physicality constraints as
4 In principle it would be more accurate to make use of the fact that the luminosity of a shell i, Li ∝ ρ2
observations, and therefore that deviations in d and t are correlated: δd = −
uncertainty in the temperature of a shell is enormous compared with the uncertainty in its density.
(η > 1)
− ln(cid:0) 1
2(cid:1) + ln(C)
(13)
This function has several useful properties: (1) it has a
value of zero for 0 ≤ x ≤ x0; (2) it increases rapidly for
x < 0 or x > x0; (3) it has a continuous first derivative
everywhere for η ≥ 2 (or everywhere but x = 0 or x = x0
for 1 < η < 2), making it well-suited for optimization rou-
tines that rely on gradient information to increase their
efficiency; and (4) h(x) = e−g(x) is a well-behaved proba-
bility distribution function. C is of order one; for arbitrary
values of η > 1 it can be determined numerically. g(x) and
h(x) are shown in Figure 2.
We employ g(x) to characterize the two physicality re-
quirements, Mr ≥ 0 and dMr/dr ≥ 0, in their difference
, is tightly constrained by the
δt
4 . In practice, however, this is unnecessary since the relative
i T 1/2
i
equation form (Equations 8 and 9). In both cases x0 is
set to a large value that we don't expect to exceed. In the
former case we set it to 1018 M⊙; in the latter we can use
the equivalent of d log(Mr)/d log(r) ≤ 3, corresponding
to the upper limit for a monotonically decreasing density
profile. We can ignore the normalization and set K = 1
since we will be weighting the physicality penalties against
the fidelity penalty. The complete cost function f for the
minimization is thus:
N
f =
[ A1 g(Yi) + A2 g(Zi) ] + χ2
(14)
Xi=1
The weights A1 and A2 and the exponent η are chosen to
determine whether the departure from the unconstrained
profile is reasonable to achieve a physically acceptable so-
lution.
Conveniently, this technique has a built-in gauge of the
fidelity, the χ2 value of the constrained profiles (although
it should be noted that it is not distributed as the clas-
sical χ2). If χ2 . 1 then we consider the new profile to
be a reasonable excursion within the profile's uncertainty
envelope. Moreover, we identify the constrained profiles
(ρ, T , Mr) as a closer approximation to the true profiles,
under the assumption that our model is correct. Con-
versely, χ2 ≫ 1 indicates that the (ρ, T ) profiles require
excessive alteration to produce a physically sensible mass
profile. This could indicate that our assumptions of spher-
ical symmetry and/or hydrostatic equilibrium are invalid,
that there is significant spatial or spectral substructure, or
that our thermal emission model is inadequate. Regardless
of the root cause, the accuracy of such a profile is suspect.
Nulsen & Bohringer (1995) (hereafter NB95) developed
a non-parametric approach to this problem which also
makes use of the fact that the enclosed gravitating mass
of a cluster must be monotonically increasing. With their
method one obtains a series of interdependent constraints
on the mass at each point in r. These constraints are trans-
lated into a set of likelihood functions which are assumed
to be independent for computational ease. These likeli-
hood functions are jointly maximized to derive the mass
profile. Our method is similar in that it makes use of a
likelihood function based upon excursions within an uncer-
tainty envelope. The methods differ significantly, however,
in the deprojection algorithm (NB95 use onion-peeling),
and in the enforcement of the mass constraints. In NB95,
the mass constraints are absolute; in our method they can
be invoked to any degree -- they can be rigidly enforced,
completely ignored, or somewhere in between. The χ2
metric of the fidelity is a powerful tool. If the physical-
ity weight is set to an arbitrarily large value, we force the
profile to monotonically increase, in effect obtaining the
result of the NB95 technique. As a biproduct we obtain
the plasma temperature and density profiles which are re-
quired, and the χ2 value immediately tells us how likely
it is that the temperature and density measurements con-
spired to achieve this condition. In cases where rigid im-
position of the constraints yields an unacceptably high χ2
value, one can experiment with different weights and ex-
amine the resulting profiles to ascertain, for example, if
the problem is due to a single outlier, or is more systemic,
possibly indicative of a breakdown of hydrostatic equilib-
rium. Thus the flexibility of our method with regard to the
5
mass constraints allows us to extract information about
the dynamical state of the X-ray plasma and the presence
of statistical anomalies in the data.
4. choosing a model: the f test and markov
chain monte carlo sampling
In the previous section we assumed that the underly-
ing model is correct. There is some uncertainty, however,
about the form the model should take. As mentioned pre-
viously, the best candidates for this type of analysis are
relaxed cooling flow clusters. These systems are currently
not well understood, yet the way we model the cool plasma
can have a significant effect on the resulting mass profile,
and conclusions we may draw about the properties of dark
matter particles or the evolution of large scale structure
(Arabadjis, Bautz & Garmire 2002).
A standard procedure for choosing between a simple
model Ms and a complex model Mc is to utilize the F
test (Bevington 1969). This is done by computing the F
value for the data set D:
F =
χ2(MsD) − χ2(McD)
χ2(MsD)/ν(Ms)
(15)
and comparing it to the standard F distribution. Here
χ2(MsD) and χ2(McD) are the sum of the squares of the
error-weighted residuals in the spectroscopic least-squares
fit to the simple and complex models, respectively, and
ν(Ms) is the number of degrees of freedom in the simple
model.
Unfortunately, the standard F test is not applicable in
this context. As Protassov et al. (2002) have pointed out,
the standard F test is valid only in cases where the simple
model is nested within the complex model. In the present
case, the simple model lies on a boundary of the complex
model. That is, the simple model is a special case of the
complex model, with the normalization of the second emis-
sion component in the core set to zero. This means that
its F distribution may deviate significantly from the norm
(see Figure 3). Instead we must construct an empirical F
distribution by sampling the probability distribution func-
tion of the simple model, simulating a data set for each
sample item, and applying both models to the simulated
data. Once our F distribution has been constructed, we
can judge the significance of the extra emission component
based upon the location within the distribution of the F
value for the data (Protassov et al. 2002).
We employ the Markov Chain Monte Carlo (MCMC)
sampling technique Neal (1993); van Dyk et al. (2001);
Lewis & Bridle (2002); Hobson & McLachlan (2002) to
build a large sample of data realizations from which to
construct an empirical F distribution. Let P (x) repre-
sent the posterior probability distribution function of the
parameters x = x1, x2, ..., xN determined by fitting the
simple model Ms to a real data set D0. We can sample
P (x) by taking a rejection-based random walk through
the parameter space. We define a transition probability
T (xn, xn+1) as the probability of moving from an initial
set of parameters xn to a new set of parameters xn+1. T
depends on the value of the posterior distribution at the
original and new parameter sets. Let q(xn, xn+1) be an ar-
bitrary proposal distribution; that is, the probability that
the new proposed parameter set is xn+1 given that we are
6
presently at xn. If we accept the proposed parameter set
with probability α, then
α(xn, xn+1) =
T (xn, xn+1)
q(xn, xn+1)
(16)
which takes into account the odds of actually stepping to
the new location in parameter space, q(xn, xn+1), and the
odds of such a transition between the two locations being
accepted, T (xn, xn+1). The acceptance probability α is
calculated from
α(xn, xn+1) = min(cid:20)1,
P (xn+1) q(xn+1, xn)
P (xn) q(xn, xn+1) (cid:21)
(17)
In our case we use a particular form of MCMC sam-
pling called the Metroplis algorithm (see, e.g., Neal (1993))
which uses a symmetric proposal distribution function:
q(xn, xn+1) = q(xn+1, xn)
(18)
This prescription for wandering through parameter
space constitutes a Markov chain since each new param-
eter set is chosen according to a probability distribution
function that depends only upon the previous set of values.
In the case of the Metropolis algorithm, it is straightfor-
ward to show that P (x) is an invariant distribution of the
Markov chain. Using Equations 16 and 17 we have
P (xn) T (xn, xn+1) = P (xn) q(xn, xn+1) α(xn, xn+1)
= q(xn, xn+1) min[P (xn), P (xn+1)](19)
Making use of the symmetry of q in the Metropolis algo-
rithm, we have
= q(xn+1, xn) min[P (xn), P (xn+1)]
= P (xn+1) q(xn+1, xn) α(xn+1, xn)
resulting in
P (xn) T (xn, xn+1) = P (xn+1) T (xn+1, xn)
(20)
This statement of detailed balance demonstrates that P (x)
is a stationary distribution of the Markov chain. This
is necessary, though not sufficient, to ensure that we can
sample P (x) directly using an appropriately selected chain
of Monte Carlo simulations. The other necessary condi-
tion, ergodicity, ensures that any substring of the Markov
chain will asymptotically approach P (x) regardless of the
initial conditions, although a derivation of this property is
beyond the scope of this paper. For a complete discussion
see Neal (1993).
In many applications of MCMC sampling one pays spe-
cial attention to the finite "burn-in" period during which
the Markov chain equilibrates. The length of the burn-in
phase depends upon the sensibility of the starting point,
and the appropriateness of the scale chosen for the pro-
posal probability distribution step. This is not a consider-
ation in our case because we start each MCMC sample at
the (already known) peak of the probability distribution
function P (x).
Given P (x) computed by the MCMC process, we can
construct an empirical F distribution and perform an F
test on the significance of a second emission component in
the core. The entire procedure is as follows:
1. Model the real data set D0 with Ms; call
the best-fit parameters x
s
0.
2. Use XSPEC to calculate P (D0x) (i.e., the
likelihood).
3. Use Bayes' Theorem to calculate P (xD0)
(see Equation 7). We discard all unphys-
ical excursions in parameter space,
i.e.
where T < 0 or ρ < 0. (In practice, we
discard at the level of the model normal-
ization, not ρ, but ρ is simply a function
of the normalization and the binning ge-
ometry.)
s
4. Create a large sample of model parame-
i using P (xD0) and the Metropo-
ters x
lis algorithm form of the MCMC tech-
i, compute a fake data
nique. For each x
set Di, including instrumental effects of
the Chandra telescope and detectors, as
well as counting statistics.
s
5. Model each Di using both Ms and Mc.
6. For each pair of models tabulate its F
value given by Equation 15.
7. Bin up the set of F values, creating an un-
normalized histogram , and superimpose
the F value of the original data.
In practice this recipe is computationally intensive, not
because of any features of the MCMC sampling per se,
but because each of the faked spectra must be modelled
twice. For a sample size of 1000 simulations, XSPEC must
simulate 1000 spectra and calculate 2002 sets of best-fit
values for the model parameters (including the original
data). This fact leads us to simplify the method. First,
in order to reduce the modelling time, we have adopted a
simplified core-halo geometry. In this scheme the "core" is
represented by a single shell (in this case a sphere), while
the halo is represented by another shell. Thus Ms contains
four parameters, the temperature and density of each of
the two shells, while Mc contains six, the additional two
parameters representing the temperature and density of
a second cospatial emission component in the core. This
simplification also greatly improves the numerical stability
of the fitting procedure. The algorithm (steps 1-6 above)
is implemented in a Tcl script run within XSPEC.
Once we have completed step 7 we can distinguish be-
tween the models. The location of the F value of the data
within this empirical F distribution contains information
regarding the relative merit of Ms and Mc. We define the
significance S of the distribution as
N (F ) dF
0 N (F ) dF
0
S = R Fdata
R ∞
(21)
The signficance S = 1 − Pf , where Pf is the probability
that the simple model constitutes the better description,
and that the F value of the data is this large strictly by
chance. Thus, for a one-parameter model, S = 0.68, 0.90,
and 0.99 may be interpreted as 1-, 2- and 3-σ detections
of the additional component. We checked the sensitiv-
ity of this method by applying it to five simulated data
sets with known mixtures of hot (T = 5.0 kev) and cold
(T = 1.0 kev) X-ray plasmas in pressure equilbrium. We
use mass ratios of MC/MH = 0.000, 0.0222, 0.0500, 0.0857,
and 0.133, and run 100 MCMC simulations for each. Fig-
ure 4 shows an empirical F distribution for each of these
cases. The F value of the data is shown with a dashed
vertical line. For a multiphase plasma of which only 2.2%
is in the cold component, the detection of the multiphase
plasma is better than 2σ. At 5% and greater, the detection
is statistically highly significant.
5. application to chandra clusters
We illustrate these techniques using a X-ray observa-
tions of a sample of bright, apparently relaxed galaxy clus-
ters (see Table 1). Each of these clusters (except A1689)
contains a significant amount of cooler plasma in its core,
and is known as a classical "cooling flow cluster" since
the core radiative cooling time is shorter than the age of
the cluster. We prepared each archived data set as de-
scribed in Arabadjis, Bautz & Garmire (2002) and mod-
elled the emission using the two models described in §2.
In the first model, each shell contains isothermal plasma
(Nc = 0), while in the other, the central two shells are also
allowed a second (cooler) emission component (Nc = 2),
and the best-fit parameter values are obtained iteratively
using XSPEC. Figures 5-14 show baryon density, baryon
temperature, enclosed spherical gravitating mass, and en-
closed cylindrical gravitating mass profiles for each cluster
in the sample. The left panels show Nc = 0 models; the
right show Nc = 2 model. The unconstrained profiles are
shown as data points, with red [blue] symbols represent-
ing the hot [cool] plasma components. (Data points which
lie outside the ordinate range are indicated by arrows.)
Constrained profiles are drawn as solid curves; the density
and temperature contributions to the χ2 value of the con-
strained solutions are listed in Table 2. The last column
in that table lists the MCMC significance S of the pres-
ence of multiphase gas in each cluster core as determined
from the MCMC-derived empirical F distributions shown
in Figure 15. For the subset of five clusters which are
arguably the most relaxed, and are best described by an
NFW profile (A1689, A1835, A2029, MS1358 and MS2137;
see AB), we overlay the projected mass profiles with weak
and/or strong lensing measurements from the literature
(see Table 3 for references). Weak lensing mass profiles or
isothermal sphere fits to weak lensing data are shown in vi-
olet; strong lensing measurements are shown in green. We
adopt H0 = 67 km s−1 Mpc−1, ΩΛ = 0.7 and Ωm = 0.3.
A detailed analysis of these profiles and their conse-
quences for cosmology and dark matter candidates is in
preparation (AB), so we make only a few brief comments
here. Of the ten clusters presented here, only HydraA
did not admit a second emission component in the core
-- attempts to add one resulted in the temperature of the
second component being set to the temperature of the first
in the fitting procedure. This is consistent with previous
Chandra (David et al. 2001) and XMM-Newton (Kaastra
2004) results. Four of the clusters show evidence for mul-
tiphase core plasma at the 99% significance level -- A2029,
A2204, MS1358 and ZW3146 (see Table 2). This is con-
sistent with the study of Kaastra (2004), which finds evi-
dence for multiphase plasma in many clusters. (Note that,
with 1000 MCMC simulations per cluster, the precision of
the significance estimate is ∼3%.) For comparison, a clus-
ter which contained no second plasma component would,
on the average, show an MCMC significance of order 0.5,
or 50%. These clusters show a greater difference in their
core masses between the uni- and multiphase models. This
7
phenomenon is illustrated for two clusters, MS1358 and
MS2137,
in Figures 16 and 17. The mass of MS1358
in the multiphase model is about a factor of two larger
in than the uniphase mass, and its MCMC significance
S=0.987. MS2137, on the other hand, with an MCMC
significance of 0.271, shows very little difference between
the two mass models. This is perhaps unsurprising, as one
would expect that a significant amount of cospatial cool
plasma in its core would not only display a clear observa-
tional signature but would also affect the equilibrium con-
figuration of the plasma. MS2137 and ZW3146 also pro-
vide an interesting contrast. In the pre-Chandra/XMM-
Newton era, both clusters were reported to harbor cooling
flows with mass deposition rates in excess of 1000 M⊙y−1
(Allen 2000), yet they show remarkably disparate evidence
for multiphase plasma in their cores (S(MS2137)=0.271;
S(ZW3146)=0.989). This suggests that there may be more
than one mechanism at work responsible for the presence
of 1 keV plasma at the center of galaxy clusters.
For each constrained reconstruction we use η = 2.5
and physicality weights A1 = A2 = A12 originally set to
1 × 10−6. While this often will not rigidly enforce con-
straint equation 5, it is usually sufficient to enforce equa-
tion 6. In those cases where not even equation 6 is satis-
fied, we increased A12 by factors of 10 until the resulting
profile was nonnegative everywhere if possible. This is the
origin of the unacceptably high values of the χ2 fidelity
measures for the Nc = 2 model of A2104 and the Nc = 1
model of HydraA. Obviously, the fidelity of a constrained
profile tends to be greater when the unconstrained pro-
file shows only one or two significant outliers.
In some
cases (e.g. A1795) a very small adjustment to the temper-
ature profile produces a large change in the derived mass.
This is due to the competition between the derivatives in
equation 2 -- if a statistical fluctuation in the tempera-
ture measurement is large enough (and positive) it will
swamp the surface brightness decrement at that radius,
resulting in a very low or even negative mass.
In cases
where d log T /d log r & −d log ρ/d log r, a relatively minor
adjustment in the temperature can remove an unphysical
point from the mass profile.
The five very relaxed clusters generally show better
agreement with weak lensing mass measurements than
they do with strong lensing. The reprojected profile A1689
shows the least agreement with the weak lensing, differing
by up to a factor of two at some radii, although the lensing
profile is a singular isothermal sphere (IS) fit to weak lens-
ing measurements (King, Clowe & Schneider 2002). The
strong lensing measurement of Wu (2000), however, ex-
ceeds our profile again by a factor of 2. The error bars are
derived by assuming two values of the (unknown) strong
lensing arc redshift (0.8 and 2.0); the discrepancy is thus
unlikely to be due to incorrect source redshift. Our pro-
file is also systematically in excess of the IS fit to weak
lensing observations of A1835 (Clowe & Schneider 2002),
although it is consistent with the strong lensing point of
Allen (1998). A2029 and MS1358 show remarkable agree-
ment with weak lensing data (Menard, Erben & Mellier
2003; Hoekstra et al. 1998), although the strong lensing
measurement in MS1358 (Franx et al. 1997; Allen 1998) is
moderately discrepant, as is the strong lensing measure-
ment of MS2137 (Sand, Treu & Ellis 2002). These issues
8
will be addressed in greater detail in AB.
6. summary
We have presented a technique for calculating the dark
matter profile of a spherical, relaxed galaxy cluster. We
have formulated a technique for coping with statistical un-
certainty in the measurement of the cluster plasma tem-
perature. We have also described a method for determin-
ing whether there is a statistically significant presence of
multiphase plasma in the galaxy cluster core. We have
applied these tools to a sample of relaxed galaxy clusters
observed with Chandra, and find that 4/10 require a mul-
tiphase treatment of their core plasma. Our masses are in
broad agreement with weak lensing studies, though are of-
ten exceeded by those derived from strong lensing models.
JSA would like to thank Steve Allen, Aaron Lewis,
Jimmy Irwin and Renato Dupke for many enlightening
discussions. The authors would like to thank the anony-
mous referee for comments which have improved this pa-
per. This work was supported by SAO Grant AR3-4016X.
REFERENCES
9
Alam, S.M.K., Bullock, J.S. & Weinberg, D.H. 2002, ApJ, 572, 34
Allen, S.W. 2000, MNRAS, 315, 269
Allen, S.W. 1998, MNRAS, 296, 392
Allen, S.W. & Fabian, A.C. 1997, MNRAS, 286, 583
Allen, S.W., Fabian, A.C., Johnstone, R.M., Arnaud, K.A. & Nulsen,
P.E.J. 2001, MNRAS, 322, 589
Arabadjis, J.S., Bautz, M.W. & Garmire, G.P. 2002, ApJ, 572, 78
Arnaud, K.A. 1996, Astronomical Data Analysis Software and
Systems V, George H. Jacoby & Jeannette Barnes, eds., ASP Conf.
Ser., 101, 17
Balogh, M.L., Babul, A. & Patton, D.R., 1999, MNRAS, 307, 463
Bayes, T. 1763, Phil. Trans. Roy. Soc., 53, 370
Bevington, P.R. 1969, Data Reduction and Error Analysis for the
Physical Sciences (New York: McGraw-Hill)
Boriello, A., Salucci, P. & Danese, L. 2003, MNRAS, 341, 1109
Broadhurst, T., Huang, X., Frye, B., & Ellis, R. 2000, ApJ, 534, 15
Burkert, A. 1995, ApJ, 447, L25
Clowe, D.I. & Schneider, P. 2002, A&A, 395, 385
Dalcanton, J.J. & Bernstein, R.A. 2000,
in XVth IAP Meeting,
Dynamics of Galaxies: From the Early Universe to the Present,
eds. F. Combes, G.A. Mamon & V. Charmandaris
David, L.P., Nulsen, P.E.J., McNamara, B.R., Forman, W., Jones,
C., Ponman, T., Robertson, B. & Wise, M. 2001, ApJ, 557, 546
Fabian, A.C., Hu, E.M., Cowie, L.L. & Grindlay, J. 1981, ApJ, 248,
47
Firmani, C., D'Onghia, E., Avila-Reese, V., Chincarini, G. &
Hern´andez, X. 2000, MNRAS, 315, L29
Franx, M., Illingworth, G.D., Kelson, D.D., van Dokkum, P.G. &
Tran, K.-V. 1997, ApJ, 486, L75
Frenk, C.S. 2002, Phi. Trans. Roy. Soc., 300, 1277
Fukushige, T. & Makino, J. 2001, ApJ557533
Goodman, J. 2000, New Astron., 5, 103
Hennawi, J.F. & Ostriker, J.P. 2002, ApJ, 572, 41
Hicks, A.K., Wise, M.W., Houck, J.C. & Canizares, C.R. 2002, ApJ,
580, 763
Hobston, M.P. & McLachlan, C. 2003, MNRAS, 338, 765
Hoekstra, H., Franx, M. & Kuijken, K. 1998, ApJ, 504, 636
Hogan, C.J. & Dalcanton, J.J. 2000, Phys. Rev. D, 62, 063511
Hu, W. & Peebles, P.J.E. 2000, ApJ, 528, 61
Jing, Y.P. & Suto, Y. 2002, ApJ, 574, 538
Kaastra, J.S., Tamura, T., Peterson, J.R., Bleeker, J.A.M., Ferrigno,
C., Kahn, S.M., Paerels, F.B.S., Piffaretti, R., Branduardi-
Raymont, G. & Bohringer, H. 2004, A&A, 413, 415
Kaastra, J.S. 1992, An X-Ray Spectral Code for Optically Thin
Plasmas, Internal SRON-Leiden Report, version 2.0.
Kaplinghat, M., Knox, L. & Turner, M.S. 2000, Phys. Rev. Lett., 85,
3335
Kauffman, G., White, S.D.M., & Guiderdoni, B. 1993, MNRAS, 264,
201
ApJ, k522, 82
King, L.J., Clowe, D.I. & Schneider, P. 2002, A&A, 383, 118
Klypin, A.A., Kravtsov, A.V., Valenzuela, O., & Prada, F. 1999,
Lahav, O, Bridle, S.L., Percival, W.J., Peacock, J.A., Efstathiou, G.,
Baugh, C.M., Bland-Hawthorn, J., Bridges, T., Cannon, R., Cole,
S., Colless, M., Collins, C., Couch, W., Dalton, G., de Propris, R.,
Driver, S.P., Ellis, R.S., Frenk, C.S., Glazebrook, K., Jackson, C.,
Lewis, I., Lumsden, S., Maddox, S., Madgwick, D.S., Moody, S.,
Norberg, P., Peterson, B.A., Sutherland, W., & Taylor, K. 2001,
MNRAS, 333, 961
Lewis, A. & Bridle, S. 2002, Phys. Rev. D, 66, 103511
Lewis, A.D., Buote, D.A. & Stocke, J.T. 2002, ApJ, in press (astro-
ph/0209205)
Lewis, A.D., Stocke, J.T. & Buote 2002, ApJL, 573, L13
Liedahl, D.A., Osterheld, A.L. & Goldstein, W.H. 1995, ApJL, 438,
L115
Lloyd-Davies, E.J., Ponman, T.J. & Cannon, D.B., 2000, MNRAS,
315, 689
Loewenstein, M., 2000, ApJ, 532, 17
McGaugh, S.S. & de Blok, W.J.G. 1998, ApJ, 499, 41
Menard, B., Erben, T. & Mellier, Y. 2003, ASP Conf. Ser., 301, xxx
Mewe, R., Gronenschild, E.H.B.M. & van den Oord, G.H.J. 1985,
A&AS, 62, 197
Mewe, R., Lemen, J.R. & van den Oord, G.H.J. 1986, A&AS, 65,
511
Mohapatra, R.N., Nussinov, S. & Teplitz, V.L. 2002, Phys. Rev. D,
66, 063002
Moore, B., Ghigna, S., Governato, F., Lake, G., Quinn, T., Stadel,
J., & Tozzi, P. 1999a, ApJL, 524, L19
Moore, B., Quinn, T., Governato, F., Stadel, J. & Lake, G. 1999,
MNRAS, 310, 1147
Mushotzky, R., Figueroa-Feliciano, E., Loewenstein, M. & Snowden,
S.L. 2003, astro-ph/0302267
Mushotzky, R. F. & Szymkowiak, A.E. 1988,
in Cooling flows
in Clusters and Galaxies, Proceedings of the NATO Advanced
Research Workshop, Dordrecht, Netherlands: Kluwer Academic
Publishers, p. 53
Navarro, J.F., Frenk, C.S. & White, S.D.M. 1997, ApJ, 462, 563
Navarro, J.F., Frenk, C.S. & White, S.D.M. 1997, ApJ, 490, 493
Navarro, J.F. & Steinmetz, M. 2000, ApJ, 528, 607
Neal, R.M. 1993, Technical Report CRG-TR-93-1, University of
Toronto (ftp://ftp.cs.utoronto.ca/pub/radford/review.ps)
Nulsen, P.E.J. & Bo"ohringer, H. 1995, MNRAS, 274, 1093 [NB95]
Papoulis, A. 1984, Probability, Random Variables, and Stochastic
Processes (New York: McGraw Hill)
Peacock, J.A. et al. 2001, Nature, 410, 169
Peterson, J.R., Kahn, S.M., Paerels, F.B.S., Kaastra, J.S., Tamura,
T., Bleeker, J.A.M., Ferrigno, C. & Jernigan, J.G. 2003, ApJ, 590,
207
Pizzolato, F, Molendi, S., Ghizzardi, S. & de Grandi, S. 2003, ApJ,
592, 62
Ponman, T.J., Cannon, T.J. & Navarro, J.F., 1999, Nature, 397, 135
Protassov, R., van Dyk, D.A., Connors, A., Kashyap, V.L. &
Siemiginowska, A. 2002, ApJ, 571, 545
Sand, D.J., Treu, T. & Ellis, R.S. 2002, ApJL, 574, L129
Sarazin, C.L. 1988, X-ray Emission from Clusters of Galaxies,
Cambridge Astrophysics Series, Cambridge: Cambridge University
Press
Schmidt, R.W., Allen, S.W. & Fabian, A.C., 2001, MNRAS, 327,
Shapiro, P.R. & Iliev, I.T., 2000, ApJL, 542, L1
Smail, I., Ellis, R., Ritchett, M.J. & Edge, A.C. 1995, MNRAS, 273,
1070
277
Spergel, D.N. & Steinhardt, P.J. 2000, Phys. Rev. Lett., 84, 17
Sun, M., Jones, C., Murray, S.S., Allen, S.W., Fabian, A.C. & Edge,
A.C. 2003, ApJ, 587, 619
Swaters, R.A., Madore, B.F. & Trewhella, M. 2000, ApJ, 531, L107
Swaters, R.A., Madore, B.F., van den Bosch, F.C. & Balcells, M.
Tyson, J.A., Kochanski, G.P. & Dell'Antonio, I.P. 1998, ApJ, 498,
2003, ApJ, 583, 732
L107
van Dyke, D.A., Connors, A., Kashyap, V.L. & Siemiginowska, A.
2001, ApJ, 548, 243
von Mises, R. 1964, Mathematical Theory of Probability and
Statistics (New York: Academic Press)
Weisskopf, M.C., Brinkman, B., Canizares, C., Garmire, G., Murray,
S. & Van Speybroeck, L.P. 2002, PASP, 114, 1
Wu, X.-P. 2000, MNRAS, 316, 299
Wu, K.K.S., Fabian, A.C. & Nulsen, P.E.J., 2000, MNRAS, 318, 889
Zentner, A.R. & Bullock, J.S. 2002, Phys. Rev. D, 66, 043003
10
PU (x1 ,x2 )
x1
PC (x1 ,x2 )
x1
x2
x2
unconstrained
peak
unconstrained
peak
constrained peak
estimate Pest (M)
constrained peak P0 (M)
Fig. 1. -- Schematic representations of the unconstrained probability distribution function (PU , top), and the constrained distribution (PC ,
bottom) for a two-dimensional parameter space. The estimated peak of the constrained distribution is shown in blue; the true peak is shown
in violet. (For clarity we have not renormalized PC after removing the physically disallowed region of parameter space.)
11
Fig. 2. -- The penalty function g(x) and the associated (approximate) probability distribution function h(x) = e−g(x) for η = 7/2. Here
h(x) nearly, but not precisely, normalized; we have set C = 1 (see Equation 13).
12
p
a r ameter s p
a ce
Ms
Mc
Fig. 3. -- The non-nested relationship between the simple and complex models. The simple model corresponds to the complex model with
one of the normalizations set to 0.
13
Fig. 4. -- Empirical F distributions for models Ms and Mc of five simulated data sets. The cold to hot plasma mass ratio, assuming that
the two are in pressure equilibrium, is shown for each case, along with the significance of the presence of the cool component as determined
through 100 MCMC simulations (see Equation 21). The F value of the original data set is indicated by a vertical dashed line.
14
Fig. 5. -- Baryon density, baryon temperature, spherically enclosed mass, and cylindrically enclosed (projected) mass profiles of A1689, for
Nc = 0 (left panels) and Nc = 2 (right panels). The hot plasma is shown in red, the cool component in blue. Arrows represent points which
lie outside the ordinate range. Reconstructions adhering to constraint equation 6 are shown as solid black curves. Weak (violet) and strong
(green) gravitational lensing measurements are shown for comparison in the bottom panels. (A solid violet line represents an isothermal
sphere fit to the weak lensing data set.) See Table 3 for lensing references.
15
Fig. 6. -- Same as Figure 5, for A1795.
16
Fig. 7. -- Same as Figure 5, for A1835.
17
Fig. 8. -- Same as Figure 5, for A2029.
18
Fig. 9. -- Same as Figure 5, for A2104. Note that there is no reconstruction solution in the Nc = 2 case which is consistent with constraint
equations 5 and 6.
19
Fig. 10. -- Same as Figure 5, for A2204.
20
Fig. 11. -- Same as Figure 5, for HydraA. Note that there is no reconstruction solution in the Nc = 1 case which is consistent with constraint
equations 5 and 6, and that the Hydra A data set did not admit Nc = 2 emission models.
21
Fig. 12. -- Same as Figure 5, for MS1358.
22
Fig. 13. -- Same as Figure 5, for MS2137.
23
Fig. 14. -- Same as Figure 5, for ZW3146.
24
Fig. 15. -- Empirical F distributions and the MCMC significance S of a second cospatial core plasma component for each cluster in the
sample. The F value of the original Chandra data set is denoted with a vertical dashed line. (Note that Hydra A data did not admit an
Nc = 2 emission model.)
25
Fig. 16. -- Mass profiles for models Nc = 0 (red) and Nc = 2 (blue) of MS1358, overlaid for comparison. Weak (violet) and strong (green)
gravitational lensing measurements are shown for comparison.
26
Fig. 17. -- Same as Figure 16, for MS2137.
27
Table 1
Galaxy Cluster
Sample
cluster
z
0.181
A1689
0.0631
A1795
0.2523
A1835
0.0765
A2029
0.1554
A2104
A2204
0.1523
HydraA 0.0522
MS1358
MS2137
ZW3146
0.328
0.313
0.2906
28
Fidelity measures of baryon density (χ2
Table 2
ρ) and temperature (χ2
T ) of each constrained
mass profile (see Equation 12), and multiphase core plasma significance S. (Note
that Hydra A data did not admit an Nc = 2 emission model.)
cluster
χ2
ρ (Nc = 0)
χ2
T (Nc = 0)
χ2
ρ (Nc = 2)
χ2
T (Nc = 2)
S
0.00150
0.0777
0.00229
0.00774
0.00251
0.000994
A1689
A1795
A1835
A2029
A2104
A2204
HydraA 0.197
MS1358
MS2137
ZW3146
0.00663
0.0101
0.0293
0.467
7.45
1.13
1.17
0.624
0.759
2.43
0.381
1.08
1.25
0.00825
0.0156
1.188
0.00773
0.190
0.00947
--
0.00454
0.0110
0.0150
1.14
1.29
1.41
1.20
4.51
0.284
--
0.356
1.38
0.913
0.342
0.823
0.483
0.998
0.661
0.999
--
0.987
0.271
0.989
29
Lensing comparisons for the very relaxed cluster subset.
Table 3
cluster
weak lensing
strong lensing
King, Clowe & Schneider (2002) Wu (2000)
Clowe & Schneider (2002)
Menard, Erben & Mellier (2003) --
A1689
A1835
A2029
MS1358 Hoekstra et al. (1998)
MS2137 --
Allen (1998)
Franx et al. (1997); Allen (1998)
Sand, Treu & Ellis (2002)
|
astro-ph/9701182 | 1 | 9701 | 1997-01-23T16:50:31 | Driftscan Surveys in the 21cm Line with the Arecibo and Nancay Telescopes | [
"astro-ph"
] | Driftscan methods are highly efficient, stable techniques for conducting extragalactic surveys in the 21cm line of neutral hydrogen. Holding the telescope still while the beam scans the sky at the sidereal rate produces exceptionally stable spectral baselines, increased stability for RFI signals, and excellent diagnostic information about system performance. Data can be processed naturally and efficiently by grouping long sequences of spectra into an image format, thereby allowing thousands of individual spectra to be calibrated, inspected and manipulated as a single data structure with standard tools that already exist in astronomical software. The behavior of spectral standing waves (multi-path effects) can be appraised and excised in this environment, making observations possible while the Sun is up. The method is illustrated with survey data from Arecibo and Nancay. | astro-ph | astro-ph |
Driftscan Surveys in the 21cm Line with the Arecibo and
Nan¸cay Telescopes
F. H. Briggs
E. Sorar
Kapteyn Astronomical Institute
University of Pittsburgh
[email protected]
[email protected]
R. C. Kraan-Korteweg
DAEC, Observatoire de Paris, Meudon
[email protected]
W. van Driel
USN, Observatoire de Paris, Meudon
[email protected]
Abstract
Driftscan methods are highly efficient, stable techniques for conducting extragalactic
surveys in the 21cm line of neutral hydrogen. Holding the telescope still while the beam scans
the sky at the sidereal rate produces exceptionally stable spectral baselines, increased stability
for RFI signals, and excellent diagnostic information about system performance. Data can
be processed naturally and efficiently by grouping long sequences of spectra into an image
format, thereby allowing thousands of individual spectra to be calibrated, inspected and
manipulated as a single data structure with standard tools that already exist in astronomical
software. The behavior of spectral standing waves (multi-path effects) can be appraised and
excised in this environment, making observations possible while the Sun is up. The method
is illustrated with survey data from Arecibo and Nan¸cay.
Keywords: methods: data analysis - techniques:
galaxies: distances and redshifts - radio lines: galaxies
image processing - telescopes - surveys -
1 The technique
The designs of the Arecibo and Nan¸cay telescopes make them particularly well suited to taking
data in a driftscan mode. Both have large collecting areas and are therefore sensitive survey
instruments. They obtain the large collecting area by having much of their structure fixed to the
ground. When they are used to track specific celestial coordinates, the on-axis gain changes and
the far out sidelobes move in unpredictable ways, causing spill-over on the ground and RFI to
be time variable. These instabilities increase the level of systematic uncertainties, thus increasing
the difficulty of detecting weak signals. Driftscan observations avoid these problems since all
components of the structures are fixed relative to the ground, thus achieving the full sensitivity of
the large reflecting areas. Similar arguments apply to more conventional radio telescopes, when
1
Figure 1: Nan¸cay Raw Data Image. Passband calibrated spectra have been loaded into the image
in time sequence increasing from right to left. There are two slightly overlapping, 6.4 MHz wide
spectral bands, as marked at the right border. A trace of the continuum as a function of right
ascension is drawn under the image.
Figure 2: Processed Nan¸cay images for three declinations. Spectra are loaded in horizontal lines in
this image, with right ascension increasing upwards. Overlap of the spectral sub-bands has been
removed, and labels on the horizontal axis indicate velocity in km s−1. Eleven detected signals
are marked.
variation in the spillover causes fluctuations in the spectral baselines and when RFI entering the
receiving system is modulated by gain variations in the far sidelobes as the antenna tracks celestial
sources. This report is based on experience obtained in extragalactic HI surveys using the Arecibo
and Nan¸cay Telescopes.
Sorar (1994) and Briggs conducted the Arecibo HI Strip Survey in the driftscan mode in
order to determine the HI-mass function for nearby extragalactic objects by surveying long strips
at constant declination. The observations covered approximately 6000 independent sightlines to
a depth of 7500 km s−1, but since the strips were retraced on many days in order to increase
the integration time on each sightline, nearly a million individual spectra had to be calibrated,
regrouped and averaged. The details of the observational technique were developed by Sorar
(1994), and the results and followup to the survey are summarized by Zwaan et al (1996, these
proceedings).
The method is also in use for surveys at Nan¸cay. One project, being conducted by Kraan-
Korteweg, van Driel, Binggeli, and Briggs, will test the completeness to a depth of 2300 km s−1 of
deep optical catalogs of dwarfs and low surface brightness galaxies in the CnV I cloud (Binggeli
et al 1990). A sample of the raw data from the HI survey are shown in Figure 1. The spectral
passband calibration for this data has been performed by computing the average spectrum of the
entire 3.5 hour dataset, and each spectrum was divided by the average spectrum as it was loaded
in time sequence into the columns of the image. Passage of a telescope beam over a background
continuum source is registered in the image as a dark band. The residual Galactic HI emission
causes the splotchy horizontal band across the image around the HI rest frequency.
Once the data is loaded in image format, images become the units in which the data is stored,
manipulated and displayed. For example, continuum subtraction, averaging of data from different
days and smoothing can be accomplished using procedures in familiar astronomical image process-
ing packages. Figure 2 shows the processed spectra for a total of ∼2500 sightlines in three adjacent
declination strips. The figure results from approximately 82,000 individual spectra. There are 11
detected extragalactic signals resulting from 9 separate galaxies, plus a number of interesting
features associated with Galactic HI.
A second project now in progress at Nan¸cay will observe ∼6% of the sky to a depth of
4500 km s−1 with noise level ∼23 mJy (5σ) for velocity resolution 20 km s−1. The survey is
well matched to detecting nearby examples of gas-rich systems such as HI 1225+01 (Giovanelli &
Haynes 1989, Chengalur, Giovanelli & Haynes 1995) and the circum-galactic ring in Leo (Schneider
1989), as well as detecting a sample of several hundred normal galaxies.
2
Figure 3: Schematic of multipath scattering at the Arecibo Telescope. Representative off-axis
rays, R1, R2, and R3, are seen traversing paths of different length before arriving at the Arecibo
line feed F . In this diagram, R1 takes a direct path to the feed, while R2 and R3 are scattered
from different points (S2 and S3) on the support structure. Radiation is actually scattered in
many directions at the scattering points S2 and S3, but those rays that are scattered downward
parallel to the incoming on-axis rays are directed to be optimally reflected from the main reflector
directly toward the feed.
2 Spectral "standing waves"
Faint periodic fluctuations in the background noise are visible in the low velocity ranges of each
strip in Figure 2. These are commonly called "standing waves" by radio spectroscopists. In these
Nan¸cay data, they occur in the correlator quadrants that have strong Galactic HI emission, but not
in the higher velocity quadrants. The standing waves arise because Galactic HI signal enters the
receiving system by two paths of different length. In an ideal telescope this would not happen, but
it is common in the present design of radio telescopes, since radiation entering the radio receiver
may have been weakly scattered by structure that crosses the telescope's aperture; this additional
scattering can deflect off-axis radiation into the beam of the telescope. When two copies of a signal
take different paths to the receiver, the signal taking the longer path suffers a delay, and when the
autocorrelation function is computed in the spectrometer, a correlation spike is obtained in the
delay channel corresponding to the extra path length. In the Fourier transformation of the ACF
to obtain the power spectrum, a single-channel spike transforms to a sinusoidal variation across
the band. Thus, the path delay corresponds directly to the number of cycles of standing wave
across the spectral band. For the high velocity ranges of Figure 2, there is no interfering signal
entering the system with a time delay and with sufficient strength to generate standing waves,
and therefore the noise characteristics in this redshift range are better behaved.
Any broadband signal entering the receiving system in multiple copies with different delays
can give rise to these multipath effects. A strong source of such radiation is the Sun during
daytime observations, but terrestrially generated broadband RFI, Galactic emission, strong radio
continuum sources and spillover can also do it. The standing waves are often dominated by
a single periodicity, representing, for example, a round trip delay between the feed cabin and
the dish surface. However, a further complication is that antenna structures are often sufficiently
complex that several scatterering locations can contribute (as illustrated for the Arecibo Telescope
in Figure 3), producing signals in several delay channels of the ACF. When the Fourier transform is
computed, the spectral passband produced by the combination of multiple sinusoidal components
can be very complicated. Figure 4 illustrates the complexity of standing wave patterns caused
by the Sun in some Arecibo observations made during a late afternoon observation. Note that
while the feed was positioned at the same antenna coordinates, this same pattern repeated on
successive (solar) days. There was a slow change in the pattern after several days as the Sun
moved in declination.
As well as varying in amplitude, standing waves can drift in phase with time, causing them to
fail to subtract exactly when the simpler forms of passband calibration are applied. This is true
of the processing that has been done to obtain Figure 2, and substantial residuals remain in the
low velocity range. There is a time range in the observations shown in Figure 4 where the drift is
so rapid that the wave moves by a full turn of phase in just a few minutes. Fortunately, there is a
straightforward way to tackle this problem and remove even these complicated patterns from the
driftscan data in an unbiased way. It begins by performing the Fourier transform of the power
3
Figure 4: Arecibo daytime standing waves. Left. The data image for a single circular polarization
after the standard calibration and continuum removal. The image contains 1250 seven second
time steps for a duration of ∼2.4 hours. The bright, spatially resolved galaxy M94 can be seen
midway through the image in the low velocity range. There is a faint linear feature running
vertically in the image at slightly lower velocity -- this is Galactic HI emission that is aliased into
the spectrum through the baseband filters. Right. Harmonic content of each row in the image for
first 63 harmonics. The amplitudes have been averaged by 4 time steps in this image.
spectrum Sti(ν) to obtain the amplitudes and phases of the standing wave components during
each time step ti. These harmonics can be written as complex coefficients An,ti = An,ti exp(iφn,ti),
where n labels the harmonic by the number of full periods of the wave across the spectral band
and ti indicates that the coefficients are expected to change with time, as the spectra are recorded
in discrete time steps. In principle, the Fourier analysis of a string of real numbers, such as these
spectra, produces one half as many complex harmonic coefficients as there were spectral channels
in the power spectrum, but, in practice, the standing waves arise from significant signal only at
low values of n. Thus, the standing wave content of an image-formatted database can be described
by another "image" (or table) of complex numbers, with the same number of time steps, but many
fewer values for An than are required for the number of channels in the spectrum S(ν).
Figure 4 shows the time behavior of the harmonic content of the first 63 harmonics in the
Arecibo data image in the left part of the figure. The standing waves appear with a variety of
periodicities, with several long-lived bursts in the harmonics around n = 20, which corresponds
to differential delays ∼1µsec, the round-trip light travel time between the Arecibo feed support
structure and the surface of the reflector. There is substantial variability depending on where
the scatterer is located on the support structure. There are also some bursts of signal in the
lowest harmonics, possibly due to differential delays between scattered paths such as R3 and R2
in Figure 3, which are both scattered downward into the dish from different points on the support
structure; alternatively, there could be scattering from point S2 directly in the direction of the
feed F , which would also cause a fairly short differential delay and thus a long period standing
wave. The harmonic signature of a bright galaxy can also be seen near the midpoint of the data
set. The faint, periodic, horizontal striping is a result of the slightly variable correlator dump
time, causing the record integration time to beat with the 7 second grid spacing to which the data
was interpolated.
Phase drift of an n harmonic standing wave is described by watching the phase term φn(t)
from An,ti vary with time. Figure 5 shows an example of the amplitude and phase data for a
single harmonic n = 15. The time variation of both An(t) and φn(t) can be efficiently tracked
in time by sliding a window of 16 to 128 time steps along the table of An,ti and then taking the
Fourier transform of the complex time series of each harmonic within the windowed region.
A recipe for tracking and modeling a standing wave is summarized as follows:
(1) Compute the table (or image) of complex harmonic coefficients An,ti
(2) Separate the complex coefficients of the nth harmonic as a function of time into a single long
vector (such as the data plotted in Figure 5).
(3) Subdivide the vector into short enough time spans that the rate of drift is nearly constant over
that time window. The choice of length for the time window depends on how rapidly the rate
of drift changes, since the window must be short enough to track changes in the drift rate but
also long enough to be immune to signals that are localized in a single beam. Of course, a choice
for the window length of 2N time steps (with N equal to an integer in the range 4 to 7) helps to
increase the efficiency in computing the transform.
4
(4) The phase interpolation needed to track the waves is simplified if there is overlap of the
windows, and therefore the window was advanced by either 2N −1 or 2N −2 time steps before recom-
puting the transform. Thus, a mathematical summary is: each window contains p = 2N points
taken from An,ti with ti running from tm to tm+p−1. The Fourier transform of this series of p
numbers produces p complex coefficients an,δk = an,δk exp (iθn,δk). Here an,δk is the strength of
the component drifting at rate δk with phase θn,δk at time tw = tm+p/2. In principle, there may
be well be a range of different standing wave drift rates δ contributing at any given time, since
many locations on the support structure are capable of scattering. In practice, for each window,
we tabulated a δS and a θn,δS corresponding to the δk and θn,δk of the strongest an,δk component
in each time window, after checking for significance relative to the noise level.
(5) Depending of the degree of overlap of the windows, each time step ti falls in either 2 or 4
windows. Each window that produced valid measurements for δS and θn,δS can be used to form
an estimate of the standing wave phase at ti. Thus, a reasonable method for tracking the phase
for the nth harmonic is to compute an estimate < φn,ti > for the phase at each time from the
weighted vector average of the overlapping windows:
ηe(i<φn,ti >) = hX Wn,w(ti − tw)ρe(i(δS,w(ti−tw)+θn,δS ,w)i / X Wn,w(ti − tw)ρ
The sums are taken over the 2 (or 4) windows that overlap at ti. The weighting factors include
Wn,w, which indicates the statistical significance of the solution for the window w, and a factor
(ti − tw)ρ, which gradually transfers the weight among the windows by assigning the greatest
emphasis to the solution whose window is centered closest to ti. The factor ρ is one of many
possible weight adjustments. An amplitude η also results from the vector average; η is close to
unity when there is close agreement between the phase determined by all the windows included
in the average.
(6) Once a solution for phase tracking has been performed (such as shown in the lower panel of
Figure 5), a model for the temporal behavior of the wave amplitudes can be made by smoothing
the time sequence of An,ti exp(i(φn,ti− < φn,ti >)). When the resulting wave amplitudes surpass
a set threshold for significance, they can be stored in a table of harmonic coefficients that can
subsequently be used to generate models for the standing waves as a function of time (as shown
in Figure 6) and then correct the observations by subtracting the standing wave model from the
data image. Note that this technique succeeds at doing little damage to the celestial signals,
since galaxies typically fall in only one beam and the fits are derived from many beams. This
approach is far superior to simply "nulling" the An component, since this would throw away
genuine information from the sky that is specific to each beam.
While our experience at Arecibo and Nancay has shown that the strongest standing waves are
due to the sun, we have also seen similar, but weaker, standing waves at night at Arecibo when
observing in the vicinity of the strong radio source Taurus A (the Crab Nebula); in this case, the
pattern repeated at the same sidereal time day after day. Very strong standing waves have also
been generated by faulty equipment that generated intermittent, broadband noise; these waves
had fixed phase in successive scans, since the source of the rfi was fixed to the Earth and the
antenna was stationary during the observation.
3 Final Comments
The driftscan technique for spectral line surveying offers a number of advantages over pointed
observations: (1) Data quality is maximized since the telescope is still relative to the ground;
spillover, far sidelobes and on-axis gain are constant. (2) System stability can be monitored with
5
e
d
u
t
i
l
p
m
A
]
g
e
d
[
e
s
a
h
P
Figure 5: Amplitude and phase for the n = 15 harmonic standing wave. Top panel. Coefficient
amplitudes plotted as a function of time step (light noisy curve). The heavy solid points are vector
averages of the amplitude (for 20 time steps), computed after application of the phase tracking
algorithm. Heavy points are plotted only when the amplitude surpasses a set threshold. Bottom
panel. Standing wave phases (points). Smooth interpolated curve results from the phase tracking
algorithm. The curve is lighter in regions where the signal to noise ratio for the wave is low.
Figure 6: Standing wave edits. Left panels. Versions of the data image in Figure 4, but with
standing wave models subtracted. The data images are presented on the same grayscale wedge as
Figure 4. Right panels. Standing wave models built from harmonic analysis of the data image in
Figure 4, followed by application of a phase tracking algorithm and reconstruction of a noise-free
images of the standing waves. The upper model includes harmonics n = 4 through 25, when they
are deemed significant. The lower model includes n = 1 through 25. The grayscale is about 3×
more sensitive for the models than for the data images.
6
high precision. (3) The observations are 100 percent efficient since data can be taken continuously.
(4) The telescope scheduling is simple, and observing staff has little real-time responsibility, since
the telescope is sitting still with the brakes on. (5) Spectral passband fluctuations due to standing
waves (or any other instability that varies on time scales longer than a few minutes) can be tracked
and removed, permitting observations to be made during the daytime. (6) The data is naturally
processed in image-format; the analysis is very simple and efficient, and allows large projects to
be tackled using existing software, which is proven and familiar.
Acknowledgments
We are grateful to the telescope staffs at Arecibo and Nan¸cay for their assistance with the obser-
vations. The Nan¸cay radio observatory is operated by the Observatoire de Paris and associated
as USR B704 to the French Centre National de Recherche Scientifique (CNRS), with financial
support of the Region Centre. The Arecibo Observatory is part of the National Astronomy and
Ionosphere Center, which is operated by Cornell University under agreement with the U.S. Na-
tional Science Foundation. The research by RCKK is being supported with an EC-grant. The
Arecibo HI Strip Survey received extensive support through NSF Grant AST 91-19930.
Binggeli, B., Tarenghi, M, & Sandage, A. 1990, A&A, 228, 42
Briggs, F.H. 1996, this volume
Chengalur, J.N., Giovanelli, R., & Haynes, M.P. 1995, AJ, 109, 2415
Giovanelli, R., & Haynes, M.P. 1989, Ap.J., 346, L5
Schneider, S.E. 1989, Ap.J., 343, 94
Sorar, E. 1994, Ph.D. Thesis, University of Pittsburgh
Zwaan, M., Sprayberry, D., & Briggs, F.H. 1996, this volume.
7
This figure "figure1.jpg" is available in "jpg"(cid:10) format from:
http://arxiv.org/ps/astro-ph/9701182v1
This figure "figure2.jpg" is available in "jpg"(cid:10) format from:
http://arxiv.org/ps/astro-ph/9701182v1
This figure "figure3.gif" is available in "gif"(cid:10) format from:
http://arxiv.org/ps/astro-ph/9701182v1
This figure "figure4.jpg" is available in "jpg"(cid:10) format from:
http://arxiv.org/ps/astro-ph/9701182v1
This figure "figure6.jpg" is available in "jpg"(cid:10) format from:
http://arxiv.org/ps/astro-ph/9701182v1
|
astro-ph/0410526 | 1 | 0410 | 2004-10-22T01:03:44 | From the Earliest Seeds to Today's Supermassive Black Holes | [
"astro-ph"
] | I review scenarios for the assembly of supermassive black holes (MBHs) at the center of galaxies that trace their hierarchical build-up far up in the dark matter halo "merger tree". Monte Carlo realizations of the merger hierarchy in a LCDM cosmology, coupled to semi-analytical recipes, are a powerful tool to follow the merger history of halos and the dynamics and growth of the MBHs they host. X-ray photons from miniquasars powered by intermediate-mass "seed" holes may permeate the universe more uniformly than EUV radiation, make the low-density diffuse intergalactic medium warm and weakly ionized prior to the epoch of reionization breakthrough, and set an entropy floor. The spin distribution of MBHs is determined by gas accretion, and is predicted to be heavily skewed towards fast-rotating Kerr holes, to be already in place at early epochs, and not to change significantly below redshift 5. Decaying MBH binaries may shape the innermost central regions of galaxies and should be detected in significant numbers by LISA. | astro-ph | astro-ph |
From the earliest seeds to today's supermassive
black holes
Piero Madau
Department of Astronomy and Astrophysics, University of California, Santa Cruz CA
95064, USA
Abstract. I review scenarios for the assembly of supermassive black holes (MBHs)
at the center of galaxies that trace their hierarchical build-up far up in the dark mat-
ter halo "merger tree". Monte Carlo realizations of the merger hierarchy in a ΛCDM
cosmology, coupled to semi-analytical recipes, are a powerful tool to follow the merger
history of halos and the dynamics and growth of the MBHs they host. X-ray photons
from miniquasars powered by intermediate-mass "seed" holes may permeate the uni-
verse more uniformly than EUV radiation, make the low-density diffuse intergalactic
medium warm and weakly ionized prior to the epoch of reionization breakthrough, and
set an entropy floor. The spin distribution of MBHs is determined by gas accretion,
and is predicted to be heavily skewed towards fast-rotating Kerr holes, already in place
at early epochs, and not to change significantly below redshift 5. Decaying MBH bi-
naries may shape the innermost central regions of galaxies and should be detected in
significant numbers by LISA.
1 Massive black holes and galaxy formation
The strong link observed between the masses of supermassive black holes (MBHs)
at the center of most galaxies and the gravitational potential wells that host
them suggests a fundamental mechanism for assembling black holes and forming
spheroids in galaxy halos. The mBH-σ relation [12] [16] implies a rough pro-
portionality between MBH mass and the mass of the baryonic component of
the bulge. It is not yet understood whether this relation was set in primordial
structures, and consequently how it is maintained throughout cosmic time with
such a small dispersion, or indeed which physical processes established such a
correlation in the first place [48][18][6].
In cosmologies dominated by cold dark matter (CDM) galaxy halos experi-
ence multiple mergers during their lifetime, with those between comparable-mass
systems ("major mergers") expected to result in the formation of elliptical galax-
ies [21]. Simple models in which MBHs are also assumed to grow during major
mergers and to be present in every galaxy at any redshift -- while only a frac-
tion of them is "active" at any given time -- have been shown to explain many
aspects of the observed evolution of quasars [7][8][25]. The coevolution of MBHs
and their host galaxies in hierarchical structure formation scenarios gives origin
to a number of of important questions, most notably:
• Did the first MBHs form in subgalactic units far up in the merger hierarchy,
well before the bulk of the stars observed today? The seeds of the z ∼ 6 quasars
black holes with masses ∼
> a few hundred M⊙ [9][24].
2
Piero Madau
discovered in the Sloan Digital Sky Survey had to appear at very high redshift,
> 10, if they are accreting no faster than the Eddington rate. In hierarchical
z ∼
cosmologies, the ubiquity of MBHs in nearby luminous galaxies can arise even
if only a small fraction of halos harbor MBHs at very high redshift [32].
• How massive were the initial seeds, and is there a population of relic pre-
galactic MBHs lurking in present-day galaxy halos? A clue to these questions
may lie in the numerous population of ultraluminous off-nuclear ("non-AGN")
X-ray sources that have been detected in nearby galaxies [34]. Assuming isotropic
emission, the inferred masses of these "ULXs" may suggest intermediate-mass
• Can coalescing MBH binaries at very high redshift be detected in signif-
icant numbers by the planned Laser Interferometer Space Antenna (LISA)? If
MBHs were common in the past (as implied by the notion that many distant
galaxies harbor active nuclei for a short period of their life), and if their host
galaxies undergo multiple mergers, then MBH binaries will inevitably form in
large numbers during cosmic history. MBH pairs that are able to coalesce in less
than a Hubble time will give origin to the loudest gravitational wave events in
the universe.
• If was first proposed by [10] that the heating of the surrounding stars
by a decaying MBH pair would create a low-density core out of a preexisting
cuspy (e.g. ρ∗ ∝ r−2) stellar density profile. If stellar dynamical processes can
efficiently drive wide MBH binaries to the gravitational wave (GW) emission
stage, what is the cumulative dynamical effect of multiple black hole mergers on
galaxy stellar cusps?
ionized at z ∼
• Active galactic nuclei powered by supermassive holes keep the universe
< 4, structure the intergalactic medium (IGM), and probably regu-
late star formation in their host galaxies. Intermediate-mass holes accreting gas
from the surrounding medium may shine as "miniquasars" at redshifts as high
as z ∼ 20. What is the thermodynamic effect of miniquasars on the IGM at early
times?
• Besides their masses, astrophysical black holes are completely characterized
by their spins, S = aGmBH/c, 0 ≤ a/mBH ≤ 1. The spin of a MBH is expected to
have a significant effect on its observational manifestation, such as the efficiency
of converting accreted mass into radiation and the existence and direction of jets
in active nuclei. What is the expected distribution of MBH spins and how does
this evolve with cosmic time?
In this talk I will review some recent developments in our understanding of
the assembly, growth, emission history, and environmental impact of MBHs from
early epochs to the present. Unless otherwise stated, all the results shown below
refer to the currently favoured (by a variety of observations) ΛCDM world model
with ΩM = 0.3, ΩΛ = 0.7, h = 0.7, Ωb = 0.045, σ8 = 0.93, and n = 1.
Earliest seed MBHs
3
2 MBHs as Population III remnants
The first stars in the universe must have formed out of metal-free gas, in dark
> 5 × 105 M⊙ [14] condensing from the high-σ
matter "minihalos" of total mass ∼
peaks of the primordial density field at redshift z = 20 − 30. Numerical sim-
ulations of the fragmentation of primordial clouds in standard CDM theories
all show the formation of Jeans unstable clumps with masses exceeding a few
hundred solar masses; because of the slow subsonic contraction -- a regime set
up by the main gas coolant, molecular hydrogen -- further fragmentation into
sub-components is not seen, and a single very massive star forms from the inside
out [5][1][41].
Fig. 1. Mass function of minihalos of mass M formed at z = 15, 20, 25 which, by
the later time z0, will have merged into a more massive halo of total mass M0. Solid
curves: z0 = 0.8, M0 = 1012 h−1 M⊙ ("Milky Way" halo). Dashed curves: z0 = 3.5,
M0 = 2 × 1011 h−1 M⊙ (older "bulge"). If only one seed hole formed in each ∼ 106 M⊙
minihalo collapsing at z ∼ 20 (and triple hole interactions and binary coalescences were
neglected), several thousands relic IMBHs and their descendants would be orbiting
throughout present-day galaxy halos [29][23].
At zero metallicity mass loss through radiatively-driven stellar winds or
nuclear-powered stellar pulsations is expected to be negligible, and Population
III stars will likely die losing only a small fraction of their mass (except for
100 < m∗ < 140 M⊙). Nonrotating very massive stars in the mass window
< 260 M⊙ will disappear as pair-instability supernovae [4], leaving
no compact remnants and polluting the universe with the first heavy elements
[44][40]. Stars with 40 < m∗ < 140 M⊙ and m∗ > 260 M⊙ are predicted in-
stead to collapse to black holes with masses exceeding half of the initial stellar
< m∗ ∼
140 ∼
4
Piero Madau
mass [20]. Barring any fine tuning of the initial mass function of Pop III stars,
intermediate-mass black holes (IMBHs) -- with masses above the 4 -- 18 M⊙ range
of known "stellar-mass" holes -- may then be the inevitable endproduct of the
first episodes of pregalactic star formation [29]. Since they form in high-σ rare
density peaks, relic IMBHs are expected to cluster in the bulges of present-day
galaxies as they become incorporated through a series of mergers into larger
and larger systems (see Fig. 1). The presence of a small cluster of IMBHs in
galaxy nuclei may have several interesting consequences associated with tidal
captures of ordinary stars (likely followed by disruption), capture by the central
supermassive hole, and gravitational wave radiation from such coalescences [33].
Accreting pregalactic IMBHs may be detectable as ultra-luminous, off-nuclear
X-ray sources [29][26].
3 The first miniquasars
Physical conditions in the central potential wells of young and gas-rich proto-
galaxies may have been propitious for black hole gas accretion. Perhaps seed
black holes grew efficiently in small minihalos just above the cosmological Jeans
mass (with shallow potential wells), or maybe gas accretion had to await the
buildup of more massive galaxies (with virial temperatures above the thresh-
old for atomic cooling). This issue is important for the detectability of high-z
miniquasars: it also determines whether the radiation background at very high
redshifts had an X-ray component component able to preheat and partially ionize
the IGM.
As mentioned above, gas condensation in the first baryonic objects is possible
through the formation of H2 molecules, which cool via roto-vibrational transi-
tions down to temperatures of a few hundred kelvins. In the absence of a UV
photodissociating flux and of ionizing X-ray radiation, three-dimensional sim-
ulations of early structure formation show that the fraction of cold, dense gas
available for accretion onto seed holes or star formation exceeds 20% for halos
more massive than 106 M⊙ [28]. On th eother hand, a zero-metallicity progenitor
star in the range 40 < m∗ < 500 M⊙ emits about 70,000 photons above 1 ryd per
stellar baryon [43]. The ensuing ionization front will completely overrun the host
halo, photoevaporating most of the surrounding gas [54]. Black hole remnants
of the first stars that created H II regions are then unlikely to accrete significant
mass until new cold material will be made available through the hierarchical
merging of many gaseous subunits.
Accretion onto IMBHs may be an attractive way to (partially) reionize the
low-density IGM [30][42]. A large fraction of the UV radiation from massive
stars may not escape the dense sites of star formation, or may be deposited
locally in halo gas that recombines almost immediately. The harder radiation
emitted from miniquasars is instead more likely to escape from the hosts into
intergalactic space, and may then produce more 'durable' (albeit partial) ion-
ization in the diffuse IGM. High-resolution hydrodynamics simulations of early
structure formation in ΛCDM cosmologies are a powerful tool to track in de-
Earliest seed MBHs
5
tail the thermal and ionization history of a clumpy IGM and guide studies of
early reheating. We [27] have used Enzo, an adaptive mesh refinement (AMR),
grid-based hybrid (hydro+N-body) code developed by Bryan & Norman (see
http://cosmos.ucsd.edu/enzo/) to solve the cosmological hydrodynamics equa-
tions and simulate the effect of a miniquasar turning on at very high redshift in
a volume 1 Mpc on a side (comoving). We first identify in a low-resolution pure
N-body simulation the Lagrangian volume of a resolved protogalactic halo with
a total mass 7× 105 M⊙ at z = 25, above the cosmological Jeans mass. We then
generate new initial conditions with an 1283 initial static grid that covers a 0.5
Mpc volume centered around the identified high-σ peak. During the evolution,
refined grids (for a maximum of 5 additional levels) are introduced with twice
the spatial resolution of the parent (coarser) grid to home in, with progressively
finer resolution, on the densest parts of the "cosmic web". The simulation follows
the non-equilibrium chemistry of the dominant nine species (H, H+, H−, e, He,
He+, He++, H2, and H+
2 ) in primordial gas, and includes radiative losses from
atomic and molecular line cooling.
Fig. 2. Gas distribution (at overdensity 2) in the inner 0.5 Mpc of the simulation
box at z = 15.5. Dark matter halos of different masses (identified with a halo-finder
algorithm) are plotted as dark spheres.
At z = 21, a miniquasar powered by a 150 M⊙ black hole accreting at the
Eddington rate is turned on in the protogalactic halo. The miniquasar shines
for a few Salpeter times (i.e. down to z ∼ 15) and is a copious source of soft
X-ray photons, which permeate the IGM more uniformly than possible with ex-
treme ultraviolet (EUV, ≥ 13.6 eV) radiation [38] and make it warm and weakly
ionized prior to the epoch of reionization breakthrough [50]. A spectrum with
ν Lν = const (like the nonthermal component observed in ULXs) was assumed
for photons with energies in the range 0.2-10 keV, to which the simulation box is
6
Piero Madau
transparent. X-rays alone do not produce a fully ionized medium, but partially
photoionize the gas by repeated secondary ionizations. A primary nonthermal
photoelectron of energy E = 1 keV in a medium with residual ionization (from
the recombination epoch) x = 2 × 10−4 will create over two dozens secondary
electrons, depositing a fraction fion ≈ 37% of its initial energy as secondary
ionizations of hydrogen, and only fheat ≈ 13% as heat [47]. The timescale for
electron-electron encounters resulting in a fractional energy loss f = ∆E/E,
tee ≈ 140 yr Ef (cid:18) 1 + z
20 (cid:19)−3 (cid:18) ln Λ
20 (cid:19)−1
x−1
(1)
(where E is measured in keV), is typically much shorter that the electron
Compton cooling timescale off cosmic microwave background (CMB) photons,
tC = (7 × 106 yr) [(1 + z)/20]−4, and thus the primary photoelectron will ionize
and heat the surrounding medium before it is cooled by the CMB. Once the
IGM ionized fraction increases to x ≈ 0.1, the number of secondary ionizations
per ionizing photon drops to a few, and the bulk of the primary's energy goes
into heat (fheat ≈ 0.6) via elastic Coulomb collisions with thermal electrons.
Fig. 3. Mass (upper curves) and volume (lower curves) weighted ]gas fraction vs. gas
temperature in the simulation box. A miniquasar was turned on at z = 21.
Figure 3 shows the mass and volume-weighted gas fraction in the simulation
box vs. temperature as four different redshifts. The heating effect of X-rays from
the miniquasar is clearly seen, with gas temperatures between 103 and 104 K at
z < 20. Strong Jeans mass filtering takes place, and subsequent minihalos will
no longer be able to accrete gas due to the smoothing effect of gas pressure. The
low-density IGM acquires a uniform "entropy floor" [39] that (a) greatly reduces
gas clumping, curtailing the number of photons needed to maintain reionization,
and (b) results in significantly lower gas densities in the cores of minihalos that
suppress rapid H2 formation. The latter effect may imply that X-rays inhibit
rather than enhance star formation [19][28].
Earliest seed MBHs
7
4 MBH binaries and galaxy cores
Frequent galaxy mergers will inevitably lead to the formation of MBH binaries.
As dark matter halos assemble, MBHs get incorporated into larger and larger
halos, sink to the center owing to dynamical friction, accrete a fraction of the
gas in the merger remnant to become supermassive, form a binary system, and
eventually coalesce [3]. In a stellar background a "hard" binary shrinks by cap-
turing the stars that pass close to the holes and ejecting them at much higher
velocities, a super-elastic scattering process that depletes the nuclear region and
turns a stellar "cusp" into a low-density core. Rapid coalescence eventually en-
sues due to the emission of gravitational radiation. Observationally, there is clear
evidence in early-type galaxies for a systematic trend in the distribution of sur-
face brightness profiles, with faint ellipticals showing steep power-law profiles
(cusps), while bright ellipticals have much shallower stellar cores [11]. Detailed
N-body simulations have confirmed the cusp-disruption effect of a hardening
MBH binary [35], but have shed little light on why bright ellipticals have lower
central concentrations than faint ellipticals.
Fig. 4. Mass deficit produced at z = 0 by shrinking MBHs in our merger tree as a func-
tion of nuclear MBH mass (filled dots). Left panel: cusp regeneration case. Right panel:
core preservation case. Galaxies without a core (i.e. those that have never experienced
a MBH-MBH merger or with their cusp recently regenerated) are shown as vertical
arrows at an arbitrary mass deficit of 106 M⊙. Empty squares: mass deficit inferred in
a sample of galaxies by [36]. Empty circles: same for the "core" galaxies of [11]. Empty
triangles: same for the "cuspy" galaxies of [11], assuming a flat core within the upper
limit on the core size. (From [52].)
8
Piero Madau
shrink the binary down to a separation ∼
The role of MBH binaries in shaping the central structure of galaxies is best
understood within the framework of a detailed model for the hierarchical assem-
bly of MBHs over cosmic history [51]. Stellar cusps can be efficiently destroyed
over cosmic time by decaying binaries if stellar dynamical processes are able to
< 10% of the separation at which the
binary becomes hard. More massive halos have more massive nuclear holes and
experience more merging events than less massive galaxies: hence they suffer
more from the eroding action of binary MBHs and have larger cores. In [52] we
found that a model in which the effect of the hierarchy of MBH interactions is cu-
mulative and cores are preserved during galaxy mergers produces at the present
epoch a correlation between the "mass deficit" (the mass needed to bring a flat
inner density profile to a r−2 cusp) and the mass of the nuclear MBH, with a
normalization and slope comparable to the observed relation (see Fig. 4). Mod-
els in which the mass displaced by the MBH binary is replenished after every
major galaxy merger appear instead to underestimate the mass deficit observed
in "core" galaxies. In [52] a simple scheme was applied to hardening pairs, in
which the "loss cone" is constantly refilled and a constant density core forms due
to the ejection of stellar mass. The effect of loss-cone depletion (the depletion
of low-angular momentum stars that get close enough to extract energy from a
hard binary) is one of the major uncertainties in computing the decay timescale,
and makes it difficult to construct detailed scenarios for coalescing black hole
binaries.
Fig. 5. Number of MBH binary coalescences observed per year at z = 0, per unit red-
shift, in different mBH = M1 + M2 mass intervals. Each panel also lists the integrated
event rate, dN/dt, predicted by [45]. The rates (solid lines) are compared to a case in
which triple black hole interactions are switched off (dotted lines). Triple hole interac-
tions increase the coalescence rate at very high redshifts, while, for 10 < z < 15, the
rate is decreased because of the reduced number of surviving binaries. Dashed lines:
rates computed assuming binary hardening is instantaneous, i.e. MBHs coalesce after
a dynamical friction timescale.
Earliest seed MBHs
9
5 Gravitational radiation from inspiraling MBH binaries
MBH binaries, with masses in the range 103 − 107 M⊙, are one of the primary
target for LISA [17][55][45]. Interferometers operate as all-sky monitors, and the
data streams collect the contributions from a large number of sources belonging
to different cosmic populations. To optimize the subtraction of resolved sources
from the data stream, it is important to have a detailed description of the ex-
pected rate, duration, amplitude, and waveforms of events. Figure 5 shows the
number of MBH binary coalescences per unit redshift per unit observed year
predicted by [45], using a detailed model of MBH binaries dynamics. The ob-
served event rate is obtained by dividing the rate per unit proper time by the
(1 + z) cosmological time dilation factor. Each panel shows the rate for differ-
ent mBH = M1 + M2 mass intervals, and lists the integrated event rate, dN/dt,
across the entire sky. The number of events per observed year per unit redshift
peaks at z = 2 for 107 < mBH < 109 M⊙, at z = 3− 4 for 105 < mBH < 107 M⊙,
and at z = 10 for mBH < 105 M⊙, i.e. the lower the black hole mass, the higher
the peak redshift. Beyond the peak, the event rate decreases steeply with cosmic
time.
In the stationary case, i.e., assuming no orbital decay, the GW emission
spectrum of a MBH pair in a circular orbit of radius a is a delta function at
rest-frame frequency fr = ω/π, where ω = pG(M1 + M2)/a3 is the Keplerian
angular frequency of the binary. Orbital decay due to GW emission results in a
shift of the emitted frequency to increasingly larger values as the binary evolution
proceeds. Typically, the timescale for frequency shift is long compared to the
wave period, and short compared to the duration of the observation. Only close
to the innermost stable circular orbit (ISCO), the GW frequency changes at a
rate comparable to the frequency itself. The rest-frame energy flux (energy per
unit area per unit time) associated to the GW is
dE
dAdt
=
π
4
c3
G
r h2,
f 2
(2)
where the strain amplitude (sky-and-polarisation averaged) at comoving distance
r(z) is
h =
8π2/3
101/2
G5/3M5/3
c4r(z)
f 2/3
r
,
(3)
M is the "chirp mass" of the binary, and all the other symbols have their stan-
dard meaning. The strain is averaged over a wave period. The important quan-
tity to consider is the number of cycles n spent in a frequency interval ∆f ≃ f
around a given frequency f . In general, n = f 2/ f ∝ f −5/3. For a periodic signal
10
Piero Madau
at frequency f lasting for a time interval longer than the observation time τ , we
have simply n = f τ . The characteristic strain in an observation of (observed)
duration τ is then
where f = fr/(1 + z) is the observed frequency. In Figure 6 hc is plotted for dif-
ferent MBH binaries at different redshifts, compared to the LISA hrms multiplied
by a factor of 5, assuming a 3-year observation.
and
r
,
hc = h√n ∝ f −1/6
hc = hpf τ ∝ f 7/6
r
n < f τ,
,
n > f τ,
(4)
(5)
Fig. 6. Characteristic strain hc for MBH binaries with different masses and redshifts.
From top to bottom, the first three curves refer to systems with log(M1/ M⊙) = 7, 6, 5,
respectively, and M2 = 0.1M1. The solid, long-dashed, and short-dashed lines assumes
the binary at z = 1, 3, 5, respectively. A 3-year observation is considered. The lowest
solid curve assumes an equal mass binary M1 = M2 = 103 M⊙ at z = 7. The small
diamonds on each curve mark, from left to right, the observed frequency at 1 year, 1
month and 1 day before coalescence. The thick curve is LISA 5hrms, approximatively
the threshold for detection with S/N ≥ 5. (From [46]).
At frequencies higher than the "knee", the time spent around a given fre-
quency is less than 3 years, and hc ∝ f −1/6. The signal shifts toward higher
frequency during the observation, and reaches the ISCO and the coalescence
phase in most cases. The lowest curve represents a low mass, high redshift equal
mass binary. In terms of their detectability by LISA, they represent a somewhat
different class of events. Contrary to the case of more massive binaries present
at lower z, the final coalescence phase of light binaries lies at too high freque-
cies, well below the LISA threshold. For frequencies much below the knee, the
characteristic strain is proportional to f 7/6, as the timescale for frequency shift
Earliest seed MBHs
11
is longer than 3 years. The signal amplitude is then limited by the observation
time, not by the intrinsic properties of the source. The source will be observed
as a "stationary source", a quasi-monochromatic wave for the whole duration
of the observation. An increase in the observation time will result in a shift of
the knee toward lower frequencies. The time needed for the sources to reach the
ISCO starting from the knee frequency is, approximatively, the observing time.
As recently shown by [46], the GW signal from MBH binaries will be resolved
(assuming a 3-year LISA observation) into ∼ 100 discrete events, 40 of which will
be observed above threshold until coalescence. These "merging events" involve
< 6.
relatively massive binaries, M ∼ 105 M⊙, in the redshift range 2 ∼
The remaining ∼ 60 events come from higher redshift, less massive binaries
> 6) and, although their S/N integrated over the duration
(M ∼ 5×103 M⊙ at z ∼
of the observation can be substantial, the final coalescence phase is at too high
frequency to be directly observable by LISA. The total number of detected events
> 90% of all coalescences of massive black hole binaries at
accounts for a fraction ∼
< 5. The residual confusion noise from unresolved massive black hole binaries
z ∼
is expected to be at least an order of magnitude below the estimated stochastic
LISA noise.
< z ∼
6 MBH spins
The spin of a MBH is determined by the competition between a number of
physical processes. Black holes forming from the gravitational collapse of very
massive stars endowed with rotation will in general be born with non-zero spin
[13]. An initially non-rotating hole that increases its mass by (say) 50% by swal-
lowing material from an accretion disk may be spun up to a/mBH = 0.84 [2].
While the coalescence of two non-spinning black holes of comparable mass will
> 0.8 [15],
the capture of smaller companions in randomly-oriented orbits may spin down
a Kerr hole instead [22].
immediately drive the spin parameter of the merged hole to a/mBH ∼
12
Piero Madau
Fig. 7. Normalized distribution of mass ratios, q = M2/M1, of coalescing MBH binaries
at three different epochs. Note that at low redshift MBHs typically capture much
smaller companions.
In [53] we have made a first attempt at estimating the distribution of MBH
spins and its evolution with cosmic time in the context of hierarchical structure
formation theories, following the combined effects of black hole-black hole coa-
lescences and accretion from a gaseous disk on the magnitude and orientation of
MBH spins. Here I will briefly summarize our findings. Binary coalescences ap-
pear to cause no significant systematic spin-up or spin-down of MBHs: because
of the relatively flat distribution of MBH binary mass ratios in hierarchical mod-
els (shown in Fig. 7), the holes random-walk around the spin parameter they
are endowed with at birth, and the spin distribution retains significant memory
of the initial rotation of "seed" holes.
It is accretion, not binary coalescences, that dominates the spin evolution
of MBHs (Fig. 8). Accretion can lead to efficient spin-up of MBHs even if the
angular momentum of the inflowing material varies in time. This is because, for
a thin accretion disk, the hole is aligned with the outer disk on a timescale that
is much shorter than the Salpeter time [37], leading to accretion via prograde
equatorial orbits. As a result, most of the mass accreted by the hole acts to spin
it up, even if the orientation of the spin axis changes in time. For a geometrically
thick disk, alignment of the hole with the outer disk is much less efficient, occur-
ring on a timescale comparable to the Salpeter time. Even in this case most holes
will be rotating rapidly. This is because, in any model in which MBH growth
is triggered by major mergers, every accretion episode must typically increase a
hole's mass by about one e-folding to account for the local MBH mass density
and the mBH − σ∗ relation. Most individual accretion episodes thus produce
rapidly-rotating holes independent of the initial spin.
Earliest seed MBHs
13
Fig. 8. Distribution of MBH spins in different redshift intervals. Left panel: effect of
black hole binary coalescences only. Solid histogram: seed holes are born with a ≡
a/mBH = 0.6. Dashed histogram: seed holes are born non-spinning. Right panel: spin
distribution from binary coalescences and gas accretion. Seed holes are born with a =
0.6, and are efficiently spun up by accretion via a thin disk.
Under the combined effects of accretion and binary coalescences, we find
that the spin distribution is heavily skewed towards fast-rotating Kerr holes,
is already in place at early epochs, and does not change significantly below
redshift 5. As shown in Figure 9, about 70% of all MBHs are maximally rotating
and have mass-to-energy conversion efficiencies approaching 30%. Note that if
the equilibrium spin attained by accreting MBHs is lower than the value of
a ≡ a/mBH = 0.998 used here, as in the thick disk MHD simulations of [15],
where a ≈ 0.93, then the accretion efficiency will be lower as well, ≈ 17%.
14
Piero Madau
Fig. 9. Distribution of accretion efficiencies, ǫ ≡ 1 − E, in different redshift intervals
(assuming that the energy radiated is the binding energy at the ISCO). The spin
distribution from binary coalescences and gas accretion has been calculated assuming
the holes accrete via a thin disk on prograde equatorial orbits.
Even in the conservative case where accretion is via a geometrically thick
disk (and hence the spin/disk alignment is relatively inefficient) and the initial
orientation between the hole's spin and the disk rotation axis is assumed to be
random, we find that most MBHs rotate rapidly with spin parameters a > 0.8
and accretion efficiencies ǫ >12%. As recently shown by [56][31], a direct com-
parison between the local MBH mass density and the mass density accreted by
luminous quasars shows that quasars have a mass-to-energy conversion efficiency
> 0.1 (a simple and elegant argument originally provided by [49]). This high
ǫ ∼
average accretion efficiency may suggest rapidly rotating Kerr holes, in agree-
ment with our findings. Since most holes rotate rapidly at all epochs, our results
suggest that spin is not a necessary and sufficient condition for producing a
radio-loud quasar.
I would like to thank my numerous collaborators on this subject: F. Haardt,
M. Kuhlen, P. Oh, E. Quataert, M. Rees, A. Sesana, and M. Volonteri. This
manuscript was written while the author was enjoying the hospitality of the Kavli
Institute for Theoretical Physics. Support for this work was provided by NASA
grant NNG04GK85G, and by NSF grants AST-0205738 and PHY99-07949.
References
1. Abel, T., Bryan, G., & Norman, M. 2000, ApJ, 540, 39
Earliest seed MBHs
15
2. Bardeen, J. M. 1970, Nature, 226, 64
3. Begelman, M. C., Blandford, R. D., & Rees, M. J. 1980, Nature, 287, 307
4. Bond, J. R., Arnett, W. D., & Carr, B. J. 1984, ApJ, 280, 825
5. Bromm, V., Coppi, P. S., & Larson, R. B. 1999, ApJ, 527, L5
6. Burkert, A., & Silk, J. 2001, ApJ, 554, L151
7. Cattaneo, A., Haehnelt, M. G., & Rees, M. J. 1999, MNRAS, 308, 77
8. Cavaliere, A., & Vittorini, V. 2000, ApJ, 543, 599
9. Colbert, E. J. M., & Mushotzky, R. F. 1999, ApJ, 519, 89
10. Ebisuzaki, T., Makino J., & Okumura S. K. 1991, Nature, 354, 212
11. Faber, S. M., et al. 1997, AJ, 114, 1771
12. Ferrarese, L., & Merritt, D. 2000, ApJ, 539, L9
13. Fryer, C. L., Woosley, S. E., & Heger, A. 2001, ApJ, 550, 372
14. Fuller, T. M., & Couchman, H. M. P. 2000, ApJ, 544, 6
15. Gammie, C. F., Shapiro, S. L., & McKinney, J. C. 2004, ApJ, 602, 312
16. Gebhardt, K., et al. 2000, ApJ, 543, L5
17. Haehnelt, M.G. 1994, MNRAS, 269, 199
18. Haehnelt, M. G., & Kauffmann, G. 2000, MNRAS, 318, L35
19. Haiman, Z., Abel, T., & Rees, M. J. 2000, ApJ, 534, 11
20. Heger, A., & Woosley, S. E. 2002, ApJ, 567, 532
21. Hernquist, L. 1992, ApJ, 400, 460
22. Hughes, S. A., & Blandford, R. D. 2003, ApJ, 585, L101
23. Islam, R. R., Taylor, J. E., & Silk, J. 2003, MNRAS, 340, 647
24. Kaaret, P., et al. 2001, MNRAS, 321, L29
25. Kauffmann, G., & Haehnelt, M. G. 2000, MNRAS 311, 576
26. Krolik, J. H. 2004, ApJ, in press (astro-ph/0407285)
27. Kuhlen, M., & Madau, P. 2004, in prepration
28. Machacek, M. M., Bryan, G. L., & Abel, T. 2003, MNRAS, 338, 273
29. Madau, P., & Rees, M. J. 2001, ApJ, 551, L27
30. Madau, P., Rees, M. J., Volonteri, M., Haardt, F., & Oh, S. P. 2004, ApJ, 606, 484
31. Marconi, A., Risaliti, G., Gilli, R., Hunt, L. K., Maiolino, R., & Salvati, M. 2004,
MNRAS, 351, 169
32. Menou, K., Haiman, Z., & Narayanan, V. K. 2001, ApJ, 558, 535
33. Miller, M. C. 2004, preprint (astro-ph/0409331)
34. Miller, M. C., & Colbert, E. J. M. 2004, Int. J. Mod. Phys. D13, 1
35. Milosavljevic, M., & Merritt, D. 2001, ApJ, 563, 34
36. Milosavljevic, M., Merritt, D., Rest, A., van den Bosch, F. C. 2002, MNRAS, 331,
L51
37. Natarajan, P., & Pringle, J. E. 1998, ApJ, 506, L97
38. Oh, P. S. 2001, ApJ, 553, 499
39. Oh, S. P., & Haiman, Z. 2003, MNRAS, 346, 456
40. Oh, S. P., Nollett, K. M., Madau, P., & Wasserburg, G. J. 2001, ApJ, 562, L1
41. Omukai, K., & Palla, F. 2003, ApJ, 589, 677
42. Ricotti, M., & Ostriker, J. P. 2004, MNRAS, 352, 547
43. Schaerer, D. 2002, A&A, 382, 28
44. Schneider, R., Ferrara, A., Natarajan, P., & Omukai, K. 2002, ApJ, 571, 30
45. Sesana, A., Haardt, F., Madau, P., & Volonteri, M. 2004, ApJ, 611, 623
46. Sesana, A., Haardt, F., Madau, P., & Volonteri, M. 2004, ApJ, submitted
(astro-ph/0409255)
47. Shull, J. M., & Van Steenberg, M. E. 1985, ApJ, 298, 268
48. Silk, J., & Rees, M. J. 1998, A&A, 331, L1
16
Piero Madau
49. Soltan, A. 1982, MNRAS, 200, 115
50. Venkatesan, A., Giroux, M. L., & Shull, M. J. 2001, ApJ, 563, 1
51. Volonteri, M., Haardt, F., & Madau, P. 2003, ApJ, 582, 559
52. Volonteri, M., Madau, P., & Haardt, F. 2003, ApJ, 593, 661
53. Volonteri, M., Madau, P., Quataert, E., & Rees, M. J. 2004, ApJ,
in press
(astro-ph/0410342)
54. Whalen, D., Abel, T., & Norman, M. L. 2004, ApJ, 610, 22
55. Wyithe, J.S.B., & Loeb, A. 2003 ApJ, 590, 691
56. Yu, Q., & Tremaine, S. 2002, MNRAS, 335, 965
|
astro-ph/0309557 | 1 | 0309 | 2003-09-20T15:14:29 | The Radio Afterglows of Gamma-Ray Bursts | [
"astro-ph"
] | Radio afterglow studies have become an integral part of the study of gamma-ray bursts, providing complementary and sometimes unique diagnostics on GRB explosions, their progenitors, and their environments. This brief review consists of two parts. The first section is a summary of current search strategies and the main observational properties of radio afterglows. In the second section we highlight the key scientific contributions made by radio observations, either alone or as part of panchromatic studies. | astro-ph | astro-ph |
The Radio Afterglows of Gamma-Ray Bursts
Dale A. Frail
National Radio Astronomy Observatory, Socorro, NM 87801 USA
1 Introduction
Our understanding of the gamma-ray bursts (GRBs) has advanced rapidly
since the discovery of long-lived "afterglow" emission from these events. Ra-
dio afterglow studies have become an integral part of this field, providing
complementary and sometimes unique diagnostics on GRB explosions, their
progenitors, and their environments. The reason for this is that the radio part
of the spectrum is phenomenologically rich. This can be illustrated simply by
calculating the brightness temperature (Tb ∝ Fν/(θs ν)2) for a 1 mJy cen-
timeter wavelength source at cosmological distances (∼ 1028 cm), expanding
with Vexp ≤ c one week after the burst. Since the derived Tb ∼ 1013 K is
well in excess of the TIC ∼ 1011 − 1012 K limit imposed by inverse Compton
cooling, it follows, independent of any specific afterglow model, that the radio
emission must originate from a compact, synchrotron-emitting source that is
expanding superluminally (i.e. Tb ∼ Γ × TIC, Γ >> 1). Likewise, since the
brightness temperature cannot exceed the mean kinetic energy of the elec-
trons, the emission is expected to be self-absorbed at longer wavelengths [1].
Finally, strong modulation of the centimeter signal is expected on timescales
of hours and days because the angular size θs of this superluminal source is
comparable to the Fresnel angle of the turbulent ionized gas in our Galaxy [2].
Synchrotron self-absorption, interstellar scintillation, forward shocks, reverse
shocks, jet-breaks, non-relativistic transitions and obscured star formation are
among the phenomena routinely observed.
This short review is divided into two parts. The first section (§2) is a sum-
mary of the current search strategies and the main observational properties of
radio afterglows. In the second section (§3) we highlight the key scientific con-
tributions made by radio observations, either alone or as part of panchromatic
studies. By necessity we will restrict this brief review to long-duration GRBs,
although radio afterglows have also been detected toward the newly classified
X-ray flashes, and searches have been carried out toward short bursts [3].
2
Dale A. Frail
2 Detection Statistics and Observational Properties
The search for a radio afterglow is initiated either by a satellite localization of
the burst, or by the detection of the X-ray or optical afterglow. The current
search strategy has been to use the Very Large Array (VLA)1 or the Australia
Telescope Compact Array (ATCA; for declinations, δ < −40◦)2 at 5 GHz or
8.5 GHz. These frequencies were chosen as a compromise between the need to
image the typical error box size of 30-100 arcmin2, while having the requisite
sensitivity to detect afterglows at sub-milliJansky levels. At lower frequencies
the afterglow is attenuated by synchrotron self-absorption (fν ∝ ν 2), while at
higher frequencies the field-of-view is proportionally smaller (FOV∝ ν 2). For
typical integration times (10 min at the VLA, and 240 min at the ATCA) the
rms (receiver) noise is 30-50 µJy. Follow-up observations of detected after-
glows were carried out by a network of radio facilities at centimeter, millimeter
and submillimeter wavelengths [4].
s
w
o
g
r
e
l
t
f
A
f
o
r
e
b
m
u
N
15
10
5
0
s
w
o
g
r
e
l
t
f
A
f
o
r
e
b
m
u
N
6
5
4
3
2
1
100
Flux Density (µJy)
1000
29
30
Radio Luminosity (erg/s/Hz)
31
32
33
Fig. 1. (Left) Histogram distribution of flux densities (or upper limits) at 8.5 GHz
for a complete sample of bursts. The hatched histogram shows the distribution for
the detections only. (Right) Histograms of radio luminosity from the same sample
but restricted to the subset of bursts with known redshifts. The hatched histogram
shows the distribution for bursts with detected radio afterglows only.
In the five year period beginning in 1997 and ending in 2001 approximately
1500 radio flux density measurements (or upper limits) were made toward 75
bursts [5]. From these 75 GRBs, there are a total of 32/36 successful X-
ray searches, 27/70 successful optical searches, and 25/75 successful searches.
These afterglow search statistics illustrate a well-known result, namely that
the detection probability for X-ray afterglows is near unity, while for optical
1 The NRAO is a facility of the National Science Foundation operated under co-
operative agreement by Associated Universities, Inc.
2 The Australia Telescope is funded by the Commonwealth of Australia for opera-
tion as a National Facility managed by CSIRO.
The Radio Afterglows of Gamma-Ray Bursts
3
afterglows and radio afterglows it is 40% and 33%, respectively. The origin
of these optically "dark bursts" could either be due to intrinsic effects (i.e.
inadequate search due to rapid evolution of the afterglow and/or an under-
energetic GRB) [6, 7], or an extrinsic effect (i.e. extinction of the optical flux
caused by circumburst dust or by the intergalactic medium) [8, 9].
To accurately derive the fraction of "radio quiet" bursts it is necessary to
incorporate both detections and upper limits in a statistically sound manner.
This has been done in Fig. 1 where flux density distribution at 8.5 GHz is
shown for a sample of 44 GRBs, toward which measurements or upper limits
have been made between 5 and 10 days after a burst. The time since the burst
is an important variable since radio light curves do not exhibit the simple
power-law decays seen in X-ray and optical afterglows, but rise to a peak on
average about one week after the burst and decay on timescales of a month.
The mean of the 19 detections in Fig. 1 is 315±82 µJy. Adding in the non-
detections, and using the Kaplan-Meier estimator [10] shifts this to 186±40
µJy. Approximately 50% of all bursts have radio afterglows at 8.5 GHz above
110 µJy, while fewer than 10% exceed 500 µJy. The relatively small range
of peak flux densities in Fig. 1 suggests that the fraction of "radio quiet"
bursts is largely determined by instrumental sensitivities. With the arcsecond
localizations provided by the Swift satellite (launch in 2004) it will be possible
to routinely detect all afterglows with centimeter radio emission above 100
µJy. Increasing the fraction of detected radio afterglows significantly above
50% will require the sensitivity improvements provided by the Expanded Very
Large Array3 (complete in 2010).
From this sample of peak flux densities we also derive the peak spectral
L (1 + z)1+β−α, where Fν ∝
radio luminosity in Fig. 1 given by Lν = 4πFν d2
tανβ and α = 1/2 and β = 1/3 has been assumed, corresponding to an
optically thin, rising light curve. The GRB redshifts lie in the range between
z =0.36 to z=4.5. The peak of the distribution is centered on 1031 erg s−1
Hz−1 and is similar to low-luminosity FRI radio galaxies like M87. More
interestingly, a comparison between this GRB sample and a sample of Type
Ib/c supernovae [11] shows that the later is four orders of magnitude less
luminous. Since radio emission is sensitive to the relativistic energy content of
the shock, independent of the initial geometry of the explosion, this has been
used to argue that the majority (<97%) of nearby Type Ib/c supernovae do
not produce a GRB-like event, such as that seen toward SN 1998bw [12].
3 Phenomenology and Interpretation
In this section we will follow the evolution a GRB and its radio afterglow
depicted schematically in Fig. 2. The observations span four orders of mag-
nitude in time (0.1-1000 days) and three orders of magnitude in frequency
3 http://www.aoc.nrao.edu/evla/
4
Dale A. Frail
(0.8-660 GHz), so it should be no surprise that radio light curves exhibit a
rich phenomenology. To interpret these observations we will rely on the highly
successful "standard fireball model" [13]. In this model there is an impulsive
release of kinetic energy (∼1051 erg) from the GRB event which drives an
ultra-relativistic outflow into the surrounding medium whose hydrodynami-
cal evolution is governed by the kinetic energy released, the density structure
of the circumburst medium and the geometry of the outflow. Synchrotron
emission is produced by this relativistic shock which accelerates electrons to
a power-law distribution. It is through the study of temporal (and spectral)
evolution of afterglow light curves that we can gain insight into the physical
conditions of the shock and the central engine that produced it.
−2
t
1−5 days
10−30 days
1 day
1/2
t
−1/3
t
Reverse Shock
Forward Shock
θ,
Reverse shocks: Γ
Interstellar scintillation:
E/n
Synchrotron self−absorption:
n
θ > 0.2
Wide−angle jets:
Circumburst environment:
Non−relativistic evolution:
Host galaxies:
jet
ISM vs. Wind, n
E, n, B,
θ
Obscured Star Formation
Non−Relativistic
−2.2
t
−1.2
t
50−100 days
Host Galaxy
Fig. 2. A schematic radio afterglow light curve. Timescales and scalings for the
temporal evolution are indicated. The list summarizes aspects of the flux evolution
which are unique to the radio bands (Lorentz factor, Γ ; source size, θ; energy, E;
density, n; jet opening angle, θjet; density profile; magnetic field strength, B; and
obscured star formation rate).
Despite response times as short as 2 hrs, centimeter searches (§2) are rarely
successful until a day or more after a burst. Broadband afterglow spectra
show that centimeter emission is attenuated as a result of synchrotron self-
absorption [14]. Typical observed values for the self-absorption frequency νa
are 5-10 GHz. It is interesting to note that the flux density below νa has
the form Fν ∝ ν 2, not the 5/2 spectral slope usually seen toward most radio
sources. This is because the relativistic shock accelerates electrons to a power-
law distribution (with energy index p given by N(γe) ∝ γ−p
e ) above a minimum
energy γm, which initially radiate their energy most of their energy at νm >>
The Radio Afterglows of Gamma-Ray Bursts
5
νa. The flux below νa depends only the angular size of the source and the
fraction of the shock energy that goes into accelerating electrons [1], and thus
it is a useful diagnostic of the ratio of the energy of the shock and the density
of the circumburst medium (E/n).
As this optically thick radio source expands, a monotonic rise in the flux
would be expected. It was therefore a considerable surprise when early obser-
vations of GRB 970508 showed erratic, short term (∼ hrs) and narrow band
(∼GHz) fluctuations in the centimeter emission [15]. The origin of these varia-
tions [2] was traced to the scattering of the radio emission, owing to the small
angular size of the fireball, as it propagates through the turbulent ionized
gas of our Galaxy. This is a large and complex subject [16, 17], but for the
purpose of this review it is sufficient to note that for typical lines of sight the
modulation of the flux densities is near a maximum at frequencies near 5-10
GHz. Coincidently, this is the same frequency range where νa typically lies
and where the majority of radio observations are being made. While interstel-
lar scintillation adds a certain degree of complexity to interpreting afterglow
light curves, it also allows us to use the Galaxy as a large lens to effectively
resolve the fireball. The observed "quenching" of diffractive scintillation from
GRB 970708 four weeks after the burst [18, 19] lead to estimate of the angular
size, demonstrating superluminal expansion and providing an early confirma-
tion of the fireball model.
In many instances [20, 21, 22, 23, 24] bright, short-lived radio "flares"
are detected at early times (t < 3 d). The emission is much brighter than
expected from a backward extrapolation of the light curve, and the level of
fluctuation is too great to be accounted for by interstellar scintillation. One
of the best-known examples is the radio flare of GRB 990123 [20], which was
accompanied by a 9th magnitude optical flash [25]. This prompt optical and
radio emission is thought to be produced in a strong reverse shock which
adiabatically cools as it expands back through the relativistic ejecta [26]. The
strength and lifetime of this reverse shock emission is sensitive to the initial
Lorentz factor, Γ◦, of the shock and the density structure of the circumburst
medium [27, 24]. To properly constrain these values requires that the peak
of the emission be measured. This is difficult to do with optical observations,
which require a response time on the order of the burst duration, while radio
observations require a response time of only 12-48 hrs.
On a timescale of days to weeks after the burst, the subsequent evolution
of the radio afterglow (Fig. 2) can be described by a slow rise to maximum,
followed by a power-law decay. The radio peak is often accompanied by a
sharp break in the optical (or X-ray) light curves [28, 29]. The most commonly
accepted (but not universal) explanation for these achromatic breaks is that
GRB outflows are collimated. The change in spectral slope, α, where Fν ∝
tανβ, occurs when the Γ of the shock drops below θ−1
, the inverse opening
angle of the jet [30, 31]. Since the radio emission at νR initially lies below the
synchrotron peak frequency νm the jet break signature is distinctly different
than that at optical and X-ray wavelengths. Prior to the passage of νm the
j
6
Dale A. Frail
jet break is expected to give rise to a shallow decay t−1/3 or plateau t0, in
the optical thin (νa < νR) or thick (νa > νR) regimes, respectively. Another
recognizable radio signature of a jet-like geometry is the "peak flux cascade",
in which successively smaller frequencies reach lower peak fluxes (i.e. Fm ∝
ν 1/2
m ). Taken together, these observational signatures can be used to infer the
opening angles θj of wide angle jets. Such jets are hard to detect at optical
wavelengths because the break is masked by the host galaxy, which typically
dominates the light curve between a week and a month after the burst [32, 33].
Once the real geometry of the outflow is known [34, 35] the energy released
in the GRB phase and the afterglow phase can be determined.
As noted above, the radio band is fortuitously located close to νa and
as such it is a sensitive probe of the density structure of the circumburst
medium. Extensive broadband modeling [36] has yielded densities in the range
0.1 cm−3 < n< 100 cm−3, with a canonical value of order n≃10 cm−3. Such
densities are found in the diffuse interstellar clouds of our Galaxy, commonly
associated with star-forming regions. A density of order 5-30 cm−3 is also
characteristic of the interclump medium of molecular clouds, as inferred from
observations of supernova remnants in our Galaxy (e.g., Chevalier 1999 and
references therein). Based on X-ray and optical observations alone, there have
been claims of high n≫ 104 cm−3 [38, 39] or low n≪ 10−3 cm−3 [40] cir-
cumburst densities. However, in several of these cases when the radio data
has been added to the broadband modeling (i.e. constraining νa), there is no
longer any support for either extreme of density [22, 33].
One unsolved problem on the structure of the circumburst environment
is the absence of an unambiguous signature of mass loss from the presumed
massive progenitor star in afterglow light curves [41]. Although there are some
notable exceptions (e.g., Price et al. 2002), most GRB light curves are best
fit by a jet expanding into a constant density medium instead of a radial
density gradient, ρ ∝ r−2 [36]. Part of the solution may lie in reduced mass
loss rates due to metalicity effects, or the motion of the star through a dense
molecular cloud [43], both of which act to shrink the radius that the pre-burst
wind is freely expanding. It is equally likely that our failure to distinguish
between different models of the circumburst medium is due to the lack of
early afterglow flux measurements, especially at millimeter and submillimeter
wavelengths where the largest differences arise [44, 45]. The resolution of this
conflict is important as it goes to the heart of the GRB progenitor question.
At sufficiently late times, when the rest mass energy swept up by the
expanding shock becomes comparable to the initial kinetic energy of the ejecta
(∼100 days), the expanding shock may slow to non-relativistic speeds [46]. A
change in the temporal slope is expected at this time (Fig. 2) with αN R =
(21 − 15p)/10 for a constant density medium, independent of geometry. This
dynamical transition provides a simple and power method to derive the kinetic
energy of the outflow which has expanded to be quasi-spherical at this time.
In contrast, most energy estimates made at early times require knowledge
of the geometry of the outflow [40, 47, 35]. Using the late-time radio light
The Radio Afterglows of Gamma-Ray Bursts
7
curves and the robust Taylor-Sedov formulation for the dynamics we can infer
quantities such as the kinetic energy, ambient density, magnetic field strength,
and the size of the fireball. The radius can be checked for consistency with the
equipartition radius and the interstellar scintillation radius. This method has
been used for GRB 970508 [19] and for GRB 980703 (Berger, priv. comm.),
yielding energies of order f ew × 1050 erg, in agreement with other estimates.
Finally, the radio light curves at late times may flatten due to the presence
of an underlying host galaxy. Most GRBs studied to date have optical/NIR
hosts but only about 20% have been seen at centimeter and submillimeter
wavelengths [48, 49, 50]. This radio emission, if produced by star formation,
implies star formation rates SRF∼ 500 M⊙ yr−1 and Lbol > 1012 L⊙, clearly
identifies these GRB hosts as ultraluminous starburst galaxies which are all
but obscured by dust at optical wavelengths. This is an emerging area with
great potential for studying cosmic star formation with a sample of galax-
ies selected quite differently than other methods. Preliminary studies have
already shown that GRB-selected galaxies are significantly bluer than other
radio-selected samples [50].
Acknowledgements. DAF would like to thank his many collaborators
in the radio afterglow network, especially Shri Kulkarni and Edo Berger.
References
1. Katz, J. L. and Piran, T. ApJ, 490, 772, (1997).
2. Goodman, J. New Astr., 2(5), 449 -- 460, (1997).
3. Hurley, K., et al. ApJ, 567, 447 -- 453, (2002).
4. Frail, D. A., et al. in AIP Conf. Proc. 526: Gamma-ray Bursts, 5th Huntsville
Symposium, 298 -- 302, (2000).
5. Frail, D. A., et al. AJ, 125, 2299 -- 2306, (2003).
6. Fynbo, J. U., et al. A&A, 369, 373 -- 379, (2001).
7. Berger, E., et al. ApJ, 581, 981 -- 987, (2002).
8. Piro, L., et al. ApJ, 577, 680 -- 690, (2002).
9. Djorgovski, S. G., et al. ApJ, 562, 654 -- 663, (2001).
10. Feigelson, E. D. and Nelson, P. I. ApJ, 293, 192 -- 206, (1985).
11. Berger, E. et al. A Radio Survey of Type Ib and Ic Supernovae: Searching for
Engine Driven Supernovae. ApJ, in press; astro-ph/0307228, (2003).
12. Kulkarni, S. R., et al. Nature, 395, 663 -- 669, (1998).
13. M´esz´aros, P. Ann. Rev. Astr. Ap., 40, 137 -- 169, (2002).
14. Galama, T. J., et al. ApJ, 541, L45 -- L49, (2000).
15. Frail, D. A., et al. Nature, 389, 261 -- 263, (1997).
16. Rickett, B. J. Ann. Rev. Astr. Ap., 28, 561 -- 605, (1990).
17. Galama, T. J., et al. ApJ, 585, 899 -- 907, (2003).
18. Waxman, E., Kulkarni, S. R., and Frail, D. A. ApJ, 497, 288 -- 293, (1998).
19. Frail, D. A., Waxman, E., and Kulkarni, S. R. ApJ, 537, 191 -- 204, (2000).
20. Kulkarni, S. R., et al. ApJ, 522, L97 -- L100, (1999).
21. Frail, D. A., et al. ApJ, 538, L129 -- L132, (2000).
22. Harrison, F. A., et al. ApJ, 559, 123 -- 130, (2001).
8
Dale A. Frail
23. Yost, S. A., et al. ApJ, 577, 155 -- 163, (2002).
24. Berger, E., et al. ApJ, 587, L5 -- L8, (2003).
25. Akerlof, C., et al. Nature, 398, 400 -- 402, (1999).
26. Sari, R. and Piran, T. ApJ, 517, L109 -- L112, (1999).
27. Soderberg and Ramirez-Ruiz. Flaring up: Radio Diagnostics of the Kinematic,
Hydrodynamic and Environmental Properties of GRBs. MNRAS, in press;
astro-ph/astro-ph/0210524, (2002).
28. Harrison, F. A., et al. ApJ, 523, L121 -- L124, (1999).
29. Berger, E., et al. ApJ, 545, 56 -- 62, (2000).
30. Rhoads, J. E. ApJ, 525, 737 -- 749, (1999).
31. Sari, R., Piran, T., and Halpern, J. P. ApJ, 519, L17 -- L20, (1999).
32. Berger, E., et al. ApJ, 556, 556 -- 561, (2001).
33. Frail, D. A., et al. ApJ, 590, 992 -- 998, (2003).
34. Frail, D. A., et al. ApJ, 562, L55 -- L58, (2001).
35. Bloom, J. S., Frail, D. A., and Kulkarni, S. R. GRB Energetics and the GRB
Hubble Diagram: Promises and Limitations. ApJ in press, astro-ph/0302210,
(2003).
in' t Zand, J., et al. ApJ, 559, 710 -- 715, (2001).
36. Panaitescu, A. and Kumar, P. ApJ, 571, 779 -- 789, (2002).
37. Chevalier, R. A. ApJ, 511, 798 -- 811, (1999).
38. Dai, Z. G. and Lu, T. ApJ, 537, 803 -- 809, (2000).
39.
40. Panaitescu, A. and Kumar, P. ApJ, 554, 667 -- 677, (2001).
41. Chevalier, R. A. and Li, Z. ApJ, 536, 195 -- 212, (2000).
42. Price, P. A., et al. ApJ, 572, L51 -- L55, (2002).
43. Wijers, R. A. M. J. in Gamma-ray Bursts in the Afterglow Era, 306 -- 311, (2001).
44. Panaitescu, A. and Kumar, P. ApJ, 543, 66 -- 76, (2000).
45. Yost, S. et al. A Study of the Afterglows of Four GRBs: Constraining the
Explosion and Fireball Model. ApJ, in press; astro-ph/0307056, (2003).
46. Wijers, R. A. M. J., Rees, M. J., and M´esz´aros, P. MNRAS, 288, L51 -- L56,
(1997).
47. Berger, E., Kulkarni, S. R., and Frail, D. A. ApJ, 590, 379 -- 385, (2003).
48. Berger, E., Kulkarni, S., and Frail, D. A. ApJ, 560, 652 -- 658, (2001).
49. Frail, D. A., et al. ApJ, 565, 829 -- 835, (2002).
50. Berger, E., et al. ApJ, 588, 99 -- 112, (2003).
|
astro-ph/9709290 | 2 | 9709 | 1997-09-30T07:19:23 | Dynamical Friction for Compound Bodies | [
"astro-ph"
] | In the framework of the fluctuation-dissipation approach to dynamical friction, we derive an expression giving the orbital energy exchange experienced by a compound body as it moves interacting with a non homogeneous discrete background. The body is assumed to be composed of particles endowed with a velocity spectrum and with a non homogeneous spatial distribution. The Chandrasekhar formula is recovered in the limit of a point-like satellite with zero velocity dispersion and infinite temperature moving through an homogeneous infinite medium. In this same limit, but dropping the zero satellite velocity dispersion ($\sigma_S$) condition, the orbital energy loss is found to be smaller than in the $\sigma_S=0$ case by a factor of up to an order of magnitude in some situations. | astro-ph | astro-ph |
Mon. Not. R. Astron. Soc. 000, 1 -- 10 (0000)
Printed 13 September 2016
(MN LATEX style file v1.4)
Dynamical friction for compound bodies
R. Dom´ınguez-Tenreiro & M.A. G´omez-Flechoso⋆
Dept. F´ısica Te´orica, C-XI. Univ. Aut´onoma de Madrid, E-28049 Madrid
Accepted ..... Received .....; in original form 1997 May 14
ABSTRACT
In the framework of the fluctuation-dissipation approach to dynamical friction, we
derive an expression giving the orbital energy exchange experienced by a compound
body as it moves interacting with a non homogeneous discrete background. The body
is assumed to be composed of particles endowed with a velocity spectrum and with
a non homogeneous spatial distribution. The Chandrasekhar formula is recovered in
the limit of a point-like satellite with zero velocity dispersion and infinite temperature
moving through an homogeneous infinite medium. In this same limit, but dropping
the zero satellite velocity dispersion (σS) condition, the orbital energy loss is found to
be smaller than in the σS = 0 case by a factor of up to an order of magnitude in some
situations.
Key words: methods: analytical -- celestial mechanics, stellar dynamics, galaxy
dynamics.
1
INTRODUCTION
A satellite moving through a field of gravitating particles
experiences a dissipative frictional force known as dynam-
ical friction. It can be understood in terms of the satellite
wake, that exerts a drag force on the satellite itself (Kalnajs
1972; Binney & Tremaine 1987; Weinberg 1989; Bekenstein
& Zamir 1991). It can also been understood in terms of
the underlying basic physics as the friction resulting from
the fluctuating gravitating forces acting on the satellite as
a consequence of the non-continuos character of the particle
system. The fact that fluctuating forces cause dissipation is
quite a general scheme in Physics. It is at the basis of phys-
ical phenomena such as electric resistance in conductors or
viscous friction in liquids (Reif 1965).
Dynamical friction has important consequences in the
evolution of astronomical systems, mainly because it causes
a decay of orbiting bodies, so that merger timescales and dis-
sipation rates by dynamical friction are closely related. The
merger scenario is at the basis of a great deal of processes
in Astronomy. Not only is it the framework for the general
galaxy formation picture in hierarchical cosmological mod-
els, but also for more particular aspects of the evolution of
a number of astronomical systems, such as galactic nuclei,
cD galaxies in rich galaxy clusters, compact galaxy groups
and so on.
A dynamical friction formula was first obtained, in a
⋆ Present address: Observatoire de Gen`eve, Ch. des Maillettes
51, Ch-1290 Sauverny (Switzerland)
c(cid:13) 0000 RAS
kinematical approach, by Chandrasekhar (1943). He has cal-
culated the rate of momentum exchange between the test
and field particles as the result of a sum of uncorrelated
two-body encounters, obtaining the expression:
M
dv
dt
= −
4πG2M 2
v3
ln Λ ρ(< v) v,
(1)
where M and v are the test particle mass and speed, re-
spectively, ρ(< v) is the density of background particles
with velocity less than v, G is the gravitation constant and
Λ = pmax
, with pmax and pmin the maximum and minimum
pmin
impact parameters contributing to the drag. Keplerian or-
bits for both the test and field particles and an infinite and
homogeneous background have been assumed in the deriva-
tion of eq. (1)
Chandrasekhar's formula is widely employed to quantify
dynamical friction in a variety of situations, even if in most
astronomical problems the background is neither infinite nor
homogeneous. This formula is known to give the correct or-
der of magnitude, but it suffers from several drawbacks, that
arise from the very physical assumptions made in its deriva-
tion. Furthermore, it cannot describe some situations, as for
example the drag experienced by a satellite placed outside
the edge of a finite gravitating system. In fact, the satel-
lite would be decelerated on physical grounds, because it
causes a perturbation to the system. However, according to
eq. (1), the drag would vanish. For this reason, other works
on dynamical friction followed Chandrasekhar's pioneering
study, either from a numerical (Lin & Tremaine 1983; White
1983; Bontekoe & van Albada 1987; Zaritsky & White 1988)
or an analytical point of view. Analytical descriptions have
2
R. Dom´ınguez-Tenreiro & M.A. G´omez-Flechoso
the advantage that they help understanding the underly-
ing physics. Several methods have been developed: Fokker-
Planck equation based (Rosenblunth, Mc Donald & Judd
1957; Binney & Tremaine 1987) polarization cloud approach
(Marochnik 1968; Kalnajs 1972; Binney & Tremaine 1987;
Weinberg 1989; Bekenstein & Zamir 1991), resonant parti-
cle interactions (Lynden-Bell & Kalnajs 1972; Tremaine &
Weinberg 1984; Palmer & Papaloizou 1985; Weinberg 1986).
Berkenstein and Maoz (1992, hereafter BM92) and
Maoz (1993, hereafter M93) introduced a fluctuation-
dissipation approach to dynamical friction. The fluctuation-
dissipation theorem (Kubo 1959) relates the friction coeffi-
cient to the time integral of the correlation function for the
fluctuations causing the friction. They showed that dynam-
ical friction fits into this general scheme, which provides a
powerful technique to study it. This approach is in fact a
return to Chandrasekhar's original attempt to give a sta-
tistical description of dynamical friction. Other stochastic
approaches to dynamical friction are those of Cohen (1975)
and Kandrup (1980). M93 derived a formula for the drag
experienced by an object which travels in an arbitrary mass
density field, assumed to be stationary and formed by par-
ticles much lighter than the object.
A common feature of all the previous approaches is that
they treat the satellite as being rigid and without struc-
ture in the velocity space. This approximation can be good
enough in a number of astronomical situations, but in oth-
ers, these two ingredients could play a crucial role. As a first
example of such a situation, let us consider the dynamical
evolution of compact groups of galaxies (e.g. Mamon 1993).
In this case, the velocity dispersion of individual galaxies is
comparable to the velocity dispersion of the common halo
that hosts them. As a second example, we recall that in the
problem of the interaction of two comparable mass galax-
ies, the energy exchange due to dynamical friction cannot
be calculated in the previous frameworks, because their pre-
sumably comparable velocity dispersions need to be taken
into account.
In this paper we present a fluctuation-dissipation study
of the orbital changes experienced by a non-rigid satellite,
composed of gravitating particles with a finite velocity dis-
persion, as it interacts with a general background. An exten-
sion of BM92 and M93 techniques has allowed us to calculate
the rate of energy exchange between them as a result of fluc-
tuations in the gravitational forces of both the background
and the satellite.
The paper is organized as follows: in Sect. 2, the phys-
ical formulation of the method and the general expressions
giving the energy exchange rate are presented. In Sect. 3,
we calculate the instantaneous energy variations for general
backgrounds at rest and with a Maxwellian velocity distribu-
tion. Some particular limits are dealt with in Sect. 4. Finally,
in Sect. 5, we summarize and discuss our results. Two Ap-
pendices follow, where the results of the calculation of the
correlation matrix and of an integral needed are given.
2 PHYSICAL FORMULATION
We will study the energy exchange between two self-
gravitating equilibrium systems called the background (B)
and the satellite (S). The background consists of NB equal
mass particles, each of them with a mass mB. Its total mass
and typical size are MB = mBNB and RB, respectively.
These particles exhibit a velocity spectrum with dispersion
σB and zero mean at t = 0 (which is equivalent to consid-
ering the origin of the reference system placed at the halo
center of mass at t = 0). The satellite in turn is composed of
NS equal mass particles of mass mS, total mass MS, typical
size RS, and a velocity distribution with dispersion σS and
mean equal to the center of mass velocity of the satellite,
vCM S.
As we are mainly interested in the effects that a non
vanishing σS would have on the evolution of the whole sys-
tem, we can consider that σS ∼ σB ≡ σ. The virial theorem
tells us that σ ∼ GN mR−1, so that σS ∼ σB implies that
NSmSR−1
(2)
S ∼ NBmBR−1
B .
As both the satellite and the background are assumed
to have a finite size, we can safely consider that
(3)
N 1/n
Rγ(cid:19)m
α ≫(cid:18) Rβ
for whatever α, β, γ = B or S, and n, m of order unity. If
mα ≫ m α ( α is the particle class contrary to α), as could
be the case, for example, when the α and α systems are
composed of stars and dark matter particles, respectively,
then eqs. (2) and (3) imply the inequality Nα ≪ N α. But,
in any case, both Nα and N α will be assumed to be very
large.
2.1 The fluctuating forces acting on a particle
The fluctuating force, F iS (riS , t), acting at time t on a satel-
lite particle iS placed at position riS and caused by its in-
teractions with both background and satellite particles, can
be written as:
F iS (riS , t) = F B
iS (riS , t) + F S
iS (riS , t)
(4)
where
F B
iS (riS , t) = −GmSmB∇ XjB
riS − r (cid:19)
− Z drnB(r)
and
1
riS − rjB
iS (riS , t) = −Gm2
F S
(5)
(6)
riS − rjS
1
S∇ XjS6=iS
riS − r (cid:19)
Z drnS(r)
−
NS − 1
NS
are the fluctuating forces acting on iS caused by the back-
ground and satellite, respectively. Each of these forces re-
sults from subtracting to the total many-body or discrete
force a smooth part derived from the mean field potentials,
ΦB(r) and ΦS(r), due to the smooth densities, nB(r) and
nS(r), respectively. These satisfy the relations:
nα(r) = NαZ duf α
0 (r, u),
(7)
c(cid:13) 0000 RAS, MNRAS 000, 1 -- 10
∇2Φα(r) = 4πGmαnα(r)
and
Z drnα(r) = Nα,
(8)
(9)
NS
where α = B, S, f α
0 (r, u) is the one-particle distribution
function for background and satellite particles in the unper-
turbed state, and the NS−1
factor takes into account that
the iS particle does not interact with itself at time t. An
exchange of the B and S labels in eq. (5) gives the expres-
sion for the fluctuating force on the background particle iB
caused by the satellite. Changing the S labels into B in eq.
(6) we get the force on iB caused by the background fluctua-
tions. In a compact formulation, we can write the fluctuating
force on a generic class α particle, iα, due to class β particles
as:
F β
iα (riα , t) = −Gmαmβ∇
Xjβ6=iα
riα − r !
Nβ Z drnβ(r)
N
β
−
′
1
riα − rjβ
(10)
′
where now N
that α = β.
β = Nβ − δαβ takes into account the possibility
The equation of motion of either a background or a
satellite particle at time t can be written in a compact no-
tation as:
viα (t) = viα ,0 +Xβ (cid:20) 1
mα Z t
∇Φβ(riα , t′)dt′(cid:21) ,
0
− Z t
0
F β
iα (riα , t′)dt′
(11)
where viα ,0 is the particle velocity at time t = 0.
Stochastic forces are weak as compared with the smooth
global forces. In fact, the fluctuating force between a class
α and a class β particle is of order (Kandrup 1980, Saslaw
1985)
F β
iα,f luc = O(cid:0)Gmαmβn2/3(cid:1) ,
where n ∼ NαR−3
is the average particle
number density for the whole system. The smooth density
of class δ particles causes a gravitating force on a class γ
particle that is of order:
jγ ,smooth = O(cid:0)Gmγ mδNδR−2
F δ
δ (cid:1) .
Eqs. (12) and (13) give, after some algebra:
α + (1− δαβ)NβR−3
(12)
(13)
β
F β
iα,f luc
F δ
jγ ,smooth
= O "(cid:16) mβ
+ (1 − δαβ)(cid:16) mβ
Rβ(cid:19)2
mδ(cid:17)1/2(cid:18) Rδ
mδ(cid:17)3/2(cid:21)2/3
δ !,(14)
where eq. (2) has been taken into account. Now, if mS ∼
mB, then we will have
N−1/3
mα
mγ
c(cid:13) 0000 RAS, MNRAS 000, 1 -- 10
Dynamical friction for compound bodies
3
, (15)
N−1/3
δ
+ (1 − δαβ)#2/3
= O
"(cid:18) Rδ
Rβ(cid:19)2
F β
iα,f luc
F δ
jγ ,smooth
this is indeed a very small quantity. If, on the contrary, back-
ground and satellite particles have very different masses, eq.
(14) must be examined more carefully. The most unfavor-
able case is when mδ ≪ mδ, γ = δ and α = β = δ. In this
case, eq. (14) gives:
F β
iα,f luc
F δ
jγ ,smooth
and it would be sufficient to ask that
mδi4/3(cid:20) Rδ
Rβ(cid:21)4/3
δ ! ,
N−1/3
(16)
= O h mβ
mδ(cid:17)4/3
N 1/3
δ ≫(cid:16) mδ
(17)
to ensure that the fluctuating forces are much smaller than
the smooth ones.
Next we compare the timescale for fluctuation, ταβ, and
the timescale Tγδ over which the velocity of class γ parti-
cles changes appreciably as a result of the smooth forces
produced by class δ particles. The first is of order of the
time required for the nearest neighbor to travel an inter-
particle distance, ταβ = 1
σn1/3 , where σ is σα or σ α and
n = nα + (1− δαβ)nβ. As it may happen that nα and nβ are
very different, we take the maximum ταβ timescale, corre-
sponding to nmin = NminR−3
min. The second timescale is set
by Tγδ = vγ
, so that we can write
∇Φδ ∼ vγRδ
σ2
τ max
αβ
Tγδ ∼ O(cid:16)N−1/3
min
Rmin
Rδ (cid:17) ≪ 1,
where we have used the inequality (3).
(18)
The rate of energy exchange between subsystems B and
S will be obtained as an integral on time which involves
the correlation matrix (see below and BM92 and M93). The
correlation matrix is known to fall with time faster than
t−1 (Cohen 1975; Kandrup 1980; M93), so that it can be
taken to vanish for times much larger than the fluctuation
timescale, and in particular for times of the order of the
macroscopic timescale, Tγδ. As a consequence, in this work
we will consider time intervals after t = 0, δt, which are
large as compared with the fluctuation timescale, τ max
αβ , but
much shorter than the macroscopic timescale (see Reif 1965
and M93 for a detailed discussion)
τ max
αβ ≪ δt ≪ min(cid:20) vα
d
dt vα
,
∇Φγ (cid:21) .
vα
(19)
2.2 The energy exchange
The energy of a particle iα is not conserved. Its total in-
stantaneous variation at time t, due to the fluctuating forces
caused by class β particles is given by:
iα
(cid:18) dEβ
dt (cid:19)t
= F β
iα (t) · viα (t).
(20)
Taking into account the expression for viα (t) given by
eq. (11), we get for times t verifying (19)
iα
(cid:18) dEβ
dt (cid:19)t
= F β
iα (t) ·"viα,0 +
1
mα Xγ Z t
0
dt′F γ
iα (t′)# , (21)
4
R. Dom´ınguez-Tenreiro & M.A. G´omez-Flechoso
where the potential terms have been neglected because
viα,0 ≫ ∇Φγ t, as ensured by (19). The total energy
change of the iα particle during the time interval (0, δt) is
easily calculated by integrating eq. (21). Summing up on iα
we get the total energy variation of class α particles due
to the fluctuating forces caused by β particles in this time
interval:
exponential in eq.(23), we obtain that the change in the
probability function is given approximately by:
K = 1 −Xα Xiα
1
f α
∂f α
∂εiα
δεiα .
(27)
For most astronomical applications it is justified to use
an isothermal Maxwellian one-particle distribution function,
∆Eβ
dtF β
α,tot(δt) = Xiα Z δt
mα Xγ Z t
+
1
0
0
iα (t)·(cid:20) viα,0
iα (t′)# .
dt′F γ
f α
0 (r, u) =
nα(r)
Nα (2πσ2
α)3/2 exp [−βαεiα ] ,
(28)
(22)
where u is the velocity vector, σα is the velocity dispersion of
class α particles and βα = 1
is an inverse temperature.
Introducing this expression for the distribution function in
eq. (27), the factor of change becomes:
mα σ2
α
We will be interested in the average value of this quan-
tity.
2.3 Statistical averaging
The presence and motion of the iS satellite particle perturbs
the background. As a result, while the statistical (ensemble)
average of the background fluctuating forces acting on the
satellite vanishes in the unperturbed state, hF B
iS (t)i0 = 0,
they do not vanish anymore in the real perturbed system,
hF B
iS (t)i 6= 0. The same is true in general for the stochas-
tic forces caused by class β particles and acting on class
α particles: hF β
iα (t)i 6= 0, because the
phase density and the number of microstates available to
the system has changed due to the perturbation. Following
BM92 and M93, we assume that the probability of finding a
dynamical variable, Q, with a given value, P [Q], is propor-
tional to the change in the number of microstates available
to the whole system. This change is given by the factor:
iα (t)i0 = 0, but hF β
K = exp [δS] ,
(23)
where δS is the entropy change of the system as a result of
the distortion. The expected value of a dynamical function,
Q, can now be written as:
,
(24)
hQi = R dΓQ(Γ)f0(Γ)K
R dΓf0(Γ)K
where Γ are the variables defining the phase space of the
NB + NS particles and f0(Γ) is the distribution function for
the unperturbed state.
As a result of the fluctuations, a generic particle initially
with energy εiα , will change to an energy ε′iα = εiα + δεiα ,
α = B, S. Recalling the definition of entropy:
S = S0 −Z f ln f dΓ,
(25)
with S0 a constant, this energy change results into a total
entropy variation given by:
δS =Xα "−Xiα
ln f α(ε′iα ) +Xiα
ln f α(εiα )# ,
(26)
where f α(εiα ) is the class α one particle distribution func-
tion for the unperturbed state corresponding to an en-
ergy εiα . Most two-body encounters are weak, and then
δεiα ≪ εiα . Expanding the r.h.s. of eq.(26) and then the
K = 1 +Xα
βαXiα
δεiα .
(29)
The next step is to find out an expression for the energy
variation. The total energy of the system must be conserved,
so that we have:
δεiα = 0.
(30)
Xα Xiα
Moreover, for one given particle, iα, its energy variation
can be expressed as:
δεiα = ∆Eα
iα + ∆E α
iα − aiα ∆Eα
α,tot − biα ∆Eα
α,tot,
(31)
where aiα is the fraction of the total class α autointeraction
energy, ∆Eα
α,tot, absorbed by particle iα, and biα is the frac-
tion of the total interaction energy between class α and α
particles, caused by the fluctuating forces of the α subsys-
tem, absorbed by iα particle. Summing on iα in eq. (31), we
get the total energy change of subsystem α:
δεiα = ∆E α
α,tot − ∆Eα
α,tot,
Xiα
(32)
that Piα
aiα = Piα
compatible with energy conservation (eq. (30)). To write
down eq. (32) from (31), it has been taken into account
biα = 1. Energy signs are such that
∆Eγ
iα > 0 if particle iα gains energy due to the fluc-
tuating forces caused by class γ particles and conversely.
With this convention, if energy flux is from subsystem α
to subsystem α, then ∆E α
α,tot < 0 (subsystem α loses en-
ergy), and ∆Eα
α,tot > 0 (subsystem α gains energy), so that
δεi α > 0. Inserting eq. (32) in eq.
(29) and recalling eq. (22), we obtain:
Piα
δεiα < 0 and Pi α
βα(Xiα Z δt
0
1
+
K = 1 +Xα
mα Xγ Z t
m α Xγ Z t
+
1
0
0
dtF α
iα (t)·" viα,0
iα (t′) − Xi α Z δt
(t′)#) .
0
dt′F γ
dt′F γ
i α
dtF α
i α (t)·" vi α ,0
(33)
Eq. (24) allows us now to write the ensemble average
of the instantaneous energy variation at time t of a generic
class µ particle due to the stochastic forces caused by class
ν particles:
c(cid:13) 0000 RAS, MNRAS 000, 1 -- 10
(cid:28)(cid:18) dEν
0
1
iµ
+
dtF ρ
iµ (δt)·(cid:18) viµ,0
dt (cid:19)δt(cid:29) = Z dΓf0(Γ)(cid:20) F ν
iµ (t)!#"1 +Xα
mµ Xρ Z δt
· viα,0 +
· vi α,0 +
mα Xγ Z t
m α Xγ Z t
βα(Xiα Z δt
iα (t′)! −Xi α Z δt
(t′)!)#(cid:20)Z dΓf0(Γ)K(cid:21)−1
dt′F γ
i α
dt′F γ
dtF α
dtF α
i α (t)
1
1
0
0
0
0
iα (t)
(34)
To second order in the fluctuating forces, taking into
account that the average of the stochastic forces in the un-
perturbed state vanishes, and summing on the iµ subindex
to obtain the global effect, we get:
(cid:28)(cid:18) dEν
mµ (cid:20)Z 0
−δt
1
µ,tot
dt (cid:19)δt(cid:29) =Xiµ
+ (βν − βν )Xj ν Z 0
−δt
dsviµ,0C νν
dsT r(cid:2)C νν
iµiµ (s)(cid:3)
iµj ν (s)vj ν ,0# ,
where
C αβ
iγ jδ
(s) = hF α
iγ (0) ⊗ F β
jδ
(s)i0
(35)
(36)
is the correlation matrix (see Appendix A; the symbol ⊗
iγ and F β
stands for the tensorial product of vectors F α
,
jδ
the average is on the unperturbed states of particles α and
β and the invariance of the correlation matrix under time
translation has been taken into account). Because in the
unperturbed state two different particles are not correlated,
the correlation matrix defined in the previous equation van-
ishes if α 6= β. This has been taken into account to deduce
eq. (35) from eq. (34). Note that because for s ≥ δt the
correlations vanish, the lower limit of the integrals can be
extended up to −∞. Eq. (35) is the expression of the global
instantaneous energy variation of subsystem µ due to the
fluctuating forces of subsystem ν. The integrand in the first
term of the r.h.s. of eq. (35) is invariant under time rever-
sal. Extending the integral to positive s, this term is the
corresponding power spectrum at zero frequency (Wiener-
Khintchine theorem, see Reif 1965 and BM92) and, conse-
quently, it is positive and represents a heating term of class
µ particles due to the fluctuating forces caused by ν parti-
cles. It is of order O( 1
N ) ≪ 1 relative to the second term.
Terms of this kind will be neglected in this work. The sec-
ond term can be either positive or negative, depending on
the sign of (βν − βν). In the next section it will be explicitly
calculated for µ, ν = B, S.
3 THE INSTANTANEOUS ENERGY
VARIATION
3.1 Energy exchange between the satellite and
the background
Energy flows between the satellite and the background as
a result of either the global energy variation of the satellite
c(cid:13) 0000 RAS, MNRAS 000, 1 -- 10
Dynamical friction for compound bodies
5
particles due to the fluctuating forces of the background, or,
conversely, the global energy variation of the background
particles due to the fluctuating forces of the satellite. These
fluxes are given by eq. (35) when µ 6= ν.
Let us first calculate the effect due to the stochastic
forces of the background. This term could cause an energy
flux responsible for the satellite deacceleration. Eq. (35) with
µ = S, ν = B and neglecting the heating term, gives:
(cid:28)(cid:18) dEB
S,tot
dt (cid:19)δt(cid:29) =
= (βS − βB)XiS XjS Z 0
−∞
dsviS ,0C BB
iS jS (s)vjS ,0,
(37)
where the correlation matrix is given by eq. (A10) with α =
B and γ = δ = S. In the case of a point-like satellite with
mS ≫ mB, βS ≪ βB and then the satellite loses energy to
the background. In our case, however, it is not excluded in
principle that βS > βB and then the energy would flow from
the background to the satellite.
The integral over ds in eq. (37) can be calculated taking
into account eq. (A10) and the equality:
Z 0
−∞
ds
A + V s
A + V s 3 =
A · V + AV
V 2A2
⊥
A −
h A
V
V i
where A⊥ is the projection of vector A on the plane normal
to vector V . In our case either V ≡ vjS − vhB or V ≡ vjS ,
. We
(see eq. (A10)), and A ≡ rjS − rhB or A ≡ rjS − r
obtain:
′
(38)
′
S,tot
Bm2
N
B
NB
(cid:28)(cid:18) dEB
dt (cid:19)δt(cid:29) = G2m2
S(βS − βB)
×XiS XjS "XhB (cid:26) (riS − rhB ) · viS
riS − rhB 3 (cid:27)
×(cid:26) [vjS · (rjS − rhB )][(rjS − rhB ) · (vjS − vhB )]
rjS − rhB vjS − vhB 2 (rjS − rhB )⊥ 2
−
rjS − rhB vjS − vhB 2 (rjS − rhB )⊥ 2
vjS − vhB (rjS − rhB )⊥ 2(cid:27)
vjS · (rjS − rhB )⊥
NBZ drdr′nB(r)nB(r′)
rjS − rhB 2 (vjS − vhB ) · vjS
riS − r 3 rjS − r′ (cid:21) (39)
[(riS − r) · viS ]
+
+
1
where the sum on a generic subindex, iα, means Nα times
the average on velocity and positions of class α particles, and
is carried out by means of the distribution function given in
eq. (28).
To proceed further, we recall that the velocity distri-
bution of background particles is assumed to be Maxwellian
with zero mean and dispersion σB, and the velocity distribu-
tion of the satellite particles has mean equal to the satellite
center of mass velocity, vCM S, and dispersion σS. The inte-
gration over dvhB gives:
S,tot
(cid:28)(cid:18) dEB
dt (cid:19)δt(cid:29) = G2m2
Bm2
S(βS − βB)
6
R. Dom´ınguez-Tenreiro & M.A. G´omez-Flechoso
jS )erf(γjS ) − 1
riS − rhB 3 (cid:27)
×XiS XjS (cid:20)Z drhB nB(rhB )(cid:26) (riS − rhB ) · viS
×(cid:26) exp(γ2
jS − y2
(cid:27)
rjS − rhB
riS − r 3 rjS − r′ (cid:21) (40)
NBZ drdr′nB(r)nB(r′)
[(riS − r) · viS ]
, γjS ≡
v jS√2σB
+
1
′
The integration over dvjS involves the integral:
)
and taking
(r jS −rhB
r jS −rhB
where erf stands for the error function, yjS ≡
yjS · ζ with ζ ≡
I =Z dvjS fM (vjS ) exp(γ2
jS − y2
In Appendix B we show that
jS )erf(γjS ).
N
B
NB
= 1.
(41)
(42)
I =
σ2
B
σ2
T
exp(α2
CM S − x2
B +σ2
CM S)erf(αCM S),
T = σ2
S, xCM S ≡ vCM S√2σT
where σ2
and αCM S ≡ xCM S·ζ.
The integration over dviS is trivial recalling that viS =
vCM S + uiS and that for a Maxwellian distribution the
contribution of uiS to the first moments vanishes. Taking
mαnα = ρα, we finally get:
(cid:28)(cid:18)dEB
σ2
S,tot
B
mSσ2
S −1(cid:19) G2
dt (cid:19)δt(cid:29)=(cid:18)mBσ2
BZ driS drjS ρS(riS )ρS(rjS )
riS − rhB 4 (cid:27)
×(cid:20)Z drhB ρB(rhB )(cid:26) (riS − rhB ) · vCM S
×(cid:26) σ2
CM S)erf(αCM S) − 1(cid:27)
NB Z drdr′ρB(r)nB(r′)
riS − r 3 rjS − r′ (cid:21)(43)
[(riS − r) · vCM S]
CM S − x2
exp(α2
B
σ2
T
+
1
tot
(cid:28)(cid:18) dES
dt (cid:19)δt(cid:29) =(cid:28)(cid:18) dES
=(cid:28)(cid:18) dEB
dt (cid:19)δt(cid:29)
S,tot
orb
dt (cid:19)δt(cid:29) +(cid:28)(cid:18) dES
dt (cid:19)δt(cid:29)
in
(45)
orb and ES
where ES
in stand for the orbital and internal energy
of the satellite, respectively, and the second equality results
from the zero value of the flow given by eq. (44). The rate
of change of the satellite internal energy can be calculated
with the help of eq. (24) with Q =
dES
in
dt
or:
d
dt
uiS
uiS ,
Q = mSXiS
where uiS = viS − vCM S is the velocity of iS particles with
respect to its center of mass. This gives an expression similar
to eq. (43), except that now vCM S takes a zero value, so that
(46)
in
(cid:28)(cid:18) dES
dt (cid:19)δt(cid:29) = 0,
(47)
and then eq. (43) gives, at second order in the fluctuations,
the rate of change of the satellite orbital energy.
3.2 The self-interaction energies of the
background and the satellite
When µ = ν = B or S, eq. (35) describes the instantaneous
energy variation rate of class µ particles due to the stochastic
forces caused by µ particles themselves.
According to the eq. (32) these energies do not play any
role in the energy change of the subsystems S or B. They
only represent the energy change of an individual particle
(see eq. (31)).
The self-interaction energy of the background is easily
obtained from eq. (35) with µ = ν = B:
A specification of the density distribution of both the
satellite and the background is needed in order to carry out
the integrals over the space variables.
Next we calculate the variation of the background en-
ergy caused by the stochastic forces of the satellite. It can
be obtained from eq.(35) with µ = B and ν = S, neglecting
the heating term:
B,tot
(cid:28)(cid:18) dEB
dt (cid:19)δt(cid:29) =
= (βS − βB)XiB XjS Z 0
−∞
dsviB ,0C BB
iB jS (s)vjS ,0.
(48)
B,tot
(cid:28)(cid:18) dES
dt (cid:19)δt(cid:29) =
= (βB − βS)XiB XjB Z 0
−∞
dsviB ,0C SS
iB jB (s)vjB ,0.
(44)
Again, the integration over dviB makes it vanish for
a background at rest with an isotropic velocity distribu-
tion function: because no translation energy is available, the
autointeraction results only in a slow heating of the back-
ground.
Regarding the autointeraction energy of the satellite,
eq. (35) gives:
For a background at rest with a Maxwellian velocity
distribution function, this energy rate vanishes when one
performs the integration over dviB . We conclude that the
effect of fluctuating forces of the satellite acting on the back-
ground only heat it, and have no effect on a variation of the
orbital energy of either the satellite or the background.
The total instantaneous energy flow between the satel-
lite and the background is given by the difference between
the l.h.s. of the eqs. (43) and (44) (see eq. (32)). It can be
written in terms of the rate of change of the satellite orbital
and internal energies as:
S,tot
(cid:28)(cid:18) dES
dt (cid:19)δt(cid:29) =
= (βB − βS)XiS XjB Z 0
−∞
dsviS ,0C SS
iS jB (s)vjB ,0.
(49)
The correlation matrix is given by eq. (A10) with
α = S, γ = S and δ = B. Substituting the expression for
C SS
iS jB (s) in eq. (49), and performing the integral over ds
with the help of eq. (38), we obtain:
c(cid:13) 0000 RAS, MNRAS 000, 1 -- 10
′
S,tot
Bm2
N
S
NS
(cid:28)(cid:18) dES
dt (cid:19)δt(cid:29) = G2m2
S(βB − βS)
×XiS XjB "XhS (cid:26) (riS − rhS ) · viS
riS − rhS 3 (cid:27)
×(cid:26) [vjB · (rjB − rhS )][(rjB − rhS ) · (vjB − vhS )]
rjB − rhS vjB − vhS 2 (rjB − rhS )⊥ 2
−
rjB − rhS vjB − vhS 2 (rjB − rhS )⊥ 2
vjB − vhS (rjB − rhS )⊥ 2(cid:27)
vjB · (rjB − rhS )⊥
NSZ drdr′nS(r)nS(r′)
rjB − rhS 2 (vjB − vhS ) · vjB
riS − r 3 rjB − r′ (cid:21) .(50)
[(riS − r) · viS ]
+
+
1
The integrations over velocities can be carried out fol-
lowing the same steps as in the previous section and we
finally have:
σ2
S,tot
σ2
B
σ2
S
mBσ2
B −1(cid:19) G2
SZ driS drjB ρS(riS )ρB(rjB )
T Z drhS ρS(rhS )(cid:26) (riS − rhS ) · vCM S
riS − rhS 3 (cid:27)
CM S − x2
CM S)erf(αCM S )
(cid:28)(cid:18)dES
dt (cid:19)δt(cid:29) =(cid:18)mSσ2
×(cid:20)−
×(cid:26) exp(α2
rjB − rhS
NSZ drdr′ρS(r)nS(r′)
riS − r 3 rjB − r′ (cid:21) (51)
[(riS − r) · vCM S]
(cid:27)
+
1
where αCM S ≡ xCM S ·
(rhS −rjB
rhS −rjB
)
.
4 PARTICULAR LIMITS
4.1 Point-mass satellites
In eq. (43) it is implicitly assumed that the distance between
a generic satellite particle and a generic background parti-
cle cannot be smaller than a scale, dmin, that is, that there
exists a minimum effective impact parameter. As the main
contribution to the integrals appearing in this expression
comes from small rS − rB values (rS and rB are generic
satellite and background particle positions), an accurate de-
termination of the dmin scale is a crucial point when study-
ing dynamical friction. This scale, however, cannot be deter-
mined by the present approach to this problem. There exist
in literature several estimates of dmin for specific situations
(White 1976; Bontekoe & van Albada 1987). White (1976)
in fact shows that for spherical symmetric satellites, it is a
good approximation to consider them as point-like systems
with a cut-off in their distances to background particles,
and that for the case of satellites with a King mass distribu-
tion (King 1966), this cut-off is given by dmin ≃ Rt
5 , where
Rt is the satellite tidal radius. A point mass satellite has a
density profile given by ρS(rS) = MSδ(rS − rCM S), where
δ(rS − rCM S) is the delta function in three dimensions and
rCM S is the position of the satellite center-of-mass. With
this value of ρS(rS), eq. (43) gives in the point mass limit:
c(cid:13) 0000 RAS, MNRAS 000, 1 -- 10
Dynamical friction for compound bodies
7
(cid:28)(cid:18) dES
S
orb
σ2
B
B
mSσ2
S − 1(cid:19) G2M 2
dt (cid:19)δt(cid:29) =(cid:18) mBσ2
×(cid:20)Z drhB ρB(rhB )Θ(rhB )
×(cid:26) σ2
NB Z drdr′ρB(r)nB(r′)Θ(r)Θ(r′)
rCM S − r 3 rCM S − r' (cid:21)
[(rCM S − r) · vCM S ]
CM S − x2
exp(α2
B
σ2
T
×
+
1
(rCM S − rhB ) · vCM S
rCM S − rhB 4
CM S)erf(αCM S ) − 1(cid:27)
(52)
(rCM S−r hB
where now αCM S ≡ xCM S ·
rCM S−r hB
Θ ( rCM S − r −dmin) is the step function.
)
and Θ(r) ≡
4.2 Homogeneous backgrounds. The
Chandrasekhar limit
When the background density is uniform, we can set
ρB(rB) = ρB,0 and rCM S = 0 in eq. (52) and then it be-
comes:
orb
vCM S
SρB,0 ln Λ
(cid:28)(cid:18) dES
S − 1(cid:19) 4πG2M 2
xCM S exp(−x2
dt (cid:19)δt(cid:29) =(cid:18) mBσ2
B
mSσ2
2
√π
×[erf(xCM S) −
This recovers the Chandrasekhar formula for the motion
of a massive test particle in an homogeneous background, ex-
cept for the factor containing the temperature ratio (which
vanishes in the Chandrasekhar formula because mS ≫ mB
in this limit), and, now, σT = (σ2
B)1/2 at the place of
the background velocity dispersion.
S + σ2
CM S)].
(53)
In the Figure 1 we plot the ratio, R, of satellite en-
ergy loss in the Chandrasekhar limit with σS = 0 and with
σS 6= 0, assuming that TB/TS = 0. As can be seen in this
Figure, the effect of a non zero σS increases with increasing
σS and is more important at low vCM S . The Chandrasekhar
formula always overestimates the dynamical friction force,
and in some situations the effect of neglecting the satellite
velocity dispersion could cause an error as high as an order
of magnitude.
5 SUMMARY AND DISCUSSION
We have derived an expression giving the orbital energy
exchange due to dynamical friction experienced by an ex-
tended body, composed of NS bound particles endowed with
a velocity spectrum, as it moves interacting with a non ho-
mogeneous discrete background. It has been assumed that
both, the satellite and the background, have Maxwellian ve-
locity distributions and that the background is static.
Self-interactions of both satellite and background par-
ticles have been taken into account. This results in no effect
on their energy exchange.
Heating terms appear in quite a natural way in our
approach both due to interactions among particles of the
8
R. Dom´ınguez-Tenreiro & M.A. G´omez-Flechoso
sented in this paper, has resulted in the derivation of a for-
mula that takes into account the space and velocity struc-
ture of the satellite. This represents a common situation in
many astrophysical processes and, as we have shown, might
have important quantitative consequences in the setting-up
of timescales for these processes.
ACKNOWLEDGMENTS
We thank Drs. G. Gonz´alez-Casado and E. Salvador-Sol´e
for interesting informations. M.A. G´omez-Flechoso was sup-
ported by the Direcci´on General de Investigaci´on Cient´ıfica
y T´ecnica (DGICYT, Spain) through a fellowship.
The DGICYT also supported in part this work , grants
AEN93-0673 and PB93-0252.
REFERENCES
Bekenstein, J. D. & Maoz, E., 1992, ApJ, 390, 79
Bekenstein, J. D. & Zamir, D., 1991, ApJ, 359, 427
Binney, J. & Tremaine, S., 1987, Galactic Dynamics. Princeton
Univ. Press, Princeton
Bontekoe, T. R. & van Albada, T. S., 1987, MNRAS, 224, 349
Chandrasekhar, S., 1943, ApJ, 97, 255
Cohen, L., 1975, in Hayli, A., ed., Proc. IAU Symp. 60: Dynamics
of Stellar Systems. Reidel, Dordrecht, p. 33
Kalnajs, A. J., 1972, in Lecar, M., ed., Proc. IAU Colloq. 10:
Gravitational N-body Problem. Reidel, Dordrecht, p. 13
Kandrup, H. E., 1980, Phys. Rep., 63, 1
King, I., 1966, AJ, 71, 64
Kubo, R., 1959, in Lectures in Theoretical Physics, Univ. of Col-
orado, vol. 1. Intersciences Publishers, p. 120
Lin, D. N. C. & Tremaine, S., 1983, ApJ, 264, 364
Lynden-Bell, D. & Kalnajs, A., 1972, MNRAS, 157, 1
Mamon, G. A., 1993, in Combes, F. & Athanassoula, E., eds.,
Gravitational Dynamics and the N-body Problem. Meudon,
Paris, p. 188
Maoz, E., 1993, MNRAS, 263, 75
Marochnik, L. S., 1968, SvA, 11, 873
Palmer, P. L. & Papaloizou, J., 1985, MNRAS, 215, 691
Reif, F., 1965, Fundamentals of Statistical and Thermal Physics.
McGraw-Hill, New York
Rosenblunth, M. N., Mc Donald, W. M. & Judd, D. L., 1957,
Phys. Rev., 107, 1
Saslaw, W. C., 1985, Gravitational Physics of Stellar and Galactic
Systems. Cambridge Univ. Press, Cambridge
Tremaine, S. & Weinberg, M. D., 1984, MNRAS, 209, 729
Weinberg, M. D., 1986, ApJ, 300, 93
Weinberg, M. D., 1989, MNRAS, 239, 549
White, S. D. M., 1976, MNRAS, 174, 467
White, S. D. M., 1983, ApJ, 274, 53
APPENDIX A: THE CORRELATION MATRIX
The correlation matrix is defined in eq. (36) with the fluc-
tuating forces given by eq. (10). Each matrix is an object
with four index, and each index takes on two different val-
ues (B and S). This makes sixteen different possibilities,
corresponding to the tensorial product of four different pos-
sibilities for the F α
iγ (t) forces. Writing the Fourier transform
of r − r′ −1 and nα(r) with respect to r, the convolution
theorem and eq. (10) imply that:
c(cid:13) 0000 RAS, MNRAS 000, 1 -- 10
Figure 1. Ratio, R, of satellite energy loss in the Chandrasekhar
limit with σS = 0 and with σS 6= 0, assuming that TB /TS = 0.
Lines correspond to different σS /σB ratios from 0.2 to 2.0 (ratio
increases as the thickness of the line increses)
same kind or of different kind. They are a factor of 1/N
smaller than the orbital effects.
In the point-like satellite limit (or small as compared
with the background size) and constant density background,
we obtain an expression that recovers, in the limit σS/σB →
0 and TS/TB → ∞, Chandrasekhar's dynamical friction for-
mula (eq. (1)). Its comparison with the energy loss given by
eq. (1) allows for a quantification of the effects of having
a non zero σS. It has been found out that the energy loss
is always smaller in this case, and that the difference can
be up to about an order of magnitude for slow satellites as
compared with σB.
In deriving eq. (35) we have considered time intervals,
δt, that are short as compared with the time scale for the
variation of the particle velocities and positions. This al-
lows us to neglect the effects of the smooth gravitational
potential gradients. However, in order to carry out a pre-
cise calculation of dynamical friction, the whole history of
the system from an early enough time and the interactions
along the entire satellite trajectory should have been taken
into account. This would have made the problem extremely
difficult to solve. Instead, taking only a finite δt, means that
interactions with distant particles have not been accurately
considered. Nevertheless, we recall that the contribution of
particles to dynamical friction quickly decreases with dis-
tance, so that this neglect should not result in major conse-
quences.
The bound of δt has also another consequence: this ap-
proach is unable to describe slowly accumulating effects on
dynamical friction (Kalnajs 1972) or the effect of reversible
dynamical feedback (e.g. Tremaine & Weinberg 1984), be-
cause they arise as a consequence of the periodic motion of
the satellite after many revolutions.
Despite these shortcomings,
the
fluctuation-dissipation approach to dynamical friction pre-
the extension of
k2
2π2
Gmαmγ
Z dk
exp[−ik · rgα (t)] −
F α
iγ (riγ , t) = −i
×
Xgα6=iγ
where
nα
k ≡Z drnα(r)exp[−ik · r].
With this definition
k exp[ik · riγ (t)]
N′α
Nα
nα
k
,
(A1)
(A2)
nα
k
Nα
hexp[−ik · riα ]i0 =
and then the average of the stochastic forces, eq. (A1), in
the unperturbed state vanishes, as required.
(A3)
Inserting eq. (A1) in eq. (36), we get:
C αβ
iγ jδ
G2mαmβmγ mδ
4π4
k2
dl
l2
(s) = −
×Z dk
×*
Xgα6=iγ
×
Xhβ6=jδ
k ⊗ l exp[i(k · riγ (0) + l · rjδ (s))]
exp[−ik · rgα (0)] −
exp[−il · rhβ (s)] −
N′α
Nα
nα
k
l
+
0
N′β
Nβ
nβ
(A4)
(A5)
(A6)
and eq. (A3) now gives:
G2mαmβmγ mδ
C αβ
iγ jδ
k2
4π4
dl
l2
(s) = −
×Z dk
×
Xgα6=iγ Xhβ6=jδ
k ⊗ l exp[i(k · riγ (0) + l · rjδ (s))]
l
where
Ekl(gα, hβ, s) ≡ hexp{−i[k · rgα (0) + l · rhβ (s)]}i0.
Ekl(gα, hβ, s) −
N′α
Nα
N′β
Nβ
nα
k nβ
It is assumed that particles are uncorrelated in the un-
perturbed state. Then, if α 6= β, necessarily gα 6= hβ and the
average in eq. (A6) factorizes, as corresponding to uncorre-
lated variables, giving C αβ
(s) = 0. The average also fac-
iγ jδ
torizes for different particles belonging to the same particle
class, i.e., when α = β but gα 6= hα. There are N′α(N′α − 1)
such terms, and each of them has the value
. When
gα = hα, that is, when we consider the same particle, the
average does not factorize anymore. Taking this considera-
tions into account, the correlation matrix becomes:
k nα
nα
N 2
α
l
Dynamical friction for compound bodies
9
The position of a generic particle at time s satisfying
(19) can be written as:
rjα (s) = rjα ,0 + vjα ,0s
(A8)
An integration of eq. (11) would give four more terms
corresponding to the gravitational acceleration and the fluc-
tuating forces. The acceleration terms are negligible when
compared with vjα,0s (see eq. (19)) and the fluctuating
forces would give rise to third order terms.
Once this expression for the particle positions at time s
is substituted in eq. (A7), the integrals over dk and dl can
be easily calculated taking the gradient of the Fourier rep-
resentation for the Green's function of the Laplace equation
in three dimensions:
Z dk
k2
k exp[ik · A] = i
2πA
A3 .
(A9)
Finally, combining eqs. (A2), (A6), (A7), (A8) and (A9)
the correlation matrix reads:
C αα
iγ jδ (s) = G2m2
αmγ mδ
N′α
Nα
[rjδ − rhα + s(vjδ − vhα )]
rjδ − rhα + s(vjδ − vhα )3
1
(riγ − rhα )
riγ − rhα 3 ⊗
×"Xhα
Nα Z drdr′nα(r)nα(r′)
(riγ − r)
riγ − r 3 ⊗
−
×
The 0 subindex have been dropped from the r and v
(rjδ − r′ + svjδ )
rjδ − r′ + svjδ 3(cid:21) .
(A10)
vectors.
APPENDIX B:
In this appendix we calculate the integral
I =Z dvf S
M (v) exp(cid:2)γ2 − y2(cid:3) erf(γ),
where v is the velocity of class S particles, that is
v = vCM S + u,
(B1)
(B2)
whose distribution function, f S
in u
M , is a Maxwellian isotropic
f S
M (v) =
1
(2πσ2
S)3/2 exp(cid:20)−
(v − vCM S)2
2σ2
S
(cid:21) .
In the last equation y ≡ v/√2σB, γ ≡ y · ζ, and ζ is a
unit constant vector. Expression (B2) implies:
(B3)
(B4)
γ ≡ γCM S + γu
where
γCM S =
vCM S√2σB
ζ
and
γu =
uCM S√2σB
ζ
G2m2
αmγmδ
4π4
C αα
iγ jδ (s) = −
×Z dk
×hN′αEkl(gα = hα, s) − N′α
dl
l2
k2
c(cid:13) 0000 RAS, MNRAS 000, 1 -- 10
k ⊗ l exp[i(k · riγ (0) + l · rjδ (s))]
nα
k nα
l
N 2
α i .
(A7)
The definition of the error function and equation (B4)
imply that:
10
R. Dom´ınguez-Tenreiro & M.A. G´omez-Flechoso
erf(γ) = erf(γCM S) +
2
√π Z γu
0
dt exp(cid:2)−(t + γCM S)2(cid:3) . (B5)
We now write I (eq. (B1)) as an integral with respect
to the relative velocities u. Taking ζ = (0, 0, 1) we obtain:
I =
CM S(cid:1)
CM S−x2
(2πσ2
S)3/2
Z dux exp(cid:2)− (Aux +Bx)2(cid:3)
exp(cid:0)α2
×Z duy exp(cid:2)− (Auy +By)2(cid:3)Z duz exp"−(cid:18) uz√2σS(cid:19)2#
dt exp(cid:2)− (t + γCM S)2(cid:3)(cid:21) (B6)
×(cid:20)erf(γCM S) +
where xCM S ≡ vCM S /√2σT , αCM S ≡ xCM S · ζ, σ2
T ≡
S, A ≡ σT /√2σSσB, Bi ≡ σS
σ2
B + σ2
xCM S,i, with i = x, y,
and γu = uz/√2πσB. The integrals over dux and duy can
be easily performed and they give:
√π Z γu
σB
2
0
(cid:20)√π
A
erf(∞)(cid:21)2
=
π
A2
The integral can now be written as:
σ2
H
σ2
2
π
1
I =
where
I(a, b) ≡
CM S − x2
I(a, b)i exp(cid:2)α2
a(cid:17)2(cid:21)Z S
T herf(γCM S) +
aZ ∞
CM S(cid:3) (B7)
dt exp(cid:2)−(t + b)2(cid:3) (B8)
with a ≡ σS/σH , b ≡ γCM S, comes from the integration on
duz.
The I(a, b) integral can be evaluated as follows: first, we
derive it with respect to b, the we carry out the integration
with respect to t, obtaining:
ds exp(cid:20)−(cid:16) s
−∞
0
∂I(a, b)
∂ b
=√π(cid:20)
1
(1 + a2)1/2 exp(cid:18)−
b2
1 + a2(cid:19)−exp(−b2)(cid:21) (B9)
and, finally, an integration on b leads to:
I(a, b) = I(a, 0) +
π
2 (cid:20)erf(cid:20)
b
(1 + a2)1/2(cid:21) − erf(b)(cid:21)
(B10)
with I(a, 0) = 0 because erf(s) = −erf(−s). Substituting the
expression for I(a, b) in eq. (B7), we get:
I =
σ2
B
σ2
T
CM S − x2
exp(cid:0)α2
CM S(cid:1) erf(αCM S)
where now xCM S ≡ vCM S/√2σT and αCM S ≡ xCM S · ζ.
(B11)
c(cid:13) 0000 RAS, MNRAS 000, 1 -- 10
|
astro-ph/0211420 | 1 | 0211 | 2002-11-19T12:28:47 | Star Cluster Formation and Disruption Time-Scales - II. Evolution of the Star Cluster System in M82's Fossil Starburst | [
"astro-ph"
] | ABRIDGED: We obtain new age and mass estimates for the star clusters in M82's fossil starburst region B, based on improved fitting methods. Our new age estimates confirm the peak in the age histogram attributed to the last tidal encounter with M81; we find a peak formation epoch at slightly older ages than previously published, log(t_peak / yr) = 9.04, with a Gaussian sigma of Delta log(t_width) = 0.273. Cluster disruption has removed a large fraction of the older clusters. Adopting the expression for the cluster disruption time-scale of t_dis(M)= t_dis^4 (M/10^4 Msun)^gamma with gamma = 0.62 (Paper I), we find that the ratios between the real cluster formation rates in the pre-burst phase (log(t/yr) <= 9.4), the burst-phase (8.4 < log(t/yr) < 9.4) and the post-burst phase (log(t/yr) <= 8.4) are about 1:2:1/40. The mass distribution of the clusters formed during the burst shows a turnover at log(M_cl/Msun) ~ 5.3 which is not caused by selection effects. This distribution can be explained by cluster formation with an initial power-law mass function of slope alpha=2 up to a maximum cluster mass of M_max = 3 x 10^6 Msun, and cluster disruption with a normalisation time-scale t_dis^4 / t_burst = (3.0 +/- 0.3) x 10^{-2}. For a burst age of 1 x 10^9 yr, we find that the disruption time-scale of a cluster of 10^4 Msun is t_dis^4 ~ 3 x 10^7 years, with an uncertainty of approximately a factor of two. This is the shortest disruption time-scale known in any galaxy. | astro-ph | astro-ph | Mon. Not. R. Astron. Soc. 000, 000 -- 000 (2001)
Printed 27 November 2018
(MN LATEX style file v1.4)
Star Cluster Formation and Disruption Time-Scales -- II.
Evolution of the Star Cluster System in M82's Fossil
Starburst
Richard de Grijs1⋆, Nate Bastian2, and Henny J.G.L.M. Lamers2
1 Institute of Astronomy, University of Cambridge, Madingley Road, Cambridge CB3 0HA
2 Astronomical Institute, Utrecht University, Princetonplein 5, 3584 CC Utrecht, The Netherlands
Received date; accepted date
ABSTRACT
We obtain new age and mass estimates for the star clusters in M82's fossil starburst
region B, based on improved fitting methods. Our new age estimates confirm the peak
in the age histogram attributed to the last tidal encounter with M81; we find a peak
formation epoch at slightly older ages than previously published, log(tpeak/yr) = 9.04,
with a Gaussian σ of ∆ log(twidth) = 0.273. The actual duration of the burst of clus-
ter formation may have been shorter because uncertainties in the age determinations
may have broadened the peak. Our improved mass estimates confirm that the (initial)
masses of the M82 B clusters with V ≤ 22.5 mag are mostly in the range 104 − 106M⊙,
with a median mass of Mcl = 1.08 × 105M⊙. The formation history of the observed
clusters shows a steady decrease towards older ages. This indicates that cluster dis-
ruption has removed a large fraction of the older clusters.
Adopting the expression for
tdis(M ) =
tdis
4 (M/104M⊙)γ with γ ≃ 0.62 (Paper I), we find that the ratios between the real
cluster formation rates in the pre-burst phase (log(t/yr) ≥ 9.4), the burst-phase
(8.4 < log(t/yr) < 9.4) and the post-burst phase (log(t/yr) ≤ 8.4) are about 1 : 2 : 1
40 .
The formation rate during the burst may have been higher if the actual duration of
the burst was shorter than adopted.
The mass distribution of the clusters formed during the burst shows a turnover at
log(Mcl/M⊙) ≃ 5.3 which is not caused by selection effects. This distribution can be
explained by cluster formation with an initial power-law mass function of slope α = 2
up to a maximum cluster mass of Mmax = 3 × 106M⊙, and cluster disruption with a
4 /tburst = (3.0 ± 0.3) × 10−2. For a burst age of 1 × 109 yr,
normalisation time-scale tdis
we find that the disruption time-scale of a cluster of 104 M⊙ is tdis
4 ∼ 3 × 107 years,
with an uncertainty of approximately a factor of two. This is the shortest disruption
time-scale known in any galaxy.
the cluster disruption time-scale of
Key words: diffusion -- galaxies: individual: M82 -- galaxies: starburst -- galaxies:
star clusters
2
0
0
2
v
o
N
9
1
1
v
0
2
4
1
1
2
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
1
INTRODUCTION
1.1 Multiple starbursts in M82
M82 is often regarded as the "prototype" starburst galaxy.
Observations over the entire wavelength range, from radio
waves to X-rays (see, e.g., Telesco 1988 and Rieke et al.
1993 for reviews), seem to support a scenario in which tidal
interactions, predominantly with its large neighbour M81,
triggered intense star formation in the centre of this small
⋆ E-mail: [email protected]
c(cid:13) 2001 RAS
irregular galaxy during the last several 100 Myr. The result-
ing starburst, with the high star formation rate of ∼ 10M⊙
yr−1, has continued up to about 50 Myr ago (e.g., O'Connell
& Mangano 1978, Rieke et al. 1993). Energy and gas ejec-
tion from supernovae and combined stellar winds drive a
large-scale galactic wind along the minor axis of M82 (e.g.,
Lynds & Sandage 1963, McCarthy et al. 1987, Shopbell &
Bland-Hawthorn 1998).
All of the bright radio and infrared sources associated
with the active starburst are found in a small region within a
radius of ∼ 250 pc from the galaxy's centre, but most of this
volume is heavily obscured by dust at optical wavelengths.
2
R. de Grijs, N. Bastian, and H.J.G.L.M. Lamers
However, there is now compelling evidence that the ac-
tive starburst was not the only major starburst event to
have occurred in M82. A region at ∼ 400 − 1000 pc from the
centre, "M82 B" (nomenclature from O'Connell & Mangano
1978), has the high surface brightness and spectral features
expected for an ancient starburst with an age in excess of
several 100 Myr and an amplitude similar to the active burst
(O'Connell & Mangano 1978, Marcum & O'Connell 1996, de
Grijs et al. 2001, hereafter dGOG). Its spectral features re-
semble the characteristics of the anomalous "E+A" spectra
found in distant galaxy clusters, which are generally inter-
preted as the signature of a truncated burst of star forma-
tion that occurred 100 − 1000 Myr earlier (e.g., Couch &
Sharples 1987, Dressler & Gunn 1990, Couch et al. 1998).
It is thought that these starbursts are closely related to the
process by which disc galaxies are converted into elliptical or
lenticular galaxies and they are thought to result from tidal
interactions, mergers, or perhaps ram-pressure stripping by
the intergalactic medium (Butcher & Oemler 1978, Oemler
1992, Barger et al. 1996).
The significance of the detailed study of M82's star-
burst environment lies, therefore, in the broader context of
galaxy evolution. Starbursts of this scale are likely to be
common features of early galaxy evolution, and M82 is the
nearest analogue to the sample of star-forming galaxies re-
cently identified at high redshifts (z & 3; Steidel et al. 1996,
Giavalisco et al. 1997, Lowenthal et al. 1997). In addition,
M82 affords a close-up view not only of an active starburst
-- in M82 A, C and E -- but also, in region B and elsewhere,
of the multiple post-burst phases.
In a recent study focusing on the fossil starburst site,
M82 B, we found a large population of (∼ 110) evolved com-
pact star clusters (dGOG), whose properties appear to be
consistent with them being evolved (and therefore faded)
counterparts of the young (super) star clusters detected in
the galaxy's active core (O'Connell et al. 1995). Based on
broad-band optical and near-infrared colours and compari-
son with stellar evolutionary synthesis models, we estimated
ages for the M82 B cluster population from ∼ 30 Myr to over
10 Gyr, with a peak near 650 Myr (see also de Grijs 2001).
About 22 per cent of the clusters in M82 B are older than
2 Gyr, with a roughly flat distribution to over 10 Gyr. Very
few clusters are younger than 300 Myr. The full-width of
the peak at 650 Myr is ∼ 500 Myr, but this is undoubt-
edly broadened by the various uncertainties entering the
age-dating process. The selection bias of the star clusters
in dGOG is such that the truncation of cluster formation
for ages < 300 Myr is better established than the roughly
constant formation rate at ages > 2 Gyr.
Thus, we suggested steady, continuing cluster formation
in M82 B at a very modest rate at early times (> 2 Gyr ago)
followed by a concentrated formation episode lasting from
400 -- 1000 Myr ago and a subsequent suppression of cluster
formation (dGOG).
1.2 Cluster disruption time-scales
the star cluster system in M82's fossil starburst region, we
determine both its cluster formation history and the char-
acteristic disruption time-scales.
The dynamical evolution of star clusters is determined
by a combination of internal and external time-scales. The
free-fall and two-body relaxation time-scales, which depend
explicitly on the initial cluster mass density (e.g., Spitzer
1957, Chernoff & Weinberg 1990, de la Fuente Marcos 1997,
Portegies Zwart et al. 2001), affect the cluster-internal pro-
cesses of star formation and mass redistribution through en-
ergy equipartition, leading to mass segregation and, even-
tually, core collapse (see, e.g., de Grijs et al. [2002a,b] for
a detailed description of mass segregation effects in young
star clusters). While the internal relaxation process will, over
time, eject both high-mass stars from the core (e.g., due to
interactions with hard binaries; see Brandl et al. 2001, de
Grijs et al. 2002a) and lose lower-mass stars from its halo
through diffusion (e.g., due to Roche-lobe overflow), the ex-
ternal process of tidal disruption and stripping by the sur-
rounding galactic field is in general more important for the
discussion of the disruption of star clusters.
Tidal disruptive processes are enhanced by "normal"
stellar evolutionary effects such as mass loss by winds and/or
supernova explosions, which will further reduce the stellar
density in a cluster, and thus make it more sensitive to ex-
ternal tidal forces.
From the bimodal age distribution of (young) open and
(old) globular clusters in the Milky Way, Oort (1957) con-
cluded that disruption of Galactic star clusters must occur
on time-scales of ∼ 5×108 yr. Around the same time, Spitzer
(1957) derived an expression for the disruption time scale
as a function of a cluster's mean density, ρc (M⊙ pc−3):
tdis = 1.9 × 108ρc yr, for 2.2 < ρc < 22M⊙ pc−3. More
advanced recent studies, based on N-body modeling, have
shown that the cluster disruption time-scale is sensitive to
the cluster mass, the fraction of binary (or multiple) stars,
and the initial mass function (IMF) adopted (e.g., Chernoff
& Weinberg 1990, de la Fuente Marcos 1997).
Boutloukos & Lamers (2001, 2002) derived an empiri-
cal relation between the disruption time and a cluster's ini-
tial mass for the Milky Way, the Small Magellanic Cloud
(SMC), M33 and the inner spiral arms of M51. They showed,
based on an analysis of the mass and age distributions of
magnitude-limited samples of clusters, that the empirical
disruption time of clusters depends on their initial mass (Mi)
as
tdis = tdis
4 (Mi/104M⊙)γ
(1)
4
where tdis
is the disruption time of a cluster of initial mass
Mi = 104M⊙. The value of γ is approximately the same
in these four galaxies, γ = 0.62 ± 0.06. However, the char-
acteristic disruption time-scale tdis
is widely different in the
different galaxies. The disruption time-scale is longest in the
SMC (∼ 8 Gyr) and shortest in the inner spiral arms of M51
(∼ 40 Myr). We will derive the disruption time-scale of clus-
ters in M82 B and allow for disruption in the determination
of the cluster formation history.
4
In the determination of the cluster formation history from
the age distribution of magnitude-limited cluster samples,
cluster disruption must be taken into account. This is be-
cause the observed age distribution is that of the surviving
clusters only. Therefore, in this study of the evolution of
In Section 2, we will summarise the observations on
which our discussion is based, and in Section 3 we ob-
tain new age and mass estimates for the clusters in M82
B, based on the 3/2DEF method (3/2-dimensional energy-
fitting method; Bik et al. 2002) of fitting the observed spec-
c(cid:13) 2001 RAS, MNRAS 000, 000 -- 000
Star Cluster Formation and Disruption Time-Scales -- II.
3
tral energy distribution to that of cluster evolution models.
In Section 4 we derive the cluster formation history and the
characteristic cluster disruption time-scale. In Section 5 we
discuss the mass distribution of the clusters formed during
the burst. The methods and assumptions are discussed in
Section 6; finally, we will summarise and conclude the paper
in Section 7.
2 OBSERVATIONS
The observations on which our analysis of the M82 clus-
ter formation history is based were described in detail in
dGOG. Briefly summarised, we observed region B with the
Wide Field Planetary Camera 2 (WFPC2) on board the
Hubble Space Telescope (HST) through the F439W, F555W
and F814W filters, roughly corresponding to the standard
Johnson-Cousins B, V and I passbands, respectively. We
obtained WFPC2 observations using two pointings, so that
the entire region B was covered by the Planetary Camera
(PC) chip, the highest-resolution optical
imaging instru-
ment available on board HST at that time, with a pixel
size of 0.0455′′. The effective integration times used for the
F439W, F555W, and F814W observations were 4400s, 2500s
and 2200s, respectively, for the western half of M82 B and
4100s, 3100s and 2200s, respectively, for the eastern half
closest to the galaxy's starburst core.
In addition, we imaged the entire region with HST's
Near-Infrared Camera and Multi-Object Spectrometer
(NICMOS) (Camera 2; pixel size 0.075′′) in both the F110W
and F160W filters (comparable to the Bessell J and H filters,
respectively), in a tiled pattern of 2×4 partially overlapping
exposures. The integrations, with effective integration times
of 768s for each of the eastern and western halfs of M82
B and each filter, were taken in MULTIACCUM mode to
preserve dynamic range and to correct for cosmic rays.
We subsequently obtained integrated photometry for
the extended objects (i.e., star clusters) in M82 B down to
a 50 per cent completeness limit of V ≃ 23.3 mag. Because
of the highly variable background and the numerous spuri-
ous features due to dust lanes and background variations,
we decided to select only genuine star cluster brighter than
V = 22.5 for further analysis, corresponding to close to 100
per cent completeness (dGOG).
3 OBSERVED CLUSTER MASS AND AGE
DISTRIBUTIONS
For the final 113 objects obtained following the procedures
outlined in the previous section, dGOG used (B − V ) versus
(V − I) colour-colour diagrams to disentangle age and ex-
tinction effects, since the age and extinction vectors are not
entirely degenerate for this choice of optical colours.
Accurate age and mass determinations are essential for
the analysis of cluster formation and disruption time-scales
performed in this paper. Combining the observed luminosi-
ties of the M82 B star cluster population with the appropri-
ate, age-dependent mass-to-light ratios from spectral syn-
thesis models provided us with photometric mass estimates
with an accuracy of well within an order of magnitude. In-
dependent dynamical mass estimates from high-resolution
c(cid:13) 2001 RAS, MNRAS 000, 000 -- 000
spectroscopy are available only for a few of the most lumi-
nous super star clusters seen in the nearest starburst galax-
ies (M82, the Antennae, NGC 1569 and NGC 1705), and are
approximately 106M⊙ (Ho & Filippenko 1996a,b, Smith &
Gallagher 2000, Mengel et al. 2002). Because of the prox-
imity of M82, we have been able to probe the young cluster
population in M82 B to fainter absolute magnitudes, and
thus lower masses, than has been possible before in other
starburst galaxies. Other young star cluster samples are bi-
ased towards high masses by selection effects due to their
host galaxies' greater distances.
Keeping the importance of the age dependence on our
photometric mass estimates in mind, we decided to redeter-
mine the M82 B cluster ages using the full parameter space
available. We obtained more accurate age determinations by
also including the NICMOS observations, and solving for the
best-fitting ages matching the full spectral energy distribu-
tions (SEDs) of our sample clusters, from the F439W to the
F160W passband. We compared the observed cluster SEDs
with the model predictions for an instantaneous burst of star
formation, assuming a Salpeter IMF from 0.1 − 100M⊙ with
power-law slope αIMF = 2.35 to obtain our new estimates for
the cluster age t, initial mass Mcl and extinction E(B − V );
for the latter we adopted the (Galactic) extinction law of
Scuderi et al. (1996). For ages t ≤ 109 yr we used the Star-
burst99 models (Leitherer et al. 1999), while for older ages
we used the most recent single stellar population models by
Bruzual & Charlot (2000, hereafter BC00). Examples of our
model fits to the observed cluster SEDs are shown in Fig. 1.
We realise that recent determinations of the stellar
IMF deviate significantly from a Salpeter-type IMF at low
masses, in the sense that the low-mass stellar IMF is sig-
nificantly flatter than the Salpeter slope. The implication
of using a Salpeter-type IMF for our cluster mass deter-
minations is therefore that we have overestimated the indi-
vidual cluster masses (although the relative mass distribu-
tion of our entire cluster sample remains unaffected). There-
fore, we used the more modern IMF of Kroupa, Tout &
Gilmore (1993, hereafter KTG) to determine the correc-
tion factor, C, between our masses and the more realistic
masses obtained from the KTG IMF (both normalised at
1.0M⊙). This IMF is characterised by slopes of α = −2.7
for m > 1.0M⊙, α = −2.2 for 0.5 ≤ m/M⊙ ≤ 1.0, and
−1.85 < α < −0.70 for 0.08 < m/M⊙ ≤ 0.5. Depending on
the adopted slope for the lowest mass range, we have there-
fore overestimated our individual cluster masses by a factor
of 1.70 < C < 3.46 for an IMF containing stellar masses in
the range 0.1 ≤ m/M⊙ ≤ 100.
The fitting of the observed cluster SEDs to the
Starburst99 and BC00 models was done using a three-
dimensional maximum likelihood method, 3/2DEF, with the
initial mass Mi, age and extinction E(B − V ) as free param-
eters (see Bik et al. 2002). For the 46 clusters with upper
limits in one or more filters we used a two-dimensional maxi-
mum likelihood fit, using the extinction probability distribu-
tion for E(B −V ). This distribution was derived for the clus-
ters with well-defined SEDs over the full wavelength range
(see Bik et al. 2002). We obtained reliable age estimates
based on the full SED modeling for 81 of the 113 clusters.
The new age and mass estimates for these 81 clusters are
listed in Table 1, where the individual clusters are identified
by their dGOG ID. The dGOG mass estimates in Table 1
4
R. de Grijs, N. Bastian, and H.J.G.L.M. Lamers
Table 1. New age, mass and extinction determinations for the M82 B clusters, and comparison with previously published values.
Object
log(Age/yr)
This paper
log(m/M⊙ )a
E(B − V ) (mag)
log(Age/yr)
log(m/M⊙ )
dGOG
min
best
max
min
best max
min
best
max
E(B − V )
(mag)b
B1-01
B1-02
B1-04
B1-05
B1-06
B1-07
B1-08
B1-09
B1-10
B1-11
B1-12
B1-14
B1-16
B1-17
B1-18
B1-20
B1-22
B1-24
B1-25
B1-26
B1-27
B1-28
B1-29
B1-32
B1-34
B1-35
B1-36
B1-37
B1-38
B1-39
B1-40
B1-43
B2-01
B2-04
B2-05
B2-07
B2-08
B2-12
B2-13
B2-14
B2-17
B2-18
B2-21
B2-22
B2-25
B2-26
B2-28
B2-29
B2-30
B2-31
B2-32
B2-33
B2-34
B2-36
B2-37
B2-38
B2-39
B2-40
B2-41
B2-43
B2-44
B2-45
B2-47
B2-48
B2-50
B2-51
B2-52
B2-54
B2-55
B2-56
B2-57
B2-59
B2-60
B2-62
B2-64
B2-65
B2-67
B2-68
B2-69
B2-70
9.16
8.96
5.10
9.74
5.10
8.96
8.51
5.10
5.10
8.71
8.71
8.51
8.56
8.66
8.86
9.01
8.56
8.66
8.91
5.10
8.76
8.86
5.10
8.76
5.10
8.61
8.96
9.16
5.10
8.81
5.10
9.01
9.01
9.01
5.10
8.66
8.66
9.74
9.30
8.71
8.96
5.10
6.86
8.96
5.10
8.71
5.10
5.10
8.41
5.10
8.81
5.10
5.10
9.16
6.84
5.10
8.71
8.41
8.96
8.66
5.10
8.71
9.01
5.10
5.10
8.96
8.96
5.10
8.56
8.86
8.86
5.10
5.10
8.76
8.96
8.91
8.81
8.86
8.86
8.76
9.88
9.28
8.46
10.05
6.58
9.01
8.66
8.91
8.71
8.86
9.01
8.86
9.95
9.01
8.91
9.16
8.76
8.86
9.01
8.61
8.86
8.86
8.76
9.01
6.84
8.86
10.14
9.40
8.81
9.34
8.66
9.48
9.94
9.23
8.46
8.96
9.01
9.76
9.83
9.44
9.48
7.49
7.00
10.20
8.51
8.76
7.49
6.66
8.71
8.51
9.16
8.71
6.48
9.16
7.14
8.76
8.76
8.46
9.01
8.86
8.66
9.01
9.26
8.61
7.68
10.14
9.01
8.66
8.71
9.26
9.23
8.21
8.76
8.96
9.23
9.11
8.91
9.16
8.96
8.81
10.05
9.36
8.81
10.15
8.76
9.16
8.66
9.01
8.86
8.86
9.21
9.01
10.30
9.16
8.91
9.16
8.86
8.91
9.26
8.66
8.91
8.86
8.91
9.01
8.36
9.01
10.30
9.44
8.96
9.44
8.76
9.68
10.25
9.28
8.71
9.41
10.26
9.80
9.90
10.14
9.80
8.66
7.60
10.30
8.81
8.76
8.36
8.71
9.01
8.71
9.60
9.01
8.71
9.21
7.65
8.96
8.81
8.56
9.01
8.91
8.76
9.70
9.32
8.81
8.51
10.30
9.01
8.71
8.76
9.36
9.41
8.96
8.86
9.01
9.30
9.16
8.96
9.30
8.96
8.81
5.20
4.66
4.14
5.76
4.33
5.53
5.14
3.14
3.14
5.36
4.81
5.11
4.94
4.47
5.54
5.48
5.23
5.15
5.00
3.09
4.44
5.97
3.52
4.64
3.31
5.02
5.13
5.75
3.57
4.29
2.79
4.35
4.43
5.23
4.14
5.41
5.49
6.86
5.38
4.75
5.21
4.09
4.46
4.53
3.90
5.60
3.51
2.95
4.99
3.21
5.05
3.03
2.97
5.32
4.42
3.08
4.56
5.07
5.91
5.15
3.57
5.00
4.12
3.32
3.54
5.13
4.66
3.39
3.87
4.76
5.21
3.69
3.46
4.05
4.25
5.02
5.22
4.51
4.09
4.97
5.62
4.90
5.21
5.95
4.53
5.60
5.24
4.24
4.33
5.38
4.88
5.24
5.63
4.56
5.56
5.65
5.29
5.20
5.10
4.26
4.49
5.97
4.71
4.69
3.39
5.08
5.90
5.88
4.63
4.58
3.97
4.69
5.09
5.44
5.21
5.55
5.60
6.88
5.71
5.13
5.56
4.79
4.57
5.36
5.01
5.62
4.09
3.21
5.12
4.29
5.23
4.27
3.32
5.34
4.70
4.34
4.57
5.08
5.92
5.20
4.61
5.09
4.36
4.52
4.24
5.90
4.67
4.44
3.94
5.00
5.42
4.73
4.48
4.12
4.48
5.12
5.25
4.69
4.14
4.99
5.74
4.98
5.34
6.03
5.54
5.69
5.24
4.34
4.41
5.41
5.02
5.28
5.87
4.68
5.56
5.65
5.33
5.23
5.23
4.32
4.51
5.97
4.78
4.72
4.44
5.12
6.01
5.91
4.71
4.70
4.03
4.87
5.34
5.47
5.35
5.78
6.35
6.91
5.77
5.63
5.81
5.44
5.17
5.47
5.14
5.62
4.65
4.22
5.19
4.39
5.50
4.42
4.19
5.36
5.11
4.52
4.62
5.14
5.92
5.23
4.64
5.51
4.42
4.61
4.69
6.01
4.71
4.47
3.98
5.10
5.55
5.00
4.54
4.14
4.55
5.20
5.27
4.77
4.16
4.99
0.00
0.00
0.06
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.04
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.08
0.00
0.00
0.00
0.02
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.04
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.06
0.00
0.00
0.02
0.02
0.00
0.18
0.00
0.52
0.06
0.02
0.00
0.04
0.00
0.06
0.10
0.04
0.04
0.00
0.00
0.04
0.02
0.10
0.00
0.02
0.04
0.06
0.02
0.12
0.08
0.00
0.00
0.04
0.00
0.02
0.00
0.00
0.02
0.08
0.18
0.32
0.00
0.00
0.08
0.02
0.26
0.04
0.00
0.10
0.02
0.00
0.44
0.16
0.06
0.12
0.12
0.54
0.02
0.00
0.00
0.00
0.02
0.04
0.02
0.04
0.20
0.00
0.06
0.06
0.00
0.00
0.02
0.00
0.00
0.04
0.24
0.02
0.02
0.00
0.00
0.08
0.04
0.00
0.02
0.20
0.16
0.70
0.06
0.60
0.08
0.04
0.72
0.64
0.08
0.22
0.22
0.50
0.24
0.02
0.08
0.12
0.12
0.16
0.56
0.08
0.04
0.68
0.14
0.46
0.20
0.34
0.10
0.68
0.28
0.60
0.20
0.28
0.14
0.62
0.34
0.50
0.00
0.12
0.40
0.26
0.74
0.24
0.36
0.64
0.04
0.48
0.58
0.24
0.60
0.28
0.78
0.58
0.02
0.16
0.70
0.04
0.04
0.06
0.12
0.60
0.36
0.12
0.64
0.52
0.34
0.06
0.58
0.04
0.22
0.22
0.72
0.64
0.12
0.14
0.12
0.12
0.18
0.04
0.04
10.24
9.27
8.78
10.14
8.67
9.03
8.68
8.89
8.76
8.84
9.03
8.89
10.00
9.03
8.85
9.10
8.78
8.83
9.05
8.68
8.86
8.86
8.81
8.98
7.55
8.88
10.14
9.32
8.83
9.28
8.68
9.70
10.00
9.32
8.66
9.19
9.72
9.70
9.76
10.13
9.46
9.23
8.90
10.26
8.71
8.77
7.26
8.71
8.89
8.65
9.83
8.83
8.56
9.20
7.51
8.83
8.83
8.51
9.00
8.81
8.70
9.93
9.28
8.68
8.31
8.63
9.01
8.68
8.78
9.23
9.26
8.75
8.78
8.95
9.26
9.03
8.90
9.20
8.85
8.81
5.92
4.90
5.25
6.04
5.46
5.53
5.24
4.21
4.33
5.36
4.81
5.13
5.63
4.50
5.50
5.60
5.26
5.15
4.97
4.33
4.46
5.90
4.68
4.65
3.94
4.98
5.91
5.80
4.60
4.53
3.97
4.90
5.16
5.51
5.28
5.56
5.89
6.83
5.66
5.64
5.52
5.69
5.82
5.39
5.05
5.63
3.91
4.21
5.05
4.34
5.70
4.23
4.08
5.38
4.97
4.42
4.65
5.10
5.86
4.54
4.61
5.70
4.39
4.50
4.57
3.90
4.67
4.44
4.03
4.98
5.40
4.82
4.49
4.09
4.52
5.01
5.11
4.70
4.04
4.97
0.31
0.04
0.58
0.18
0.44
0.12
0.09
0.00
0.18
0.09
0.06
0.34
0.40
0.17
0.15
0.07
0.25
0.13
0.16
0.00
0.00
0.13
0.25
0.10
0.35
0.24
0.24
0.18
0.16
0.15
0.13
0.11
0.00
0.15
0.32
0.43
0.77
0.22
0.19
0.43
0.26
1.03
1.13
0.15
0.29
0.14
0.60
0.25
0.37
0.24
0.37
0.24
0.29
0.00
0.69
0.00
0.00
0.13
0.20
0.00
0.12
0.37
0.00
0.24
0.48
0.00
0.00
0.00
0.00
0.00
0.26
0.65
0.18
0.00
0.00
0.20
0.35
0.07
0.00
0.03
Notes: a based on a Salpeter-type IMF; comparison with results from more modern IMFs (e.g., KTG) implies that we have overestimated our individual
cluster masses by a factor of about 1.7 -- 3.5, depending on the adopted IMF slope for the lowest stellar masses (see Sect. 3); b E(B − V ) = AV /3.1.
c(cid:13) 2001 RAS, MNRAS 000, 000 -- 000
Star Cluster Formation and Disruption Time-Scales -- II.
5
(a)
(c)
(e)
(b)
(d)
(f)
Figure 1. Examples of a few fits of the Starburst99 and BC00 models (dashed lines) to the observed cluster SEDs sampled in the
B, V, I, J, and H passbands (data points with error bars), representative of clusters throughout the entire age range (age increases from
panel a to f). The best-fitting model SEDs are based on the 3/2DEF maximum likelihood method.
were determined by combining the absolute magnitudes of
the clusters with their age-dependent mass-to-light ratios.
Our analysis in both dGOG and the present paper is
based on the assumption of solar metallicity, which should
be a reasonable match to the young objects in M82 (Gal-
lagher & Smith 1999; see also Fritze-v. Alvensleben & Ger-
hard 1994), but the effects of varying the metallicity between
one-fifth and 2.5 times solar are small compared to the pho-
tometric uncertainties (see also dGOG). We therefore assert
that our results are not significantly affected by possibly
varying metallicities. Ongoing analysis (Parmentier, de Grijs
& Gilmore, 2002) suggests that this assumption was indeed
justified. We should caution, however, that we have only
been able to sample the star clusters located close to the
surface of region B, as evidenced by the derived extinction
estimates: dGOG find AV . 1 mag in general, while our
new determinations (this paper) restrict the extinction even
more, to AV . 0.2 mag for the clusters with well-determined
ages, and possibly up to AV ∼ 0.40 − 0.55 for some of the
others.
In Fig. 2 we compare the age and mass estimates from
the more sophisticated 3/2DEF method used in this pa-
per with the corresponding parameters obtained by dGOG
based on their location in the (B − V ) vs. (V − I) di-
c(cid:13) 2001 RAS, MNRAS 000, 000 -- 000
6
R. de Grijs, N. Bastian, and H.J.G.L.M. Lamers
agrams. The black dots represent clusters for which the
range between the minimum and maximum plausible ages,
log(Age[max])−log(Age[min]) ≤ 1.0; the open circles are ob-
jects with more uncertain age determinations. We conclude
that the new age and mass determinations are consistent
with the dGOG values; this consistency is better for the
photometric cluster mass estimates than for their ages, as
shown by the larger scatter in Fig. 2a with respect to panel
b. However, we also note that the age and mass distributions
obtained for the subsample with well-determined ages and
for the full sample are internally consistent.
Fig. 3 shows the distribution of the M82 B clusters in
the age vs. mass plane. It is immediately clear that the lower
mass limit increases with increasing cluster age. The vari-
ous (solid, dashed and dotted) lines overplotted on the figure
show that this is indeed the expected effect of normal evolu-
tionary fading of a synthetic single stellar population of an
instantaneously formed cluster. For ages ≤ 109 yr, we show
the unreddened fading line for clusters with a limiting mag-
nitude of V = 22.5 at the distance of M82 (m−M = 27.8; see
dGOG), and various choices for the IMF (indicated in the
figure are the IMF slope αIMF, and the lower and upper mass
cut-offs), predicted by the Starburst99 models ("SB99"). For
older ages (t ≥ 109 yr), we show its extension predicted by
the BC00 models. These predicted lower limits agree well
with our data points.
4 THE DERIVED CLUSTER FORMATION
HISTORY
4.1 Cluster age and mass distributions
The cluster age and mass distributions for the M82 B
cluster sample are shown in Figs. 4a and b. Our new
age estimates confirm the peak in the age histogram at-
tributed to the last tidal encounter with M81 by dGOG.
For the subsample of clusters with well-determined ages
(shaded histogram), we find a peak formation epoch at
log(tpeak/yr) = 9.04 (tpeak ≃ 1.10 Gyr), with a Gaussian
σ of ∆ log(twidth) = 0.27, corresponding to a FWHM of
∆ log(twidth) = 0.64. The corresponding numbers for the
full sample are log(tpeak/yr) = 8.97 (tpeak ≃ 0.93 Gyr) for
the peak of cluster formation, and for σ, ∆ log(twidth) = 0.35
(FWHM, ∆ log(twidth) = 0.82). In dGOG (see also de Grijs
2001), we concluded that there is a strong peak of cluster
formation at ∼ 650 Myr ago, which have formed over a pe-
riod of ∼ 500 Myr, but very few clusters are younger than
log(t/yr) ≃ 8.5 (t ≃ 300 Myr). Our new age estimates date
the event triggering the starburst to be slightly older than
the value of 650 Myr derived by dGOG. Our estimate of the
duration of the burst should be considered an upper limit,
because uncertainties in the cluster age determination may
have broadened the peak in Fig. 4a.
At the distance from the centre of region B, one would
expect M82's differential rotation (cf. Shen & Lo 1995) to
have caused the starburst area to disperse on these time-
scales. The reason why the fossil starburst region has re-
mained relatively well constrained is likely found in the com-
plex structure of the disc. It is well-known that the inner
∼ 1 kpc of M82 is dominated by a stellar bar (e.g., Wills et
al. 2000) in solid-body rotation. From observations in other
galaxies, it appears to be a common feature that central bars
are often surrounded by a ring-like structure. If this is also
true for M82, it is reasonable to assume that stars in the ring
are trapped, and therefore cannot move very much in radius
because of dynamical resonance effects. The phase mixing
around the ring might be slow enough for a specific part of
the ring to keep its identity over a sufficient time so as to
appear like region B (see de Grijs 2001): if the diffusion ve-
locity around the ring is sufficiently small, any specific region
would remain self-constrained for several rotation periods.
In addition, since the density in the region is high (see Sect.
6.1), simple calculations imply that the area's self-gravity
is non-negligible compared to the rotational shear, therefore
also prohibiting a rapid dispersion of the fossil starburst re-
gion.
The improved mass estimates also confirm that the (ini-
tial) masses of the young clusters in M82 B with V ≤ 22.5
mag are mostly in the range 104 − 106M⊙, with a me-
dian of 105M⊙ (dGOG). In fact, based on a Gaussian fit,
we find that the mean mass of our M82 B cluster sample
is log(Mcl/M⊙) = 5.03 and 4.88 for the subsample with
well-determined ages and the full sample, respectively, cor-
responding to Mcl = 1.1×105 and 0.8×105M⊙, respectively.
4.2 Cluster disruption and the cluster formation
history
In Figs. 4c and d we show the formation rate and the mass
spectrum of the observed clusters in M82 B. The open cir-
cles were derived for the entire sample, whereas the filled
data points represent only those clusters for which we could
obtain good age estimates. These distributions depend on
the cluster formation history and on the cluster disruption
time-scale of M82 B.
As shown by Boutloukos & Lamers (2002, hereafter Pa-
per I), with only a few well-justified assumptions, the mass
and age distributions of a magnitude-limited sample of star
clusters in a given galaxy can be predicted both accurately
and robustly, despite the complex physical processes under-
lying the assumptions (for a full discussion see Paper I). If
all of the following conditions are met:
(i) the cluster formation rate, dN(Mcl)
constant;
dt
= S × M −α
cl
, is
(ii) the slope α of the cluster IMF is constant with
N (Mcl)dM ∝ M −α
cl dM ;
(iii) stellar evolution causes clusters to fade as Fλ ∼ t−ζ,
as predicted by cluster evolution models; and
4 × (Mcl/104M⊙)γ , where tdis
4
(iv) the cluster disruption time depends on their initial
mass as tdis = tdis
is the disrup-
tion time-scale of a cluster with Mcl = 104M⊙. It is well-
established, however, that the disruption time-scale does not
only depend on mass, but also on the initial cluster density
and internal velocity dispersion (e.g., Spitzer 1957, Cher-
noff & Weinberg 1990, de la Fuente Marcos 1997, Portegies
Zwart et al. 2001). Following the approach adopted in Pa-
per I, however, we point out that if clusters are approxi-
mately in pressure equilibrium with their environment, we
can expect the density of all clusters in a limited volume
of a galaxy to be roughly similar, so that their disruption
time-scale will predominantly depend on their (initial) mass
(with the exception of clusters on highly-eccentric orbits). In
c(cid:13) 2001 RAS, MNRAS 000, 000 -- 000
Star Cluster Formation and Disruption Time-Scales -- II.
7
Figure 2. Comparison between (a) the age and (b) mass estimates of dGOG based on BVI photometry and new estimates using
additional JH observations and a full-parameter fitting routine as described in the text. The black dots represent clusters for which the
total age range obtained is (log(Age[max]) − log(Age[min])) ≤ 1.0; the open circles are objects with more uncertain age determinations.
They show a larger scatter than the black dots, but both distributions are internally consistent.
the opposite case that the initial cluster density ρ depends
on their mass M in a power-law fashion, e.g., ρ ∼ M x with
x being the (arbitrary) power-law exponent, the disruption
time-scale will also depend on mass if tdis ∼ M aρb (Paper
I).
then it can be shown easily that the age distribution of the
observed cluster population will obey the following approx-
imate power-law behaviours (see Paper I):
• dNcl/dt ∝ tζ(1−α) for young clusters due to fading;
• dNcl/dt ∝ t(1−α)/γ for old clusters due to disruption.
Similarly the mass spectrum of the observed clusters will be
• dNcl/dMcl ∝ M (1/ζ)−α
cl
for low-mass clusters due to
For the clusters in M82 B the situation is more com-
plex, because the cluster formation rate was certainly not
constant. In fact the distribution in Fig. 4c suggests that
we can distinguish three phases, which we will discuss sepa-
rately.
We will assume that the cluster formation rate has been
constant within each phase, but may differ strongly among
the phases. We adopt a cluster IMF of slope α = 2.0 (Harris
& Pudritz 1994; McLaughlin & Pudritz 1996; Elmegreen &
Efremov 1997; Zhang & Fall 1999; Bik et al. 2002). We will
also adopt a slope ζ = 0.648 for the evolutionary fading of
clusters in the V band, derived from the Starburst99 clus-
ter models (see also Paper I). For the mass scaling of the
disruption time-scale we adopt γ = 0.62 (Paper I).
fading;
• dNcl/dMcl ∝ M γ−α
cl
for high-mass clusters due to dis-
ruption.
So both distributions will show a double power law with
slopes determined by α, ζ and γ. The crossing points are
determined by the cluster formation rate and the character-
istic disruption time-scale of a cluster with an initial mass
of 104M⊙, tdis
4 .
In Paper I we showed that the observed age and mass
distributions of the star cluster systems in four well-studied
galaxies indeed show the predicted double power-law be-
haviour. From the analysis of these observed distributions
Boutloukos & Lamers (2002) showed that the value of γ is
approximately constant, (γ = 0.62 ± 0.06), under the very
different environmental conditions in the four galaxies, but
the characteristic disruption time-scales differ significantly
from galaxy to galaxy.
c(cid:13) 2001 RAS, MNRAS 000, 000 -- 000
(i) The pre-burst phase, at log(t/yr) ≥ 9.4. The age dis-
tribution of this earliest phase is represented by only three
age-bins of high age at log(t/yr) ≥ 9.4. The predicted slope
is (1 − α)/γ = −1.61. We have fitted a straight line through
the last three age-bins with this slope. We see that the pre-
dicted slope matches the observations within the uncertain-
ties.
(ii) The burst phase, in the approximate time interval
8.4 < log(t/yr) ≤ 9.4. Its distribution can be fitted with
a line of the same predicted slope of −1.61. This is also
shown in Fig. 4c. The predicted slope fits the observations
very well. We see that this disruption line of the burst-phase
is higher than in the pre-burst phase by (0.3 ± 0.1) dex. This
implies that the cluster formation rate during the burst was
approximately a factor of 2 higher than before. However,
this factor depends on the assumed duration of the burst.
8
R. de Grijs, N. Bastian, and H.J.G.L.M. Lamers
Figure 3. Distribution of the M82 B clusters in the (age vs. mass) plane. The symbol coding is as in Fig. 2. Overplotted for ages up
to 1.0 Gyr are the expected detection limits in the (age vs. mass) plane, predicted by Starburst99 for a range of IMFs; for older ages,
we use the BC00 models for a standard IMF, as indicated in the figure legend. These model predictions are based on a detection limit
of V = 22.5 and (m − M )M82 = 27.8, assuming no extinction. For a nominal extinction of AV = 0.2 mag, expected for the clusters
with well-determined ages, the detection limit is expected to shift to higher masses by ∆ log(Mcl/M⊙) = 0.08, which is well within the
uncertainties associated with our mass determinations. The features around 10 Myr are caused by the appearance of red supergiants.
The duration of the burst may have been shorter than the
value of ∆ log(t) = 1.0 suggested by the figure, because un-
certainties in the derived cluster ages may have broadened
the burst peak. If the duration of the burst was a factor
X shorter than we adopted, then the overall cluster forma-
tion rate during the burst was a factor X higher than we
estimated.
(iii) The post-burst phase, at log(t/yr) < 8.4. The age
distribution of this most recent, post-burst phase extends
from 6.0 ≤ log(t/yr) ≤ 8.4. Unless the disruption time
was extremely short with tdis
of order a few Myr (which
is contradicted by the analysis of the mass distribution in
the burst, see Sect. 5), the age distribution of the youngest
bins is governed by the fading of the clusters and the
detection limit. Despite the rather large uncertainties for
young ages, log(t/yr) < 7.0, we find a best-fitting slope
4
ζ(1 − α) = −0.57 ± 0.05, so that ζ = 0.57 ± 0.05 (for-
mal uncertainty, not including the large error bars) for a
mass-function slope α = 2 (dotted line in Fig. 4c). This is,
within the rather large uncertainties, similar to the value of
ζ ≃ 0.65 expected from the Starburst99 and BC00 models.
Therefore we have assumed a fading line with the predicted
slope ζ(1 − α) = −0.65 for the youngest three age bins. The
age bins in the range of 7.0 ≤ log(t/yr) ≤ 8.4 fall below the
extrapolated fading line. The difference is about one dex
at log(t/yr) ≃ 8, although the number of observed clusters
is small. This suggest that the number of clusters for ages
t & 107 yr is already affected by disruption. We can fit a
disruption line with the predicted slope of −1.61 through
the data points, despite their rather large uncertainties. We
see that the best-fitting line is displaced by (−1.6 ± 0.1) dex
compared to the pre-burst disruption line. This shows that
c(cid:13) 2001 RAS, MNRAS 000, 000 -- 000
Star Cluster Formation and Disruption Time-Scales -- II.
9
Figure 4. (a) and (b) -- Age and mass distributions of the M82 B clusters, redetermined in this paper using improved methods. The
shaded histograms correspond to the clusters with (log(Age[max]) − log(Age[min]) ≤ 1.0; the open histograms represent the entire cluster
sample. (c) -- The cluster formation rate (in number of clusters per Myr) as a function of age. Open circles: full sample; filled circles:
clusters with well-determined ages, as above. The vertical dashed lines indicate the age range dominated by the burst of cluster formation.
The dotted line is the least-squares power-law fit to the fading, non-disrupted clusters, for a constant ongoing cluster formation rate.
The solid line segments are the disruption lines of clusters formed in the pre-burst phase, during the burst, and in the post-burst phase,
as described in Section 6.1. (d) -- Mass spectrum of the M82 B clusters (number of clusters per mass bin); symbol coding as in panel (c).
by supernova-driven outflows, which remove the remaining
cool gas from the immediate starburst region (e.g., Cheva-
lier & Clegg 1985, Doane & Mathews 1993). The remarkable
minor-axis wind in M82 is a dramatic example of this pro-
cess. However, the disturbed conditions near an early burst
may discourage re-ignition at the same site when cool gas
inflows resume, shifting the location of active star formation
(e.g., dGOG).
the cluster formation rate in the post-burst phase was ap-
proximately a factor of 40 smaller than during the pre-burst
phase.
We conclude from this analysis that the cluster forma-
tion rate during the burst was at least a factor of 2 higher
than during the pre-burst phase (depending on the duration
of the burst), and that the cluster formation rate was a factor
of ∼ 40 smaller after the burst than before the burst. This
is not surprising because the intense episode of cluster for-
mation during the burst will have consumed a large fraction
of the available number of molecular clouds, leaving little
material for cluster formation at later times. In fact, it is
likely that starbursts are strongly self-limited, or quenched,
c(cid:13) 2001 RAS, MNRAS 000, 000 -- 000
10
R. de Grijs, N. Bastian, and H.J.G.L.M. Lamers
5 THE MASS DISTRIBUTION OF CLUSTERS
FORMED IN THE BURST.
In this section we analyse the mass distribution of the clus-
ters that were formed during the burst of cluster forma-
tion, roughly defined to have occurred in the age range
8.4 ≤ log(t/yr) ≤ 9.4 (see the vertical dashed lines in Fig.
4c). In Fig. 5 we show their mass distribution. Although the
M82 B cluster system does not contain large numbers of
very massive clusters, its mass distributions shows a clear
increase in cluster numbers from the high-mass end towards
lower masses, for log Mcl/M⊙ & 5. For lower masses, how-
ever, the number of clusters decreases rapidly. If the mass
spectrum of young, newly formed clusters resembles a power-
law distribution, this turn-over at log Mcl/M⊙ ≃ 5 is not ex-
pected, unless disruption effects have preferentially removed
the lower-mass clusters from our magnitude-limited sample.
In dGOG we concluded that the distribution of clus-
ter luminosities (and therefore the equivalent mass distri-
bution) in M82 B -- corrected to a common age of 50 Myr
-- shows a broader and flatter mass distribution than typ-
ical for young cluster systems, although this distribution
is subject to strong selection effects. For the small age
range considered here, the selection limit in observable mass
at log(M/M⊙) ≃ 4.2 imposed by the brightness limit at
V = 22.5 mag (Fig. 3) occurs at almost an order of mag-
nitude lower masses than the turn-over mass in Fig. 5, at
log(Mcl/M⊙) ≃ 5. The arrow in Fig. 5 indicates the mass
where the onset of significant selection effects, and therefore
of sample incompleteness, is expected to occur, based on
the selection limit in the (age vs. mass) plane shown in Fig.
3. The expected effect of the AV . 0.2 mag extinction for
the clusters with well-determined ages is a marginal shift in
mass towards higher masses of ∆ log(Mcl/M⊙) . 0.08. This
implies that the observed turnover is not a spurious effect
due to varying extinction. In addition, we do not observe a
systematic trend between cluster mass and extinction, de-
rived from the SED fitting, which would be expected if the
higher-mass clusters (and therefore brighter at the same age)
were located slightly deeper into M82 B, while the fainter
lower-mass clusters were only observed near the very surface
of the region.
Therefore, we conclude that if the initial mass spec-
trum of the M82 B clusters formed in the burst of clus-
ter formation resembled a power-law distribution down to
the low-mass selection limits, cluster disruption must have
transformed this distribution on a time-scale of . 109 yrs
into a distribution resembling a log-normal or Gaussian dis-
tribution.
We now compare the observed mass distribution of the
clusters formed in the burst with model predictions. These
predictions are based on the assumption that the clusters
were formed with a cluster IMF of slope α = 2, and that
the distribution has been modified subsequently by cluster
disruption. For the cluster disruption we adopt the same
relation tdis = tdis
4 (M/104M⊙)γ with γ = 0.62 (Eq. (1)), as
found in Paper I for four galaxies, and used in the previous
section. We can then model the mass distribution of the
observed burst clusters with only one free parameter, i.e.
the disruption time tdis
of a cluster with an initial mass of
104M⊙.
4
Figure 5. Mass distribution of the clusters formed in the burst of
cluster formation, 8.4 ≤ log(t/yr) ≤ 9.4. The shaded histograms
correspond to the clusters with (log(Age[max]) − log(Age[min]) ≤
1.0; the open histograms represent the entire cluster sample in this
age range. The arrow indicates the mass where the onset of signif-
icant selection effects, and therefore of sample incompleteness, is
expected to occur, based on the selection limit in Fig. 3. The ex-
pected effect of the AV . 0.2 mag extinction for the clusters with
well-determined ages is a shift in mass towards higher masses of
∆ log(Mcl/M⊙) . 0.08, which implies that the observed turnover
is not a spurious effect due to varying extinction.
The decreasing mass of a cluster with time is given by†
dM
dt
≃ −
M
tdis
= −
M
tdis
4 (M/104M⊙)γ
,
so that
M (Mi, t) = (M γ
i − Bt)1/γ
,
(2)
(3)
where M and Mi are, respectively, the present and the initial
mass of the cluster, both in units of M⊙, and B = γ104γ /tdis
4 .
Suppose that clusters form in a burst as
dN (Mi, t) = S(t)M −α
i dMidt
(4)
over the mass range Mmin < Mi < Mmax, and governed by a
α−1 yr−1. We assume an
formation rate S(t) in units of M⊙
instantaneous burst of cluster formation, so that S(t) is a δ
function at age tburst = 109 yr. We adopt Mmax = 3×106M⊙
(i.e., log Mmax/M⊙ = 6.5, one mass bin beyond the maxi-
mum mass of clusters formed in the burst, thus allowing for
some disruption effects) and Mmin = 102M⊙, but this lower
limit is not important since it is below the detection limit of
∼ 104 M⊙ for the burst clusters (see Fig. 2). The function
† Recent simulations, based on both Monte Carlo realisations
(e.g., Giersz 2001, his Fig. 2) and N-body modeling (e.g., Porte-
gies Zwart et al. 2002, their Fig. 2; S.J. Aarseth, priv. comm.),
show that this is a close approximation to real cluster evolution,
provided that the initial cluster has reached an equilibrium state.
c(cid:13) 2001 RAS, MNRAS 000, 000 -- 000
Star Cluster Formation and Disruption Time-Scales -- II.
11
Figure 6. Predicted mass distributions in terms of log(dN/dM )×
(M/Ntot (equivalent to the logarithm of the number of clusters
per mass bin) for an instantaneous burst of cluster formation that
occurred 1 Gyr ago. From top to bottom, the curves represent the
(normalised) mass distributions for characteristic cluster disrup-
tion times log(tdis
4 /yr) = 10.0, 9.5, 9.0, 8.5, 8.0, and 7.5. Notice the
dependence of both the turnover mass, indicated by the dotted
line, and the maximum mass on tdis
4 .
S(t) is related to the total number of clusters formed per
unit time as
S(t) =
dNtot
dt
(α − 1)(cid:18)M 1−α
min − M 1−α
max(cid:19)−1
(5)
where dNtot/dt is the total number of clusters formed per
year. It is easy to show that the present mass distribution of
clusters formed in an instantaneous burst of age tburst can
be written as
N (M )dM = S(t) 1 + γ(cid:18) M
104M⊙(cid:19)γ tburst
4 !(1−α−γ)/γ
tdis
M⊙(cid:19)−α
(cid:18) M
dM
(6)
for masses above the detection limit and smaller than
(M γ
max − Btburst)1/γ. We see that the initial mass distri-
bution, dN (Mi, t) = S(t)M −α
, has been modified by a fac-
tor that depends on the ratio tburst/tdis
4 , for given values of
α = 2.0 and γ = 0.62.
i
We have calculated the predicted mass distribution, us-
ing Eq. (6), for a burst age of 1 Gyr and for several val-
ues of the cluster disruption time-scale tdis
4 , see Fig. 6. To
match the observed peak in the cluster mass distribution,
a characteristic disruption time-scale log(tdis
4 /yr) between
7.5 and 8.0 is required, with the most likely value closer to
log(tdis
4 yr) = 8.0.
We then calculated the expected number of clusters in
the same age bins as observed (Fig. 5) and normalised the
distribution to the observed total number of clusters. The
results are shown in Fig. 7a for the subsample of the 42 clus-
ters with the most accurately determined ages, and in panel
b for the full sample of 58 clusters. The location of the pre-
dicted peak is very sensitive to the value of tdis
4 /tburst, which
follows from Eq. (6), and is shown in Fig. 6. We found the
4 ≃ 3 × 10−2tburst,
best fit for a cluster disruption time of tdis
c(cid:13) 2001 RAS, MNRAS 000, 000 -- 000
Figure 7. Comparison between the observed mass distribution of
the clusters formed in the burst of cluster formation and the pre-
dicted distribution, based on a cluster IMF with slope α = 2 and
cluster disruption with time-scales of tdis = tdis
4 (M/104M⊙)0.62.
(a) Cluster sample with the most accurate masses and ages. (b)
Full sample. The predicted number of clusters in each mass bin,
normalised to the total number of observed clusters, are shown
by the dotted, full and dashed lines for characteristic disruption
time-scales log(tdis
4 /yr) = 8.5, 8.0, and 7.5, respectively.
with an uncertainty of approximately a factor of two. This
corresponds to tburst = 3 × 107 yr if tburst = 1 Gyr. We con-
clude that the mass distribution of the clusters formed dur-
ing the burst can be explained by an initial cluster IMF with
a slope of α = 2, an upper limit to their mass of ∼ 3 × 106
M⊙ and cluster disruption with a time-scale given by Eq.
(1), tdis
4 ≃ 3 × 107 yr if the age of the burst is tburst = 1 × 109
yr. The uncertainties in both the observed and predicted
numbers of clusters of a given mass are governed by identi-
cal Poisson-type statistical errors. The uncertainties in the
mass and age scales, on the other hand, are dominated by
systematic uncertainties introduced by the adopted stellar
evolutionary synthesis models, and -- to some extent -- by the
specific extinction law adopted, although the relative mass
and age distributions are robust.
12
R. de Grijs, N. Bastian, and H.J.G.L.M. Lamers
6 DISCUSSION
6.1 The cluster disruption time-scale
We have derived the characteristic cluster disruption time-
scale from the mass distribution of the clusters formed dur-
ing the burst. In principle this disruption time-scale could
also be derived from the the age distributions of the clusters
formed during the post-burst phase, shown in Fig. 4c. How-
ever the number of observed young clusters is very small,
only a few per age bin so that the crossing point between
the (dotted) fading line and the (solid) disruption line is very
uncertain. However, the crossing point suggested in Fig. 4c
agrees approximately with that expected for the disruption
time-scale of tdis
4 ≃ 3 × 107 yr derived from the clusters
formed during the burst (see the description of the method
in Paper I).
In Paper I we have shown that the disruption time-scale
can also be derived from the mass distribution of clusters in
a magnitude limited sample if the cluster formation rate
is constant. This last assumption is certainly not justified
for M82 B. For a strongly variable cluster formation rate,
as in M82 B, the mass distribution is heavily affected such
that the crossing point between the power-law fit to the
distribution of low mass clusters (the fading line) and the
power-law fit to the distribution of the massive clusters (the
disruption line) does no longer represent the value of tdis
4 .
Therefore, we could only use the mass distribution of the
clusters formed during the burst to derived the disruption
time-scale.
An initial comparison among the characteristic cluster
disruption time-scales found in Paper I reveals large differ-
ences. However, the characteristic cluster disruption time-
scale for the clusters in region B of its low-mass host galaxy
M82 is, within the uncertainties, comparable to that in the
dense centre of the massive grand-design spiral M51. Lamers
& Portegies Zwart (in prep.) are currently analysing the
cluster disruption time-scales derived for the galaxies stud-
ied in Paper I and in this paper, using N-body simulations.
Here, we will therefore simply explore whether this similar-
ity between tdis
in the centre of M51 and in M82 B can
be understood from a comparison of the ambient density in
both regions.
4
In order to estimate the density in M82 B, we deter-
mined its total V-band luminosity, after correcting for a
nominal extinction throughout the region of AV ∼ 0.5 mag,
Vtot ∼ 9.5 from our HST observations (dGOG), so that
MV,tot = −18.3, LV,tot ≃ 1.8 × 109LV,⊙. Spectral synthe-
sis of region B suggests that it has a V-band mass-to-light
ratio, M/LV ∼ 0.5 − 1.0 (R.W. O'Connell, priv. comm.),
as expected from the Starburst99 models for a stellar popu-
lation of several 100 Myr, so that the total mass contained
in the objects providing most of the V-band luminosity is
M ∼ 1.35 × 109M⊙. If we now assume that the volume oc-
cupied by these objects is roughly similar to a sphere with
a radius determined by the radius of the area in which we
measured the total flux, we derive an average density for
M82 B of hρi ∼ 2.5M⊙pc−3, or loghρi(M⊙pc−3) ∼ 0.4.
Note, however, that we have introduced large uncertainties
by adopting the above assumptions, in particular because
we have assumed to have sampled the entire volume of the
"spherical region" M82 B. This implies that the region has
been assumed transparent, as opposed to the conclusion in
dGOG that the M82 B cluster sample contains likely only
the subset of M82 B clusters on the surface of the "sphere".
The implications of our assumptions are therefore that we
expect to have missed a large number of clusters present in
the interior of M82 B. For these missed clusters in the inte-
rior we expect the disruption time to be shorter because of
the higher densities predicted there.
For the interstellar medium in the centre of M51,
Lamers & Portegies Zwart (in prep.) derive a mean density
of hρi ∼ 0.60M⊙pc−3, based on column density estimates
by Athanassoula et al. (1987) and similar geometrical argu-
ments as used above for M82 B. Within the large uncertain-
ties involved in such back-of-the-envelope approximations,
these two estimates of the mean density in the centre of
M51 and in M82 B are remarkably similar, within an order
of magnitude, as are their characteristic cluster disruption
time-scales. Thus, we conclude that, although M82 as such is
a small, low-mass irregular galaxy, its fossil starburst region
B has achieved a similarly dense interstellar medium as the
centre of M51, so that similar cluster disruption time-scales
are not a priori ruled out.
6.2 The cluster formation rate
We have derived the cluster formation history of M82 B from
the age distribution shown in Fig. 4c. The age distribution
itself, Fig. 4a already shows a very strong peak around 109
years. Part of the steep increase between 8 < log(t/yr) < 9 is
owing to the use of logarithmic age bins. Therefore we have
transformed the age distribution into the formation history
of the observed clusters, Fig 4c. The general decrease with
age clearly shows the effect of cluster disruption. Disruption
of clusters formed at a constant formation rate and with a
mass-dependent disruption time-scale will result in a power-
law decrease of the observed cluster formation rate (Paper I),
roughly in agreement with the observations. The deviations
from the power law, e.g., the peak in the observed formation
history around 1 Gyr, reflect changes in the real formation
history, or in the disruption time.
For magnitude-limited cluster samples with a constant
cluster formation rate and a constant mass-dependent dis-
ruption time-scale, the disruption time and its dependence
on mass can be derived from the age distribution. Because
of the non-constant cluster formation rate and the small
number of clusters, such an analysis is not possible here.
Therefore we opted for the reasonable alternative of adopt-
ing the mass dependence of the disruption time-scale, i.e.,
the value of γ ≃ 0.62 in Eq. (1), which seems to be a uni-
versal value found in four galaxies with very different con-
ditions (Paper I). We also assumed that the scaling factor
tdis
4 of the disruption time-scale is constant. With these two
assumptions, we derived the ratios between the real clus-
ter formation rates in the pre-burst phase (log(t/yr) ≥ 9.4),
the burst-phase (8.4 < log(t/yr) < 9.4) and the post-burst
phase (log(t/yr) ≤ 8.4 of, roughly, 1 : 2 : 1
40 . The formation
rate during the burst may have been higher if the actual
duration of the burst was shorter than adopted.
Alternatively, we could have adopted a constant cluster
formation rate, but wildly variable cluster disruption time-
scales. In that case the cluster disruption time-scale should
have been much longer (i.e. much slower disruption) dur-
ing the burst than before the burst, and again much shorter
c(cid:13) 2001 RAS, MNRAS 000, 000 -- 000
Star Cluster Formation and Disruption Time-Scales -- II.
13
(faster disruption) after the burst. We think that this alter-
native is very unlikely because (a) disruption is a slow pro-
cess that occurs over an extended period of time, so large
changes are not to be expected, (b) there is no obvious phys-
ical process that would produce such large changes and (c)
the alternative of a variable cluster formation rate is much
more likely, because this effect is observed in several inter-
acting galaxies and in M82 regions A, C and E (dGOG).
Therefore, we are confident that the changes in the apparent
formation rates of the observed clusters are due to changes
in the real cluster formation rate.
7 SUMMARY AND CONCLUSIONS
In this paper, we have reanalysed the previously published
optical and near-infrared HST photometry of the star clus-
ters in M82's fossil starburst region B (dGOG) to obtain
improved individual age and mass estimates. We have also
extended this study to obtain estimates for the cluster for-
mation history and include the importance of cluster disrup-
tion. Our main results and conclusions can be summarised
as follows:
(i) Our new age estimates, based on improved fitting
methods, confirm the peak in the age histogram attributed
to the last tidal encounter with M81 by dGOG. For the sub-
sample of clusters with well-determined ages we find a peak
formation epoch at log(tpeak/yr) = 9.04 (tpeak ≃ 1.10 Gyr),
with a Gaussian σ of ∆ log(twidth) = 0.27, corresponding to
a FWHM of ∆ log(twidth) = 0.64. The corresponding num-
bers for the full sample are log(tpeak/yr) = 8.97 (tpeak ≃
0.93 Gyr) for the peak of cluster formation, and for σ,
∆ log(twidth) = 0.35 (FWHM, ∆ log(twidth) = 0.82). In
dGOG (see also de Grijs 2001), we concluded that there is
a strong peak of cluster formation at ∼ 650 Myr ago, which
have formed over a period of ∼ 500 Myr, but very few clus-
ters are younger than 300 Myr (log(t/yr) ≃ 8.5). Our new
age estimates date the event triggering the starburst to be
slightly older. Our estimate of the duration of the peak is an
upper limit because the observed width of the peak in the
age histogram may have been broadened by uncertainties in
the derived cluster ages.
(ii) The improved mass estimates confirm that the (ini-
tial) masses of the young clusters in M82 B with V ≤ 22.5
mag are mostly in the range 104 − 106M⊙, with a me-
dian of 105M⊙ (dGOG). We find that the mean mass of
our M82 B cluster sample is log(Mcl/M⊙) = 5.03 and 4.88
for the subsample with well-determined ages and the full
sample, respectively, corresponding to Mcl = 1.1 × 105 and
0.8×105M⊙, respectively. If the initial mass spectrum of the
clusters formed at the burst epoch was a power law, disrup-
tion effects have transformed it into a broader and flatter
distribution on time-scales of . 1 Gyr.
(iii) The (apparent) formation history of the observed
clusters shows a gradual increase to younger ages, reach-
ing a maximum at the present time. This shows that cluster
disruption must have removed a large fraction of the older
clusters. Adopting the expression for the cluster disruption
time, tdis(M ) = tdis
4 (M/104M⊙)γ with γ ≃ 0.62, that was
derived for four galaxies characterized by very different con-
ditions (Paper I), we found that ratios between the real clus-
ter formation rates in the pre-burst phase (log(t/yr) ≥ 9.4),
c(cid:13) 2001 RAS, MNRAS 000, 000 -- 000
the burst-phase (8.4 < log(t/yr) < 9.4) and the post-burst
phase (log(t/yr) ≤ 8.4) are roughly 1 : 2 : 1
40 . The formation
rate during the burst may have been higher if the actual du-
ration of the burst was shorter than adopted. We see that
the cluster formation rate in the post-burst phase is much
smaller than in the pre-burst phase, because the burst has
consumed a large fraction of the available molecular clouds,
leaving little material for the cluster formation in the post-
burst phase.
(iv) The mass distribution of the clusters formed during
the burst shows a turnover at log(Mcl/M⊙) ≃ 5.2. This
turnover is not due to selection effects, because the mag-
nitude limit of our sample would produce a turnover near
log(Mcl/M⊙) ≃ 4.5. The observed distribution can be ex-
plained by cluster formation with an initial power-law mass
function of exponent α = 2 up to a maximum cluster mass of
Mmax = 3 × 106M⊙, and cluster disruption given by Eq. (1)
with the adopted value of γ = 0.62 with only one free fitting
4 /tburst ≃ 3 × 10−2. For a burst age of 1 × 109
parameter: tdis
yr, we find that tdis
4 ≃ 3 × 107 years, with an uncertainty of
a factor of two.
(v) The time-scale of tdis
4 ∼ 30 Myr is much shorter than
derived in Paper I for any of the SMC (∼ 8 Gyr), the so-
lar neighbourhood (∼ 1 Gyr), M33 (∼ 0.13 Gyr) and even
shorter than in the inner spiral arms of M51 (∼ 40 Myr). The
characteristic disruption time-scale in M82 B is the shortest
known in any disc (region of a) galaxy.
ACKNOWLEDGMENTS
RdeG wishes to thank the Astronomical Institute of Utrecht
University for their hospitality and partial support on a visit
when this work was started. We acknowledge interesting and
useful discussions with Bob O'Connell and Sverre Aarseth
and helpful suggestions by the anonymous referee. This re-
search has made use of NASA's Astrophysics Data System
Abstract Service.
REFERENCES
Athanassoula E., Bosma A., Papaioannou 1987, A&A, 179, 23
Barger A.J., Arag´on-Salamanca A., Ellis R.S., Couch W.J., Smail
I., Sharples R.M., 1996, MNRAS, 279, 1
Bik A., Lamers H.J.G.L.M., Bastian N., Panagia N., Romaniello
M., Kirshner R., 2002, A&A, in press (astro-ph/0210594)
Boutloukos S.G., Lamers H.J.G.L.M., 2001, in: Extragalactic Star
Clusters, eds. Grebel E.K., Geisler D., IAU Symp. 207, (San
Francisco: ASP), in press
Boutloukos S.G., Lamers H.J.G.L.M., 2002, MNRAS, in press
(astro-ph/0210595; Paper I)
Brandl B., Chernoff D.F., Moffat A.F.J., 2001, in: Extragalactic
Star Clusters, eds. Grebel E.K., Geisler D., IAU Symp. 207,
(San Francisco: ASP), in press
Bruzual G., Charlot S., 2000, updated version of Bruzual G.,
Charlot S., 1996, in: Leitherer, C., et al. 1996, PASP, 108,
996 (AAS CDROM Series 7) (BC00)
Butcher H., Oemler A., 1978, ApJ, 219, 18
Chernoff D.F., Weinberg M.D., 1990, ApJ, 351, 121
Chevalier R.A., Clegg A.W., 1985, Nat., 317, 44
Couch W.J., Barger A.J., Smail I., Ellis R.S., Sharples R.M., 1998,
ApJ, 497, 188
Couch W.J., Sharples R.M., 1987, MNRAS, 229, 423
14
R. de Grijs, N. Bastian, and H.J.G.L.M. Lamers
de Grijs R., 2001, A&G, 42, 14
de Grijs R., Johnson R.A., Gilmore G.F., Frayn C.M., 2002a,
MNRAS, 331, 228
de Grijs R., Gilmore G.F., Johnson R.A., Mackey A.D., 2002b,
MNRAS, 331, 245
de Grijs R., O'Connell R.W., Gallagher J.S., 2001, AJ, 121, 768
(dGOG)
de la Fuente Marcos R., 1997, A&A, 322, 764
Doane J.S., Mathews W.G., 1993, ApJ, 419, 573
Dressler A., Gunn J.E., 1990, in: Evolution of the Universe of
Galaxies, ed. Kron R.G., (San Francisco: ASP), p. 200
Elmegreen B.G., Efremov Y.N., 1997, ApJ, 480, 235
Elson R.A.W., 1991, ApJS, 76, 185
Elson R.A.W., 1992, MNRAS, 256, 515
Elson R.A.W., Freeman K.C., Lauer T.R., 1989, ApJ, 347, 69
Fritze-v. Alvensleben U., Gerhard O.E., 1994, A&A, 285, 775
Gallagher J.S., Smith L.J., 1999, MNRAS, 304, 540
Giavalisco M., 1997, in: The Hubble Deep Field, eds. Livio M.,
Fall S.M., Madau P., (Cambridge: CUP), p. 121
Giersz M., 2001, MNRAS, 324, 218
Girardi L., Chiosi C., Bertelli G., Bressan A., 1995, A&A, 298,
87
Harris W.E., Pudritz R.E., 1994, ApJ, 429, 177
Ho L.C., Filippenko A.V., 1996a, ApJ, 466, L83
Ho L.C., Filippenko A.V., 1996b, ApJ, 472, 600
Kroupa P., Tout C.A., Gilmore G.F., 1993, MNRAS, 262, 545
Leitherer C., Schaerer D., Goldader J.D., Gonzalez Delgado R.M.,
Robert C., Kune D.F., de Mello D.F., Devost D., Heckman
T.M., 1999, ApJS, 123, 3 (Starburst99)
Lowenthal J.D., et al., 1997, ApJ, 481, 673
Lynds C.R., Sandage A.R., 1963, ApJ, 137, 1005
Marcum P., O'Connell R.W., 1996, in: From Stars to Galaxies:
The Impact of Stellar Physics on Galaxy Evolution, eds. Lei-
therer C., Fritze-von Alvensleben U., Huchra J., (San Fran-
cisco: ASP), p. 419
McCarthy P.J., Heckman T., van Breugel W., 1987, AJ, 92, 264
McLaughlin D.E., Pudritz R.E., 1996, ApJ, 457, 578
Mengel S., Lehnert M.D., Thatte N., Genzel R., 2002, A&A, 383,
137
O'Connell R.W., Gallagher J.S., Hunter D.A., Colley W.N., 1995,
ApJ, 446, L1
O'Connell R.W., Mangano J.J., 1978, ApJ, 221, 62
Oemler A., 1992, in: Clusters & Superclusters of Galaxies, ed.
Fabian A.C. (Dordrecht: Kluwer), p. 29
Oort J.H., 1957, "Stellar Populations", Rome: Pontifical Academy
of Science, discussion in session on star clusters
Parmentier G., de Grijs R., Gilmore G.F., 2002, MNRAS, sub-
mitted
Portegies Zwart S.F., Makino J., McMillan S.L.W., Hut P., 2001,
ApJ, 546, L101
Portegies Zwart S.F., Makino J., McMillan S.L.W., Hut P., 2002,
ApJ, 565, 265
Rieke G.H., Loken K., Rieke M.J., Tamblyn P., 1993, ApJ, 412,
99
Scuderi S., Panagia N., Gilmozzi R., Challis P.M., Kirshner R.P.,
1996, ApJ, 465, 956
Shen J., Lo K.Y., 1995, ApJ, 445, L99
Shopbell P.L., Bland-Hawthorn J., 1998, ApJ, 493, 129
Smith L.J., Gallagher J.S., 2001, MNRAS, 326, 1027
Spitzer L. Jr., 1957, ApJ, 127, 17
Steidel C.C., Giavalisco M., Pettini M., Dickinson M., Adelberger
K.L., 1996, ApJ, 462, L17
Telesco C.M., 1988, ARA&A, 26, 343
Vesperini E., 2000, MNRAS, 318, 841
Vesperini E., 2001, MNRAS, 322, 247
Wills K.A., Das M., Pedlar A., Muxlow T.B.W., Robinson T.G.,
2000, MNRAS, 316, 33
Zhang Q., Fall S.M., 1999, ApJ, 527, L81
c(cid:13) 2001 RAS, MNRAS 000, 000 -- 000
|
astro-ph/9708026 | 1 | 9708 | 1997-08-04T13:23:47 | Violent Relaxation, Phase Mixing, and Gravitational Landau Damping | [
"astro-ph"
] | This paper proposes a geometric interpretation of flows generated by the collisionless Boltzmann equation (CBE), focusing on the coarse-grained approach towards equilibrium. The CBE is a noncanonical Hamiltonian system with the distribution function f the fundamental dynamical variable, the mean field energy H[f] playing the role of the Hamiltonian and the natural arena of physics being the infinite-dimensional phase space of distribution functions. Every time-independent equilibrium f_0 is an energy extremal with respect to all perturbations that preserve the constraints associated with Liouville's Theorem, local energy minima corresponding to linearly stable equilibria. If an initial f(t=0) is sufficiently close to some linearly stable lower energy f_0, its evolution involves linear phase space oscillations about f_0 which, in many cases, would be expected to exhibit linear Landau damping. If f(t=0) is far from any stable extremal, the flow will be more complicated but, in general, one would anticipate that the evolution involves nonlinear oscillations about some lower energy f_0. In this picture, the coarse-grained approach towards equilibrium usually termed violent relaxation is interpreted as nonlinear Landau damping. The evolution of a generic initial f(t=0) involves a coherent initial excitation, not necessarily small, being converted into incoherent motion associated with nonlinear oscillations about some equilibrium f_0 which, in general, will exhibit destructive interference. | astro-ph | astro-ph |
Violent Relaxation, Phase Mixing, and
Gravitational Landau Damping
Henry E. Kandrup
Department of Astronomy and Department of Physics and
Institute for Fundamental Theory, University of Florida
Gainesville, FL 32611 USA
Abstract
This paper outlines a geometric interpretation of flows generated by the colli-
sionless Boltzmann equation, focusing in particular on the coarse-grained ap-
proach towards a time-independent equilibrium. The starting point is the recog-
nition that the collisionless Boltzmann equation is a noncanonical Hamiltonian
system with the distribution function f as the fundamental dynamical variable,
the mean field energy H[f ] playing the role of the Hamiltonian and the natural
arena of physics being Γ, the infinite-dimensional phase space of distribution
functions. Every time-independent equilibrium f0 is an energy extremal with
respect to all perturbations δf that preserve the constraints (Casimirs) associ-
ated with Liouville's Theorem. If the extremal is a local energy minimum, f0
must be linearly stable but, if it corresponds instead to a saddle point, f0 may
be unstable. If an initial f (t = 0) is sufficiently close to some linearly stable
lower energy f0, its evolution can be visualised as involving linear phase space
oscillations about f0 which, in many cases, would be expected to exhibit linear
Landau damping. If instead f (0) is far from any stable extremal, the flow will be
more complicated but, in general, one might anticipate that the evolution can
be visualised as involving nonlinear oscillations about some lower energy f0. In
this picture, the coarse-grained approach towards equilibrium usually termed
violent relaxation is interpreted as nonlinear Landau damping. Evolution of
a generic initial f (0) involves a coherent initial excitation δf (0) ≡ f (0) − f0,
not necessarily small, being converted into incoherent motion associated with
nonlinear oscillations about some f0 which, in general, will exhibit destructive
interference. This picture allows for distinctions between regular and chaotic
"orbits" in Γ: Stable extremals f0 all have vanishing Lyapunov exponents, even
though "orbits" oscillating about f0 may well correspond to chaotic trajectories
with one or more positive Lyapunov exponents.
1
1. Introduction and Motivation
The problem addressed in this paper is how to visualise flows generated by the collision-
less Boltzmann equation (CBE), i.e., the gravitational analogue of the electrostatic Vlasov
equation from plasma physics.
It is generally accepted that many physical problems arising in galactic dynamics and
cosmology can be modeled in terms of the CBE, perhaps allowing also for low amplitude
discreteness effects, modeled as friction and noise through the formulation of a Fokker-
Planck equation, or for a coupling to a dissipative fluid described, e.g., by the Navier-
Stokes equation. Astronomers recognise that an evolution described completely by the
CBE is special because of the constraints associated with Liouville's Theorem, and that, at
some level, the flow must be Hamiltonian, which precludes the possibility of any pointwise
approach towards a time-independent equilibrium:
in the absence of dissipation, one can
only speak meaningfully of a coarse-grained approach towards equilibrium. However, there
does not seem to be a clear sense of exactly how one ought to visualise a flow governed by
the CBE or of what sort of coarse-graining one ought to implement in order to identify an
approach towards equilibrium.
The conventional wisdom of galactic dynamics (cf. Binney and Tremaine 1987), as ar-
ticulated, e.g., by Maoz (1991), draws sharp distinctions between different aspects of the
evolution, speaking separately of phase mixing, (linear) Landau damping, and violent re-
laxation. However, such distinctions, even if useful in addressing specific physical effects,
are arguably ad hoc and, as such, may obscure the overall character of the flow. Plasma
physicists (cf. van Kampen 1955, Case 1959) are well acquainted with the fact that, appro-
priately interpreted, linear Landau damping is a phase mixing associated with the evolution
of a wave packet constructed from a continuous set of normal modes. Moreover, even though
conventional wisdom makes a sharp distinction between violent relaxation and phase mix-
ing/Landau damping, one can argue that, as is implicit in Lynden-Bell's (1967) original
paper on violent relaxation, it too is a phase mixing process.
The objective here is to present a coherent mathematical description of an evolution
described by the CBE that manifests explicitly the Hamiltonian character of the flow.
This entails a synthesis and extension of existing work in both plasma physics and galactic
dynamics (cf. Morrison 1980, Morrison and Eliezur 1986, Kandrup 1989, 1998 and numerous
references cited therein) which, in the context of galactic dynamics, has proven useful in
understanding problems related to both linear and global stability, as well as stability
in the presence of weak dissipation (cf. Kandrup 1991a,b, Perez and Aly 1996, Perez,
Alimi, Aly, and Scholl 1996). Section 2 describes the precise sense in which the CBE is
an infinite-dimensional Hamiltonian system, identifying the natural phase space, exhibiting
the noncanonical Hamiltonian structure, and then speculating on the possible meaning of
regular versus chaotic flows.
Section 3 turns to the problem of linear stability for time-independent equilibria. This is
2
addressed both in the context of the full noncanonical Hamiltonian dynamics and in terms
of a simpler canonical Hamiltonian structure associated with the tangent dynamics, i.e.,
identifying explicitly a set of canonically conjugate variables in terms of which to analyse
linear perturbations. One immediate by-product of this discussion is a simple explanation
(cf. Habib, Kandrup, and Yip 1986) of linear Landau damping which manifests explicitly
that it is in fact a phase mixing process: Even though a perturbation cannot "die away"
in any pointwise sense, one may expect a coarse-grained approach towards equilibrium in
which observables like the density perturbation δρ eventually decay to zero.
Section 4 generalises the preceding to the case of nonlinear stability, allowing for per-
turbations δf away from some equilibrium f0 which are not necessarily small. The intuition
derived from that problem is then used to motivate one possible way in which to visualise
the flow associated with a generic initial f (t = 0). The obvious point is that a generic initial
f (0) can be viewed as a (possibly strongly nonlinear) perturbation of some equilibrium f0,
the form of which, however, need not be known explicitly. To the extent that this inter-
pretation is accepted, those aspects of the flow typically denoted violent relaxation should
be viewed as nonlinear Landau damping/phase mixing (cf. Kandrup 1998). Section 5 con-
cludes by describing the mathematical issues which must be resolved to make the preceding
discussion rigorous and complete.
A simple mechanical model, which can help in visualising the basic ideas described
in this paper, is the following: Consider a point particle moving in some complicated,
many-dimensional potential V (r) which is characterised generically by multiple extremal
points but which, being bounded from below, will have a (in general nondegenerate) global
minimum. If one chooses initial data corresponding to a configuration space point r close to
but slightly above some local minimum r0 and a velocity v whose magnitude is very small,
the subsequent evolution will involve linear oscillations about r0, whether or not that point
corresponds to a global minimum. The trajectory of the point particle thus corresponds
to a regular orbit in what appears locally as a harmonic potential. If the initial deviation
from the extremal point becomes somewhat larger, because r − r0 and/or v is bigger,
one would still anticipate oscillations around r0, but these will now become nonlinear and
the particle trajectory may well correspond to a chaotic orbit. Suppose, however, that r0
is not the global minimum. In this case, one would expect that, for initial data sufficiently
far from r0, the particle will have left the "basin of attraction" associated with the local
minimum and will instead (generically) exhibit strongly nonlinear oscillations about the
global minimum (it could of course oscillate around a different nonglobal minimum!). In the
absence of dissipation, there is no pointwise sense in which the particle evolves towards the
global minimum. However, the nonlinear oscillations in different directions will in general
interfere destructively, so that any initial coherence between motions in different directions
will eventually be lost (at least for times short compared with the Poincar´e recurrence time).
It is this loss of coherence which, for the CBE, gives rise to (linear or nonlinear) Landau
damping.
3
2. The Noncanonical Hamiltonian Formulation
If one considers the Liouville equation appropriate for a collection of noninteracting
particles evolving in a fixed potential Φ(x), the natural phase space is the six-dimensional
phase space associated with the canonical pair (x, v). If, however, one considers the full
CBE, allowing for a self-consistent potential Φ[f (x, v)] determined by the free-streaming
particles, this is no longer so.
In this case, the fundamental dynamical variable is the
distribution function itself, and the natural phase space Γ is the infinite-dimensional phase
space of distribution functions. In general, it is not easy to identify conjugate coordinates
and momenta in this phase space so as to rewrite the CBE in the form of Hamilton's
equations. However, one can still capture the Hamiltonian character at a formal algebraic
level through the identification of an appropriate cosymplectic structure (cf. Arnold 1989).1
In this context, manifesting the Hamiltonian character of the flow entails identifying a
Lie bracket [ . , . ], defined on pairs of phase space functionals A[f ] and B[f ], and a Hamil-
tonian functional H[f ], so chosen that the CBE
with Φ(x, t) the self-consistent potential satisfying
∂f
∂t
+ v·
∂f
∂x
− ∇Φ·
∂f
∂v
= 0,
can be written in the form
∇2Φ = 4πGρ ≡ Z d3v f,
∂f
∂t
+ [H, f ] = 0.
(1)
(2)
(3)
1 One example of a noncanonical Hamiltonian system, well known to astronomers, is rigid body
rotations described by the standard Euler equations (cf. Landau and Lifshitz 1960). Specifically,
as described and generalised, e.g., in Kandrup (1990) and Kandrup and Morrison (1993), the Euler
equations constitute a Hamiltonian system, formulated in the three-dimensional phase space coor-
dinatised by the three components of angular momentum Ji, (i = 1, 2, 3), with the Hamiltonian
i /2Ii (the analogue of eq. 4) defined in terms of the principal moments of inertia
Ii, and the Lie bracket (the analogue of eq. 5) given as the natural bracket associated with the
three-dimensional rotation group, i.e.,
H[Ji] = P3
i=1 J 2
[a, b] = Xi,j,k
ǫijkJk(cid:0) ∂a
∂Ji(cid:1)(cid:0) ∂b
∂Jj(cid:1)
for functions a(Ji) and b(Ji). As for the CBE, there is also a Casimir (the analogue of eq. 9),
i , which restricts motion to the two-dimensional constant C surface in the
i=1 J 2
namely C[Ji] = P3
three-dimensional phase space.
Astronomers are also acquainted with infinite-dimensional Hamiltonian systems, at least those
t Ψ − ∇2Ψ =
realisable in canonical coordinates, one simple example being the scalar wave equation ∂ 2
0, which derives from the Hamiltonian
H =
1
2 Z d3x(cid:16)Π2(x) + ∇Ψ(x2(cid:17),
where Ψ and Π are canonically conjugate.
4
The Hamiltonian H may be taken as
H[f ] =
1
2 Z dΓ v2 f (x, v) −
G
2 Z dΓ Z dΓ′ f (x, v)f (x′, v′)
x − x′
,
(4)
with dΓ ≡ d3xd3v, which corresponds to the obvious mean field energy, as identified, e.g.,
by Lynden-Bell and Sanitt (1969). The bracket is then chosen to satisfy (Morrison 1980)
[A, B] = Z dΓ f n δA
δf
,
δB
δfo,
(5)
where {g, h} denotes the ordinary Poisson bracket acting on functions g(x, v) and h(x, v),
and δ/δf denotes a functional derivative. It is straightforward to show that the operation
defined by eq. (5) is a skew symmetric, bilinear form, satisfying the Jacobi identity
[g, [h, k]] + [h, [k, g]] + [k, [g, h]] = 0,
(6)
so that it defines a bona fide Lie bracket. However for this bracket one verifies immediately
that eq. (3) reduces to the CBE in the form
∂f
∂t
− {E, f } = 0,
where E represents the energy of a unit mass test particle, i.e.,
E =
1
2
v2 + Φ(x, t).
(7)
(8)
A flow governed by the CBE is strongly constrained by Liouville's Theorem, which
implies the existence of an infinite number of conserved quantities, the so-called Casimirs
C[f ]. Specificially, the flow has the property that, for any function χ(f ), the value of the
phase space integral
C[f ] =Z dΓ χ(f )
(9)
is invariant under time translation, i.e., dC/dt = 0. The simplest case corresponds to the
choice χ = f , which leads to conservation of number (or mass):
d
dt Z dΓ f ≡ 0.
(10)
By analogy with finite-dimensional systems, where Noether's Theorem relates conserved
quantities to continuous symmetries, these Casimirs reflect internal symmetries in the
infinite-dimensional phase space Γ (Morrison and Eliezur 1986).
The Casimirs play an important role in analysing the stability of equilibrium solutions
f0, where one must restrict attention to perturbations δf that satisfy δC ≡ 0 for all possible
choices of χ. As first noted by Bartholomew (1971), this demand implies that any allowed
perturbation δf is related to f0 by a canonical transformation induced by some generating
function g, i.e.,
f ≡ f0 + δf = exp({g, . })f0.
(11)
5
In addition to the Casimirs, there is also at least one other conserved quantity, namely
the mean field energy H[f ]. Specifically, it follows from the CBE that dH/dt ≡ 0. If one
considers initial data f (0) characterised by a high degree of symmetry, other conserved
quantities may also exist. For example, if the initial data correspond to a potential Φ which
is spherically symmetric, it follows that the numerical value of the angular momentum
J ≡ Z d3xd3v f x×v
(12)
is necessarily conserved. However, these conserved quantities, if they exist, are on a different
footing from the Casimirs since they reflect symmetries in the particle phase space, rather
than internal symmetries associated with the infinite-dimensional phase space of distribution
functions.
Because of the infinite number of constraints associated with the Casimirs, the evolution
of f is reduced to a lower (but presumably still infinite-) dimensional phase space hypersur-
face, say γ. One might naively believe that, in the same sense as, e.g., for the Kortweg-de
Vries equation (cf. Arnold 1989), the flow associated with the CBE is integrable. In point
of fact, however, this is almost certainly not so (cf. Morrison 1987), the important point
being that the Casimirs associated with the CBE are all "ultralocal" quantities which do
not involve derivatives of f .
At the present time, there is no universally accepted notion of what precisely one should
mean by chaos in an infinite-dimensional Hamiltonian system. However, one obvious tact
entails comparing initially nearby flows and asking whether, for some given f (t = 0), there
exist perturbations δf (t = 0) which grow exponentially. This leads naturally2 to the notion
of a functional Lyapunov exponent which, at least formally, can be defined by analogy
with the definition of an ordinary Lyapunov exponent in a finite-dimensional system (cf.
Lichtenberg and Lieberman 1992). Specifically, given the introduction of an appropriate
norm , one can write
χ = lim
t→∞
lim
δf (0)→0
1
t
δf (t)
δf (0)
.
(13)
For finite dimensional systems one knows that, independent of the choice of norm, the
analogue of eq. (13) will, for a generic phase space perturbation δz, converge towards the
largest Lyapunov exponent. Much less is known about the infinite-dimensional case. For
specificity, it thus seems reasonable to choose as corresponding to a (possibly weighted)
L2 norm defined in the phase space of distribution functions, i.e.,
δf ≡ Z dΓ M (x, v) δf (x, v)2,
(14)
where M denotes a specified function of x and v. This is, e.g., the type of norm that has
been used in proving theorems about linear stability.
2 I thank Bruce Miller and Klaus Dietz for suggesting this point to me.
6
3. Linear Stability and Gravitational Landau Damping
The key fact underlying the interpretation of flows described by the CBE and, especially,
the problem of stability, is that every time-independent equilibrium f0 is an energy extremal
with respect to "symplectic" perturbations δf of the form (11) which preserve the numerical
values of every Casimir. This implies that, if one restricts attention to the reduced phase
space γ obtained by freezing the value of each Casimir at its equilibrium value C[f0], every
equilibrium f0 corresponds to an isolated fixed point: To lowest order, the quantity δH ≡ 0
for any symplectic δf . As explained below, if f0 is a local energy minimum, so that, to
next leading order, δH ≥ 0, f0 must be linearly stable. Alternatively, if f0 corresponds to
a saddle point, so that H increases for some perturbations but decreases for others, linear
stability is no longer guaranteed, although one cannot necessarily infer that f0 must be
linearly unstable.
The proof that, to lowest order, δH vanishes for any perturbation of the form (11) and
the computation of δH to higher order are straightforward if one expands (11) perturbatively
to infer that
δf = {g, f0} +
{g, {g, f0}} + . . . ≡ δ(1)f + δ(2)f + . . . .
1
2
(15)
(16)
It is easy to see that, for any δ(1)f , the first variation δ(1)H becomes
δ(1)H = Z dΓ (cid:16) 1
2
v2 − GZ dΓ′
f ′
0
x − x′(cid:17)δ(1)f = Z dΓE0δ(1)f,
where E0 is the particle energy associated with f0. However, by combining eqs. (15) and
(16) and then integrating by parts, one finds that
δ(1)H = Z E0 {g, f0} = −Z dΓ g{E0, f0} ≡ 0,
(17)
where (cf. eq. 7) the final equality follows from fact that f0 is time-independent. Extending
this calculation to one higher order shows that the second variation
δ(2)H = −
1
2 Z dΓ {g, f0} {g, E0} −
G
2 Z dΓZ dΓ′ {g, f0} {g′, f ′
x − x′
0}
.
(18)
To help visualise what is going on, and to understand why linear stability follows if
δ(2)H is positive for all symplectic perturbations of the form (11), suppose that, in ordinary
three-dimensional space, the x-y plane corresponds to a hypersurface in the reduced γ-space
of distribution functions. One can then "warp" this plane into a curved two-dimensional
surface by assigning to each x-y pair a coordinate z which corresponds to the numerical
value assumed by the energy H. On this warped surface, the equilibrium points correspond
to those pairs (x0, y0) which are extremal in z, so that any infinitesimally displaced point
(x0 + δx, y0 + δy) assumes a new value z + δz.
If the equilibrium point is a local energy minimum, any infinitesimal displacement on
the surface necessarily increases the value of z, so that, in the neighbourhood of (x0, y0),
the surface has the geometry of an upward opening paraboloid. Any perturbation comes
7
with positive energy and corresponds to bounded motion on the paraboloid. Thus the
equilibrium is linearly stable. In principle, the same conclusion also obtains if the extremal
point is a local maximum, although one can show that, for realistic equilibria, δ(2)H is
never strictly negative.
If, however, the equilibrium corresponds to a saddle point, so
that z increases in some directions but decreases in others, the situation becomes more
complicated. In this case, the linearised dynamics implies that it is possible to combine a
very large negative energy perturbation in one direction with a very large positive energy
perturbation in another to generate a total perturbation with vanishing energy. In itself,
this does not guarantee a linear instability, but the simple geometric argument for stability
that holds for a local minimum is no longer applicable.3
That saddle points need not imply linear instability may seem surprising at first glance.
However, the following two-dimensional example makes clear exactly what can go wrong:
H =
1
2(cid:16)v2
1 + ω2
1(cid:17) −
1x2
1
2(cid:16)v2
2 + ω2
2(cid:17).
2x2
(19)
Here x1 = v1 = x2 = v2 = 0 is a time-independent extremal point in the phase space which
corresponds to a saddle but, nevertheless, the equilibrium is clearly stable. This model may
seem somewhat contrived but, as discussed in Section V of Kandrup and Morrison (1993),
such stable saddle points are not uncommon in various infinite-dimensional Hamiltonian
systems.
The preceding argument for stability or lack thereof may seem somewhat unusual be-
cause it is formulated abstractly in phase space, without the introduction of conjugate
coordinates and momenta. One might therefore hope that, by identifying an appropriate
set of conjugate variables, a more intuitive proof could be derived. In certain cases, this is
in fact possible. One knows that, when formulated in the full Γ-space, the dynamics cannot
be decomposed completely into canonical variables because of the existence of the Casimirs,
which correspond to null vectors of the cosymplectic structure. If, however, one passes to
the reduced γ space, where the values of all the Casimirs are frozen, one might expect
that, at least locally, conjugate variables do exist. Indeed, for finite-dimensional systems
it follows from Darboux's Theorem (cf. Arnold 1989) that, if the cosymplectic structure
has vanishing determinant, i.e., if there are no null eigenvectors, it is always possible to
find a set of canonically conjugate variables, at least locally (see Section V of Kandrup and
Morrison [1993] for a detailed discussion of this point).
One setting in which such a canonical formulation is possible is for the special case of
linear perturbations of an equilibrium f0 which is a function only of the one-particle energy
3 Strictly speaking, the application of this finite-dimensional argument to an infinite-dimensional
Hamiltonian system requires that the reduced phase space γ be endowed with a metric, so that one
knows what is meant by distance between points. In practice, this can be done by introducing an
appropriate L2 norm, which provides the natural extension of the Euclidean notion of distance to an
infinite-dimensional space. In this context, a proof of stability entails showing that δf (t) remains
bounded for all times.
8
E, i.e., f0 = f0(E), and for which the partial derivative FE ≡ ∂f /∂E is strictly negative.
Physically the latter restriction implies that the system does not exhibit a population in-
version; mathematically it ensures that division by FE is well defined. The basic idea, due
originally to Antonov (1960), is to split the linearised perturbation δf into two pieces, δf+
and δf−, respectively even and odd under a velocity inversion v → −v, and to view the
single linearised perturbation equation for δf as a coupled system for δf±.
When linearised about some equilibrium f0, the CBE reduces to
∂tδf − {E, δf } − {Φ[δf ], f0} = 0,
(20)
where E is the particle energy associated with f0 and Φ[δf ] denotes the gravitational po-
tential "sourced" (cf. eq. 2) by the perturbation δf . If one observes that E is an even
function of v, that the Poisson bracket is odd under velocity inversion, and that Φ[δf−]
vanishes identically, it is clear that eq. (20) is equivalent to the coupled system
∂tδf+ − {E, δf−} = 0
and
∂tδf− − {E, δf+} − {Φ[δf+], f0} = 0.
(21)
However, if one differentiates the second of these relations with respect to t, and uses the
first to eliminate ∂tδf+, it follows that
∂2
t δf− = {E, {E, δf−}} + {Φ[{E, δf−}], f0} ≡ FE Aδf−,
(22)
where A denotes a linear operator. One can then show that, given the identification of δf−
and ∂tδf− as conjugate variables, the equation
(−FE)−1∂2
t δf− = −Aδf−
(23)
can be derived from the Hamiltonian
1
2 Z
bH =
2 Z
=
1
dΓ
(−FE)
(∂tδf−)2 +
dΓ
(−FE)
(∂tδf−)2 +
1
2 Z dΓ δf−Aδf−
2Z
(−FE )
dΓ
1
{E, δf−}2 −
G
2 Z dΓ Z dΓ′ {E, δf−} {E ′, δf ′
x − x′
−}
. (24)
The connection between bH and the energy δ(2)H associated with a small symplectic per-
turbation is discussed in Kandrup (1989). In particular, one can show that bH > 0 for all
δf− if and only if δ(2)H > 0 for all symplectic perturbations.
The fact that A is a symmetric (i.e., hermitian) operator facilitates a proof that the
simple energy argument (cf. Laval, Mercier, and Pellat 1965) implies that the magnitude
of δf−, and hence δf , is bounded in time if A is a positive operator, so that δ(2)H > 0,
equilibrium f0(E) is linearly stable if and only if bH (and δ(2)H) is positive. Specifically, a
whereas the possibility of perturbations with R dΓδf−Aδf− < 0 implies the existence of
9
solutions that grow exponentially. This is easy to understand in the language of normal
modes. Since A is symmetric, it is clear that all solutions δf− ∝ exp(st) have s2 real, so
If A is positive, s2
that the evolution is either purely oscillatory or purely exponential.
must be negative, so that the modes are purely oscillatory. If, however, A is not a positive
operator, there exist modes with s2 > 0, which implies an exponential instability.4
This sort of normal mode expansion facilitates a simple geometric picture of an infinite-
dimensional configuration space of perturbations δf− which is (locally) embeddable in the
reduced γ-space. The equilibrium f0, which is necessarily an extremal point of the full
Hamiltonian H, satisfies δf− ≡ ∂tδf− ≡ 0. An arbitrary initial perturbation entails a ki-
netic energy K = R dΓ(−FE)−1(∂tδf−)2 which is necessarily positive and a potential energy
W = R dΓδf−Aδf− whose sign depends on the properties of A. If A is a positive opera-
tor, the evolution in configuration space involves a particle with "mass" (−FE)−1 moving
in an infinite-dimensional harmonic potential which corresponds to an upwards opening
paraboloid. Linear stability is therefore assured.
If, however, A is not always positive,
δf− ≡ 0 corresponds to a saddle point, rather than a local minimum, and the flow is lin-
early unstable.
In visualising all of this, there is the strong temptation to think of the normal modes
as being discrete, i.e., corresponding to honest square integrable eigenfunctions rather than
singular eigendistributions. This, however, is not necessarily justified.
Assuming completeness, one can always view any linear perturbation of an equilibrium
f0(E) with FE < 0 as a superposition of normal modes, writing δf as a formal sum
δf (x, v, t) = Xσ
Aσgσ(x, v)exp(iσt),
(25)
where gσ labels the eigenvector, σ is the corresponding frequency, which is necessarily
real, and Aσ is an expansion coefficient.5 Modulo largely unimportant technical details,
the modes then divide into two types, namely: (1) a countable set of discrete frequencies
belonging to the point spectrum, for which the corresponding eigenvectors are well-behaved
(e.g., square-integrable) eigenfunctions; and (2) a continuous set of frequencies belonging
to the continuous spectrum, for which the eigenvectors are singular eigendistributions.
The distinction between these two types of modes is extremely important (cf. Habib,
Kandrup, and Yip 1986). Because true eigenfunctions are nonsingular, they can in principle
be triggered individually, i.e., one can choose a reasonable initial δf which populates only
a single discrete mode. By contrast, because eigendistributions are singular, one cannot
sample a single continuous mode. Rather, any smooth δf sampling the continuous spectrum
must really be constructed as a wavepacket comprised of a continuous set of modes. The
4 In point of fact, one anticipates that, for this simple case, A is guaranteed to be positive. It
is believed (cf. Binney and Tremaine 1987) that any f0 depending only on E corresponds to a
spherically symmetric configuration; but assuming that the mass density ρ associated with f0 is
spherical one can prove that A is indeed positive (cf. Kandrup 1989).
5 Strictly speaking, this sum must be interpreted (cf. Riesz and Nagy 1955) as a Stiltjes integral.
10
important point then is that, when evolved into the future, such a wavepacket implies a
damping of coarse-grained observables like the density ρ. In other words, if the modes are
continuous there is a precise sense in which the perturbation "dies away" and the system
exhibits a coarse-grained approach towards the original equilibrium f0.
The physics here is analogous to what arises in ordinary quantum mechanics.
If, in
that setting, one considers a physical observable like angular momentum with a discrete
spectrum, one can construct well behaved eigenstates which, when evolved into the future,
maintain their coherence for all time: the only effect of the evolution is a coherently oscil-
lating phase. If, however, one considers an observable like position or linear momentum,
where the spectrum is continuous, this is no longer so. In this case, a normalisable initial
state must be constructed from a continuous set of singular eigendistributions, so that the
best one can do is build a localised (e.g., minimum uncertainty) wavepacket. However,
when evolved into the future such a wavepacket will necessarily spread because different
eigendistributions have different phase velocities.
It is this loss of coherence associated with the spreading of a wavepacket that corresponds
to (linear) Landau damping. In the context of plasma physics, Landau damping was derived
originally (Landau 1946) in a very different way, through the introduction of a Fourier-
Laplace transform and an analysis of poles in the complex plane. However, at least for the
electrostatic Vlasov equation (cf. Case 1959), i.e., the electrostatic analogue of the CBE,
the mathematical equivalence of these two pictures of Landau damping is well understood.
The physics underlying their equivalence is discussed in Kandrup (1998).
For the special case of perturbations of an homogeneous neutral plasma characterised
by an isotropic distribution of velocities that is everywhere nonvanishing, the modes can be
computed explicitly (cf. Case 1959), and one finds generically that
gσ(x, v) = exp(ik·x) gσ(v),
(26)
where gσ(v) is a singular eigendistribution involving a Dirac delta.
In this setting, an
examination of the perturbation associated with a given k-vector at a fixed phase space
point (x0, v0) yields no evidence of damping away. Rather, one finds persistent oscillations
∝ exp[ik·(x0 − v0t)]. This is simply a manifestation of the fact that, without the introduc-
tion of some coarse-graining, one cannot speak of the system returning to equilibrium. If,
however, a coarse-graining is implemented by integrating over any finite range of velocities,
one discovers that the resulting R d3v δf (x, v, t) will in fact damp away.
The obvious question, therefore, is: will perturbations δf of a generic equilibrium solu-
tion to the CBE correspond to discrete modes, continuous modes, or a combination of both?
Unfortunately, this is a difficult question to answer. It appears impossible to calculate the
modes explicitly for realistic equilibria, and a formal analysis is also difficult because the
operator T entering into the linearised equation ∂tδf = T δf is not elliptic and involves a
singular integral kernel. However, the normal modes can, and have, been computed for a
variety of nontrivial equilibrium solutions to the corresponding electrostatic Vlasov equation
11
(cf. van Kampen 1955, Case 1959), and the results derived thereby would seem suggestive.
Perhaps the most important result derived for the Vlasov equation is that discrete
modes are seemingly the exception, rather than the norm, arising only if the equilibrium
in question manifests nontrivial boundary conditions, e.g., the existence of a maximum
speed vm such that f (v) ≡ 0 for v > vm. In particular, one can prove that the modes are
always purely continuous if f0 is an analytic function of v in the complex plane.6 The best
known example of a nonempty point spectrum is the case of so-called van Kampen (1955)
modes, which arise precisely in those configurations where there is maximum velocity. In
the usual interpretation (cf. Stix 1962), Landau damping is understood as resulting from
a resonance between "particles" (the unperturbed f0) propagating with velocity v and a
"wave" (the perturbation δf ) that propagates with phase velocity c. Discrete van Kampen
modes correspond to perturbations which propagate with a phase velocity c for which
f0(c) = 0, so that no resonance is possible.
By analogy, one might therefore conjecture (cf. Habib, Kandrup, and Yip 1986) that,
for the gravitational CBE, most perturbations will in fact correspond to continuous modes
that damp away, but that some perturbations, especially longer wavelength disturbances
that probe the phase space boundaries of the system, could in fact correspond to discrete
modes. In this connection, it is interesting to note that there do in fact exist exact time-
dependent solutions to the CBE, seemingly appropriate for a system like a galaxy, that
exhibit finite amplitude undamped oscillations about some time-independent f0 (cf. Louis
and Gerhart 1988, Sridhar 1989). The interesting point, then is that in all these models the
time-independent f0 contains phase space "holes," i.e., regions in the middle of the occupied
phase space region where f0 → 0. Whether these sorts of solutions are generic, and whether
they could arise from reasonable initial conditions, is at the present unclear.
of discrete modes. For example, if one considers the function x(t) = P29
Finally, it should be noted that, in point of fact, one can in principle get (at least
temporary) phase mixing or loss of coherence even for the much simpler case of a finite set
p=10 cos(0.1pt) over
the finite interval 0 < t < 1000, one infers a rapid damping of the initial coherent excitation
with x = 20 to a much smaller value oscillating about x = 0 with typical amplitude x ∼ 1.
If, however, the evolution is tracked for a somewhat longer time one finds that the initial
coherence is regained. An infinite set of continuous modes differs from this toy model in two
important ways, namely (1) the recurrence time is infinitely long and (2) it is impossible to
consider a smooth initial excitation that does not damp.
4. Nonlinear Stability and Global Evolution
Suppose, once again, that attention is focused on some linearly stable equilibrium f0(E)
6 The validity of Landau's original derivation of exponential damping actually relies on the
implicit assumption that f0 is analytic. If it is not, his manipulations of contours and evaluation of
poles cannot be justified.
12
with FE < 0, but that one is now interested in the effects of larger perturbations δf , i.e., the
problem of nonlinear stability. To the extent that the normal modes of the linear problem
remain complete, one can still envision evolution in terms of these modes, the important
point, however, being that, because of nonlinearities, the modes will now interact. This is,
e.g., the basis for the standard quasilinear analyses implemented in plasma physics, which
allow for the effects of the quadratic term ∇Φ·(∂δf /∂v) which is ignored when considering
linear perturbations.
Mode-mode couplings are important in that they facilitate the transfer of energy be-
tween different modes, which makes the physics more complicated. However, one might still
anticipate that, if the modes are continuous, Landau damping can and will occur. Because
of the interactions between modes, the simple model of a dispersing quantum mechanical
wavepacket is no longer directly applicable, but the basic phenomenon of loss of coherence
is robust. Indeed, there are many examples in nonlinear dynamics of flows satisfying non-
linear evolution equations where phase mixing occurs. It thus seems reasonable to suppose
that, when considering the nonlinear evolution of some perturbation δf , one will encounter
nonlinear Landau damping. For the case of an electrostatic plasma, nonlinear Landau
damping is a well known, and reasonably well understood, phenomenon (cf. Davidson 1972
and references cited therein). Indeed, there are simple geometries where the nonlinear evo-
lution can be computed explicitly in the context of a systematic perturbation expansion,
thus facilitating analytic formulae for exactly how this phenomenon works (cf. Montgomery
1963).
Mode-mode couplings can also lead to another important possibility, namely the onset
of chaos. Because f0 is a local energy minimum, one knows that any infinitesimal pertur-
bation δf will simply oscillate, each eigenvector corresponding to motion in a "direction"
in configuration space that is orthogonal to the motion of all the other eigenvectors. This
implies that, for the fixed point f0, the Lyapunov exponents, which were defined in eq.
(13) as probing the average linear instability of the orbit generated from some initial f (0),
must all vanish identically. One might anticipate further that, when evolved into the future,
other phase space points sufficiently close to f0 will also correspond to regular orbits with
vanishing Lyapunov exponents. Thus, e.g., for finite-dimensional systems one knows that
there is a regular phase space region of finite measure surrounding every stable periodic or-
bit. However, for sufficiently large δf , where mode-mode couplings become significant and
the motion cannot be well approximated by orthogonal harmonic oscillations, one might
anticipate that many, if not all, perturbations will evolve chaotically. If true, this would
suggest that a "typical" perturbation with δH = H[f0 + δf ] − H[f0] will evolve ergodically
on (some subset of) the constant energy hypersurface in the γ-space with energy H[f0 + δf ].
This idea of the onset and development of chaos is an infinite-dimensional generalisa-
tion of what is typically found when considering the motion of a point mass in a multi-
13
dimensional nonlinear potential which has only one extremal point, a global minimum.7
Low energy orbits sufficiently close to the pit of the potential move in what is essentially
a harmonic potential, so that their motion is regular. If, however, the energy is raised one
finds generically that, unless the motions in different directions remain completely decou-
pled, there is an onset of global stochasticity which leads, for sufficiently high energies, to
well developed chaotic regions.
This configuration space description is not appropriate when considering generic equi-
libria, where the energy bH associated with a small perturbation cannot be written easily as
a functional of conjugate variables, and there is no guarantee that bH can be written as a
simple sum of kinetic and potential contributions, K and W. Modulo technical details, one
might expect that canonical phase space coordinates do exist, at least in principle, but the
energy bH associated with the tangent dynamics could in general be an arbitrary quadratic
functional bH[q, p] of the conjugate variables q and p. Moreover, even for the simple model
of an equilibrium f0(E) with FE < 0, it may not be possible to extend the canonical de-
scription to allow for arbitrarily large perturbations δf . One really needs to return to a full
phase space description.
As discussed in Section 3, if for some equilibrium f0 the second variation δ(2)H is
positive for all δf , a linearised perturbation corresponds in phase space to stable motion
on an upwards opening infinite-dimensional paraboloid. As long as this surface remains
convex, one would anticipate that stability will persist and, as such, one would expect
intuitively that the equilibrium could remain nonlinearly stable even for small but finite δf .
The normal modes of the linearised problem become coupled, but the geometric argument
for stability should remain valid. In particular, one can presumably visualise the evolution
of δf as involving nonlinear phase space oscillations about the equilibrium point f0.
If, however, f0 corresponds to a stable saddle, one might suppose that even the smallest
nonlinearities could trigger an instability (cf. Moser 1968, Morrison 1987). Thus, e.g., for
the simple toy model of two stable oscillators described by eq. (19), it is possible to trigger
an instability by introducing even very tiny mode-mode couplings which allow energy to be
transferred between modes. Indeed, as noted by Cherry (1925), if the two frequencies are
in an appropriate resonance, e.g., ω2
1, the introduction of a simple cubic coupling
implies that initial data arbitrarily close to x1 = v1 = x2 = v2 = 0 can lead to solutions in
which x1, x2, v1, and v2 all diverge in a finite time. If true, this expectation about saddle
points would suggest that, even though they can be linearly stable, they cannot represent
reasonable candidate equilibria in terms of which to model real astronomical objects.
2 = 2ω2
If a linearly stable f0 corresponds to a unique extremal point in the γ-space, the surface
which near f0 is a paraboloid will remain upwards opening even if δf is very large, so
that stability should persist for arbitrarily large perturbations. In other words, one would
7Even order truncations of the Toda potential (cf. Kandrup and Mahon 1994) provide a simple
two-dimensional example.
14
expect that the equilibrium f0 is globally stable: In this case, any phase space deformation
δf increases the energy, and the evolution of an initial δf (0) will involve nonlinear phase
space oscillations around the unique stable fixed point.
If, however, there exist multiple extremal points in the γ-space, each corresponding
to a local energy minimum, the situation is much more complicated.
In this case, one
would anticipate that, for sufficiently large δf , the distribution function can actually be
transferred from the "basin of attraction" of one equilibrium f0 to the "basin" of some
other f1. In other words, the evolution of δf (0) could yield oscillations around f1, rather
than f0. By suitably fine-tuning the perturbation, one can in principle displace the system
from any one basin to any other. However, by analogy with the behaviour observed in finite-
dimensional systems, one might expect generically that, if the perturbation is sufficiently
large, its motion can be interpreted as involving nonlinear phase space oscillations about
the global energy minimum. To the extent that this is true, one would anticipate that a
sufficiently large perturbation will tend generically to push f into the "basin of attraction"
of the equilibrium f0 that corresponds to a global energy minimum.
If one considers an initial perturbation δf (0) that is sufficiently large, the subsequent
evolution will in general be almost completely unrelated to the initial equilibrium f0 and, as
such, the way in which one visualises the evolution δf (0) is really no different from the way
in which one can, and arguably should, envision the evolution of a generic f (0). In other
words, the physical picture described above can be used equally well to visualise generic
flows associated with the initial value problem, the only difference being that, in general,
one may know nothing at all about what time-independent equilibria f0 actually exist.
Specification of an initial f (0) fixes the values of all the Casimirs for all times, thus
determining γ, the reduced infinite-dimensional phase space which constitutes the natural
arena of physics. This f (0) also fixes the numerical value of the conserved energy H and, as
such, determines the constant energy hypersurface in the γ-space to which the flow is nec-
essarily restricted. By analogy with finite-dimensional Hamiltonian systems (cf. Kandrup
and Mahon 1994) one might expect that, when evolved into the future, f (0) will exhibit a
coarse-grained approach towards an invariant measure on this hypersurface, i.e., a suitably
defined microcanonical distribution. If the flow associated with f (0) is chaotic, one might
anticipate an approach towards this invariant measure that is exponential in time. If, alter-
natively, the flow is regular, one might instead expect a power law approach. However, in
either case one might anticipate an approach towards a "phase-mixed" invariant measure.
In this context, the crucial question is then: to what extent can this invariant measure be
interpreted as corresponding to a distribution function f executing phase space oscillations
about one or more equilibrium solutions f0?
It is easy to see that, in the γ-space, there must exist one or more extremal points
with δ(1)H = 0, these corresponding to equilibrium solutions f0 for which all the Casimirs
share the same values as the Casimirs associated with f (0). Indeed, one knows that, for
sufficiently smooth initial data, the CBE admits global existence (cf. Pfaffelmoser 1992,
15
Schaeffer 1991), so that δf cannot diverge and, presumably, the Hamiltonian is bounded
from below. However, this implies that there must exist at least one f0, namely the global
energy minimum (although in principle the global minimum could be degenerate). The
question therefore becomes: in the basin of which f0 (or f0's) does the flow reside?
In principle, the evolved distribution function f could execute phase space oscillations
about any f0 with lower energy, which one presumably depending on the initial f (0). How-
ever, one might conjecture that, if the initial f (0) is sufficiently far from any equilibrium f0,
it will execute oscillations around the global minimum f0. The initial f (0) cannot exhibit
a pointwise approach towards this, or any other, f0. However, one might expect that, in
general, the initial deviation δf (0) = f (0) − f0 will exhibit nonlinear Landau damping so
that, in terms of observables like the density ρ, δf does indeed "die away," and one can
speak of a coarse-grained approach towards the equilibrium f0.
5. Conclusions and Unanswered Questions
The aim of this paper is to suggest a potentially fruitful way in which to visualise flows
described by the CBE and, in particular, the expected coarse-grained approach towards
an equilibrium. No claim is made regarding mathematical rigor, and it is not clear that
all the details are completely correct. However, the viewpoint developed here does have
the advantage that it incorporates what is known rigorously about the CBE, and that it
provides a framework in terms of which to pose precise, well defined questions.
In this
context, there are at least three basic questions which, if answered satisfactorily, would
yield important insights into the physical properties of a flow generated by the CBE:
1. Will generic initial conditions exhibit effective Landau damping, thus allowing one to
speak of an efficient coarse-grained evolution towards some equilibrium f0? In the context
of linear Landau damping, the answer to this question depends on the spectral properties of
the linearised evolution equation. If the modes are all continuous, every initial perturbation
will eventually phase mix away, so that physical observables like the density will damp
to zero.
If, however, some of the modes are discrete, it is possible to construct initial
perturbations that do not damp away. At the present time, it is not clear whether, for
realistic galactic models, the spectrum is purely continuous, although the investigation of
various toy models is currently underway (Lynden-Bell 1997, private communication).
To the extent that N -body simulations are reliable and that, for sufficiently large N ,
they capture the same physics as the CBE, the fact that most initial conditions yield
an efficient approach towards some statistical equilibrium can be interpreted as evidence
that nonlinear Landau damping is in general very effective. However, there do exist toy
models like one-dimensional gravity where one ends up with undamped oscillations. For
example, the evolution of counterstreaming initial conditions in one-dimensional systems
(either gravitational or electrostatic) can lead to a final state which corresponds seemingly
to a distribution function f exhibiting finite amplitude undamped oscillations about a (near-
16
) equilibrium f0 (cf. Mineau, Feix, and Rouet 1990). This toy model actually corroborates
the physical intuition described in this paper in the sense that, as one would expect, the
phase space contains a large "hole," i.e., a region where f0 → 0. Whether or not analogous
results obtain for two- and three-dimensional systems is as yet unclear, although the problem
is currently under investigation (Habib, Kandrup, Pogorelov, and Ryne, work in progress).
2. Are functional Lyapunov exponents the "right" way in which to identify chaos in infinite-
dimensional systems and, assuming that they are, will a generic flow associated with the
CBE be chaotic? Given this definition, will standard results from finite-dimensional chaos
remain at least approximately valid? Although not proven for generic finite-dimensional
systems, there is the physical expectation that, when evolved into the future, a chaotic
initial condition will evolve towards an invariant distribution on a time scale that is related
somehow to the spectrum of Lyapunov exponents. This implies however that, at asymptot-
ically late times, one can visualise the flow as densely filling a chaotic phase space region
of finite measure. Assuming, however, that this is true, the Ergodic Theorem provides im-
portant information about the statistical properties of the flow, implying the equivalence of
time and phase space averages (cf. Lichtenberg and Lieberman 1992).
One other point about chaos in the CBE should be stressed: The definition proposed
in this paper is, at least superficially, completely decoupled from the (also interesting)
question of whether individual orbits in a self-consistent potential generated from the CBE
are, or are not, chaotic. This latter question refers to the behaviour of nearby trajectories
in the six-dimensional particle phase space. The "natural" definition of chaos for the CBE
should presumably reflect properties of the flow in the infinite-dimensional phase space of
distribution functions.
3. For a specified initial f (0), towards which equilibrium f0 will the system evolve? Given
f (0), one can compute the numerical value of all possible Casimirs, thus identifying explicitly
the γ-space to which the evolution is restricted. The obvious problem, then, is to identify
all time-independent equilibria f0 in γ and to determine which initial conditions correspond
to which equilibria. Although unquestionably difficult, this is a problem that is both well
defined mathematically and well motivated physically. Finding all equilibria is equivalent
mathematically to finding all extremal points in γ. However, to the extent that one chooses
to visualise the flow as involving oscillations in the γ-space, there is no question physically
but that the extremal points define "basins of attraction" associated with the oscillations.
The basic points described in this paper are easily summarised:
1. The CBE is a Hamiltonian system, albeit an unusual one. The fundamental dynamical
variable is the distribution function f , not the particle x's and v's; and it is not always
possible (at least easily) to identify canonically conjugate variables.
2. Because the CBE is Hamiltonian, there can be no pointwise approach towards equilibrim.
The best for which one can hope is a coarse-grained approach towards equilibrium.
3. Even though the phase space γ associated with the dynamics is infinite-dimensional, one
17
might expect that much of one's intuition from finite-dimensional systems remains valid.
In particular, one might anticipate an asymptotic approach towards an invariant measure,
and one might hope to make meaningful distinctions between regular and chaotic flows.
4. The phenomenon normally designated as linear Landau damping can be interpreted as a
phase mixing of a continuous set of normal modes. Whether a small initial perturbation will
always eventually Landau damp/phase mix away depends on whether the normal modes
for the linearised perturbation equation are discrete or continuous.
5. To the extent that one's ordinary intuition about finite-dimensional phase spaces remains
approximately valid, the evolution of generic initial data should be interpreted as involving
nonlinear (phase space) oscillations about one or more energy extremals, which correspond
to time-independent equilibria f0. The phenomenon of violent relaxation should thus be
interpreted as nonlinear phase mixing/Landau damping which, if efficient, will facilitate a
coarse-grained approach towards equilibrium.
Acknowledgments
I am pleased to acknowledge useful discussions, collaborations, and correspondence with
Salman Habib, Donald Lynden-Bell, Bruce Miller, Phil Morrison, and Daniel Pfenniger. I
am grateful to Barbara Eckstein for comments on the exposition. Work on this manuscript
began while I was a visitor at the Observatoire de Marseille, where I was supported by the
C.N.R.S., and continued during a visit to the Aspen Center for Physics. Limited financial
support was provided by the National Science Foundation grant PHY92-03333.
18
Antonov, V. A. 1960, Astr. Zh. 37, 918 (English trans. in Sov. Astr. -- AJ 4, 859 [1961]).
Arnold, V. I. 1989, Mathematical Methods of Classical Mechanics, 2nd ed., Springer-Verlag,
Berlin.
Bartholomew, P. 1971, MNRAS 151, 333.
Binney, J. and Tremaine, S. 1987, Galactic Dynamics, Princeton University Press, Prince-
ton.
Case, K. M. 1959, Ann. Phys. (NY) 7, 349.
Cherry, T. M. 1925, Trans. Cambridge Philos. Soc. 23, 199.
Davidson, R. C. 1972, Methods of Nonlinear Plasma Theory, Academic Press, New York.
Habib, S., Kandrup, H. E., and Yip, P. F. 1986, Ap. J. 309, 176.
Kandrup, H. E. 1990, Ap. J. 351, 104.
Kandrup, H. E. 1991a, Ap. J. 370, 312.
Kandrup, H. E. 1991b, Ap. J. 380, 511.
Kandrup, H. E. 1998, Ann. N. Y. Acad. Sci., in press.
Kandrup, H. E. and Mahon, M. E. 1994, Phys. Rev. E 49, 3735.
Kandrup, H. E. and Morrison, P. J. 1993, Ann. Phys. (NY) 225, 114.
Landau, L. D. 1946, Soviet Phys. - JETP 10, 25.
Landau, L. D. and Lifshitz, E. M. 1960, Mechanics, Addison-Wesley, Reading, MA.
Laval, G., Mercier, C., and Pellat, R. 1965, Nucl. Fusion 5, 156.
Lichtenberg, A. J. and Lieberman, M. A. 1992, Regular and Chaotic Dynamics, Berlin,
Springer, 2nd ed.
Louis, P. D. and Gerhart, O. 1988, MNRAS 233, 337.
Lynden-Bell, D. 1967, MNRAS 136, 101
Lynden-Bell, D. and Sanitt, N. 1969, MNRAS 143, 167.
Maoz, E. 1991, Ap. J. 275, 687.
Mineau, P., Feix, M. R., and Rouet, J. L. 1990, Astron. Astrophys. 228, 344.
Montgomery, D. 1963, Phys. Fluids A6, 1109.
Moser, J. K. 1968, Mem. Am. Math. Soc. 81, 1.
Morrison, P. J. 1980, Phys. Lett. A 80, 383.
Morrison, P. J. 1987, Z. Naturforsch. 42a, 1115.
Morrison, P. J. and Eliezur, S. 1986, Phys. Rev. A 33, 4205.
Perez, J. and Aly. J.-J. 1996, MNRAS 280, 689.
Perez, J., Alimi, J.-M., Aly, J.-J., and Scholl, H. 1996, MNRAS 280, 700.
Pfaffelmoser, K. 1992, J. Diff. Eqns. 95, 281.
Riesz, F. and Sz.-Nagy, B. 1955, Functional Analysis, Ungar, New York.
Schaeffer, J. 1991, Commun. Part. Diff. Eqns. 16, 1313
Sridhar, S. 1989, MNRAS 201, 939.
Stix, T. H. 1962, The Theory of Plasma Waves, McGraw-Hill, New York.
van Kampen, N. G. 1955, Physica 21, 949.
19
|
astro-ph/0701313 | 2 | 0701 | 2007-01-25T07:55:20 | On the Primordial Helium Abundance and the DeltaY/DeltaO Ratio | [
"astro-ph"
] | We present a review on the determination of the primordial helium abundance, Yp, based on the study of hydrogen and helium recombination lines in extragalactic H II regions. We also discuss the observational determinations of the increase of helium to the increase of oxygen by mass Delta Y/Delta O, and compare them with predictions based on models of galactic chemical evolution. | astro-ph | astro-ph |
**FULL TITLE**
ASP Conference Series, Vol. **VOLUME**, **YEAR OF PUBLICATION**
**NAMES OF EDITORS**
On the Primordial Helium Abundance and the ∆Y /∆O
Ratio
Manuel Peimbert1, Valentina Luridiana2, Antonio Peimbert1, & Leticia
Carigi1
Abstract.
We present a review on the determination of the primordial helium abun-
dance, Yp, based on the study of hydrogen and helium recombination lines in
extragalactic H ii regions. We also discuss the observational determinations of
the increase of helium to the increase of oxygen by mass ∆Y /∆O, and compare
them with predictions based on models of galactic chemical evolution.
1. Why Yp?
The determination of Yp is important for at least the following reasons: a)
It is one of the pillars of Big Bang cosmology and an accurate determination
of Yp permits to test the standard Big Bang nucleosynthesis (SBBN), b) the
models of stellar evolution require an accurate initial Y value; this is given by
Yp plus the additional Y produced by galactic chemical evolution, which can be
estimated based on the ∆Y /∆O ratio, c) the combination of Yp and ∆Y /∆O
is needed to test models of galactic chemical evolution, d) to test solar models
it is necessary to know the initial solar abundances, which are different to the
photospheric ones due to diffusive settling, this effect reduces the helium and
heavy element abundances in the solar photosphere relative to that of hydrogen,
the initial solar abundances can be provided by models of galactic chemical
evolution, e) the determination of the Y value in metal poor H ii regions requires
a deep knowledge of their physical conditions, in particular the Y determination
depends to a significant degree on their density and temperature distribution,
therefore accurate Y determinations combined with the assumption of SBBN
provide a constraint on the density and temperature structure of H ii regions.
2. Recent Determinations of Yp
Previous reviews on Yp determinations have been presented by Peimbert et al.
(2003) and Luridiana (2003). Recent determinations of Yp are those by Izotov et al.
(1999), Izotov & Thuan (2004), Izotov et al. (2006), Luridiana et al. (2003),
Olive & Skillman (2004), Fukugita & Kawasaki (2006), and Peimbert et al. (2007).
1Instituto de Astronom´ıa, Universidad Nacional Aut´onoma de M´exico; Apdo. postal 70 -- 264;
Ciudad Universitaria; M´exico D.F. 04510; Mexico.
2Instituto de Astrof´ısica de Andaluc´ıa (CSIC), Camino bajo de Hu´etor 50, 18008 Granada,
Spain.
1
2
Peimbert et al.
A critical discussion of these determinations has been presented by Peimbert et al.
(2007). Most of the differences among these determinations are due to system-
atic effects.
3. Error Budget
The error budget of the Yp determination by Peimbert et al. (2007) is presented
in Table 1. In this table the sources of error are listed in order of importance.
The error budgets of Yp determinations by other groups are different to that
by Peimbert et al. (2007) for many reasons, each error budget depends on the
sample of H ii regions used and on the treatment given to the different sources
of error.
Table 1.
Error budget in the Yp determination
Problem
Estimated error
Collisional Excitation of the H i Lines
Temperature Structure
O (∆Y /∆O) Correction
Recombination Coefficients of the He i Lines
Density Structure
Underlying Absorption in the He i Lines
Recombination Coefficients of the H i Lines
Underlying Absorption in the H i Lines
Ionization Structure
Collisional Excitation of the He i Lines
Reddening correction
Optical Depth of the He i Triplet Lines
He i and H i Line Intensities
±0.0015
±0.0010
±0.0010
±0.0010
±0.0007
±0.0007
±0.0005
±0.0005
±0.0005
±0.0005
±0.0005
±0.0003
±0.0003
4. The Four Main Sources of Error
The most important source of error is the collisional excitation of Balmer lines.
Neutral hydrogen atoms in excited states may form not only by the usual pro-
cess of H+ recombination, but also through collisions of neutral hydrogen with
electrons. The recombination cascade that follows such excitations contributes
to the observed intensity of Balmer lines, mimicking a larger relative hydrogen
abundance and, hence, a smaller helium abundance. To estimate this contri-
bution it is necessary to have a tailored photoionization model for each object
that fits properly the temperature structure. The contribution to the Balmer
line intensities depends strongly on the temperature: therefore this effect, and
consequently the associated error in its estimate, increases for H ii regions of
lower metallicity and consequently higher temperature.
The second most important source of error is the temperature structure.
Most determinations neglect the possible presence of temperature variations
across the H ii region structure and assume that T (O iii) is representative of
the zone where the He i recombination lines form. However, other temperature
Primordial Helium
3
determinations based on different diagnostics yield lower values; furthermore,
photoionization models do not predict the high T (O iii) values observed. These
results indicate that temperature variations are indeed present in H ii regions,
and this result should be included in the Y determination (Peimbert et al. 2002).
The best procedure to take into account the temperature structure is to self-
consistently determine T (He ii) from a set of He i lines by means of the maximum
likelihood method. The Y abundances derived from T (He ii) are typically lower
by about 0.0030 − 0.0050 than those derived from T (O iii). The difference
between both temperatures does not have a significant trend with the metallicity
of the H ii region, hence the systematic error introduced by the use of T (O iii) in
the Y determination is similar for objects with different metallicities. The error
quoted under "Temperature Structure" in Table 1 is the residual error due to
the uncertainty in the T (He ii) determinations of the dataset by Peimbert et al.
(2007).
The third most important source of error is the extrapolation of the derived
Y values to zero metallicity through the ∆Y /∆O ratio. This problem will be
discussed in the next section. The fourth most important source of error is the
uncertainty on the recombination coefficients of the He i lines.
5. ∆Y /∆O from Models and Observations
To determine the Yp value it is customary to use the Y values of a set of O poor
galaxies and to extrapolate the Y values to the case of O = 0 using the following
equation:
Yp = Y − O
∆Y
∆O
,
(1)
where O is the oxygen abundance per unit mass. To obtain an accurate Yp value,
a reliable determination of ∆Y /∆O for O-poor objects is needed.
The ∆Y /∆O value derived by Peimbert et al. (2000) from observational re-
sults and models of chemical evolution of galaxies amounts to 3.5 ± 0.9. More
recent results are those by Peimbert (2003) who finds 2.93 ± 0.85 from ob-
servations of 30 Dor and NGC 346, and by Izotov et al. (2006) who, from the
observations of 82 H ii regions, find ∆Y /∆O = 4.3 ± 0.7. We have recomputed
the value by Izotov et al. considering two systematic effects not considered by
them: the fraction of oxygen trapped in dust grains, which we estimate to be
10% for objects of low metallicity, and the increase in the O abundances due
to explicit taking into account the presence of temperature fluctuations, which
for this type of H ii regions we estimate to be about 0.08 dex (Relano et al.
2002). From these considerations we obtain for the Izotov et al.
sample a
∆Y /∆O = 3.2 ± 0.6.
From chemical evolution models of galaxies it is found that ∆Y /∆O depends
on: the stellar yields, the initial mass function, the star formation rate, the age,
and the O value of the galaxy in question. Models with substantial outflows of
O-rich material can produce large ∆Y /∆O ratios but they are ruled out by the
low C/O values observed in irregular galaxies (Carigi et al. 1995, 1999, 2006).
Carigi & Peimbert (2007) have produced chemical evolution models of the
following types: closed box, inflow of gas, and outflow of gas of well-mixed
4
Peimbert et al.
Figure 1.
Two chemical evolution models for NGC 6822 with different gas
infall and outflow histories, and the same star formation rate derived from
observations. The models are the 7L and 8S (continuous and dashed lines, re-
spectively) studied by Carigi et al. (2006). The first panel shows the increase
of the helium and oxygen abundances relative to the primordial values, the
second panel shows the gaseous content as a function of time, which is widely
different for the two models, and the third panel shows the star formation
history, which is the same one in both models.
material. They find that ∆Y /∆O is practically constant for models with the
same IMF, the same age, the same star formation history, and an O abundance
smaller than ∼ 4×10−3. They find also that 2.4 < ∆Y /∆O < 4.1 for models
with different star formation histories and different values of the upper mass
limit of the IMF. The results derived from the chemical evolution models are in
very good agreement with the observations mentioned above.
Based on the observations and models discussed, Peimbert et al. (2007)
adopted a value of ∆Y /∆O = 3.3 ± 0.7 in the computation of the error budget
presented in Table 1 and the Yp value presented in Table 2.
Primordial Helium
5
For models of galactic chemical evolution that reach values of O > 4 × 10−3
at present time, the ∆Y /∆O ratio of the interstellar medium increases with the
O abundance due to two effects: the helium production by low and intermediate
mass stars and the increase of the helium yield of massive stars due to stellar
winds.
As an example of the minor role that well mixed outflows play in the
chemical evolution of galaxies in Figure 1 we present the ∆Y versus ∆O be-
havior for two chemical evolution models of NGC 6822 (Carigi et al. 2006;
Carigi & Peimbert 2007). These models have the same SFH, which was de-
rived from observations, but are drastically different in their gas flow histories
and show practically the same ∆Y /∆O behavior.
To compute stellar evolution models with O < 4 × 10−3 we propose to use
the following relation for the initial Y and O abundances
Y = 0.2474 ± 0.0028 + (3.3 ± 0.7)O.
(2)
For O > 4 × 10−3 (Carigi & Peimbert 2007), Y increases faster with the increase
of O than in the previous equation.
The ratio ∆Y /∆O can also be expressed in terms of ∆Y /∆Z if the fraction
of O/Z is known. Based on observations of galactic and extragalactic H ii
regions, we propose to assume that, for models with O < 4 × 10−3, O constitutes
55% ± 5% of the total Z value, implying ∆Y /∆Z = 2.0 ± 0.6. For O > 4 × 10−3
the fraction of Z due to O decreases due to the increase of the C/O, N/O
and Fe/O ratios with the increase of O. From the chemical composition of the
Orion nebula and M17 it is found that the fraction of Z due to O drops to
about 42% for 0.0057 < O < 0.0082 (e.g. Peimbert 2003; Esteban et al. 2004;
Garc´ıa-Rojas et al. 2007, and references therein).
Previous determinations of the ∆Y /∆Z ratio based on models or observa-
tions have been in the 1.0 to 6.0 range (e.g. Peimbert 1995; Fukugita & Kawasaki
2006, and references therein). It is interesting to note that 25 years ago Chiosi & Matteucci
(1982) determined a value of ∆Y /∆Z ∼ 2.0, in excellent agreement with the re-
cent results by Carigi & Peimbert (2007).
6. Discussion
In Table 2 we present some of the best Yp determinations of the last few years.
The Yp values and their statistical errors are the ones presented in the original
papers. After the statistical error we list the systematic error estimated by us,
which depends on one or more of the following sources: a) the change in the
published emissivities of the He i lines (Porter et al. 2005); b) the change in the
published collisional excitation coefficients of the H i lines (Anderson et al. 2000,
2002); and c) the temperature structure of the H ii region. Adopting the new
He i emissivities, the Y determination of individual H ii regions is increased by
about 0.004. The change in the H i collisional excitation coefficients goes in the
same direction (Peimbert et al. 2007), although in this case the size of this effect
varies from object to object and a tailored photoionization model for each object
is needed to obtain a good estimate of it. Both effects produce an increase in the
Yp determination: for the sample of H ii regions used by Peimbert et al. (2007),
the increase in Yp due to the adoption of the new He i emissivities amounts to
6
Peimbert et al.
about 0.0040, while the increase due to the adoption of the new H i collisional
excitation coefficients amounts to 0.0025. Finally, the Y value of individual H ii
regions decreases when the temperature structure of the H ii region is taken
into account, since in the self-consistent solutions for all the observed He i line
intensities, the lower T (He ii) values imply higher densities, the higher densities
produce a higher collisional contribution to the He i intensities, and consequently
lower helium abundances. For the objects in the sample used by Peimbert et al.
(2002) and Peimbert et al. (2007), this effect decreases the Y determinations by
amounts ranging from 0.003 to 0.009.
The Yp determination by Izotov et al. (1999) is affected by all three of
the above sources of systematic error; those by Luridiana et al. (2003) and
Izotov & Thuan (2004) are affected by sources a) and b), while the one by
Izotov et al. (2006) is affected by source b). The disagreement between the
Yp derived by Luridiana et al. (2003) with the Yp derived from WMAP under
the assumption of SBBN implies the need for "new physics". The new physics
needed to reconcile the two Yp values turned out to be the "new atomic physics"
by Anderson et al. (2000, 2002) and Porter et al. (2005). With the new physics
Peimbert et al. (2007) found agreement, within the observational errors, be-
tween the observed Yp value and that derived from the WMAP results under
the assumption of SBBN.
The Yp derived by Peimbert et al. (2007) together with SBBN implies that
Ωbh2 = 0.02054 ± 0.00639 (Steigman 2006a,b), where Ωb is the baryon closure
parameter, and h is the Hubble parameter. This value is in excellent agreement
with the value derived by Spergel et al. (2006) from the WMAP results under
the assumption of SBBN, which amounts to Ωbh2 = 0.02233 ± 0.00082.
The comparison of the Yp derived by Peimbert et al. (2007) with the Yp
derived by Spergel et al. (2006) from the WMAP data together with the as-
sumption of SBBN provides strong constraints for the study of non SBBN (e.g.
Cyburt et al. 2005; Coc et al. 2006, and references therein).
In Table 2 we also present our Yp prediction for 2010. We consider that in
the next few years it will be possible to reduce the statistical errors in the Yp
determination to about 0.0020 by obtaining a new set of observations of brighter
and slightly O-richer H ii regions than the ones that have been used so far. A
more extensive discussion of the relative advantages of these H ii regions with
respect to more metal-poor ones can be found in Peimbert et al. (2007).
7. Summary and conclusions
In this contribution we have presented some recent determinations of Yp and
discussed the reasons underlying the differences between them. For the most
recent determination, the one by Peimbert et al. (2007), we have presented the
error budget in terms of thirteen different sources of error. The ∆Y /∆O ratio,
which enters as a crucial factor in one of the three main sources of error, has
been discussed in the light of recent observations and models of galactic chemical
evolution.
The Yp determinations by different groups are slowly converging among
them as systematic errors are progressively identified and corrected for. The need
for "new physics" that had been suggested by recent results (e.g., Luridiana et al.
Primordial Helium
7
Table 2.
Primordial helium abundance values a
Source
Yp
Izotov et al. (1999), this work
Luridiana et al. (2003), this work
Izotov & Thuan (2004), this work
Izotov et el. (2006), this work
Peimbert et al. (2007)
Prediction (2010), this work
Spergel et al. (2006)
0.2452 ± 0.0015 ± 0.0100
0.2391 ± 0.0020 ± 0.0070
0.2421 ± 0.0021 ± 0.0075
0.2462 ± 0.0025 ± 0.0040
0.2474 ± 0.0028
0.2??? ± 0.0020
0.2482 ± 0.0004
a Direct Yp determinations based on observations of H ii regions, with the exception
of that by Spergel et al. (2006), which is based on the baryon to photon ratio derived
from WMAP and the assumption of SBBN.
2003) seems now to be fulfilled by new atomic physics, i.e. the atomic data by
Anderson et al. (2000, 2002) and Porter et al. (2005). On the other hand, the
temperature structure of H ii regions is still a source of systematic error if an
appropriate scheme for temperature is not adopted. The proper temperature to
determine the helium abundance is T (He ii), derived self-consistently from the
intensities of the helium lines. Adopting this temperature, the Yp value derived
from H ii regions agrees with the Yp derived from the WMAP data under the
assumption of SBBN. On the other hand, the use of T (O iii) to determine the Y
values from H ii regions produces Yp values from 0.003 to 0.006 higher than those
found adopting T (He ii), that is Yp values more than 1σ higher than the one
predicted by the WMAP observations combined with the assumption of SBBN.
It is important to continue the effort on the study of the physical conditions
in H ii regions, this effort will permit us to lower the error on the Yp determi-
nation, which in turn will permit us to improve our knowledge on the possible
importance of non SBBN.
Acknowledgments. This work was partly supported by the CONACyT
grant 46904 and by the Spanish Programa Nacional de Astronom´ıa y Astrof´ısica
through the project AYA2004-07466. VL is supported by a CSIC-I3P fellowship.
References
Anderson, H., Ballance, C. P., Badnell, N. R., & Summers, H. P. 2000, Journal of
Physics B Atomic Molecular Physics, 33, 1255
Anderson, H., Ballance, C. P., Badnell, N. R., & Summers, H. P. 2002, Journal of
Physics B Atomic Molecular Physics, 35, 1613
Carigi, L., Col´ın, P., & Peimbert, M. 1999, ApJ, 514, 787
Carigi, L., Col´ın, P., & Peimbert, M. 2006, ApJ, 644, 924
Carigi, L., Col´ın, P., Peimbert, M., & Sarmiento, A. 1995, ApJ, 445, 98
Carigi, L., & Peimbert, M. 2007, ApJ, to be submitted
Chiosi, C., & Matteucci, F. 1982, A&A, 105, 140
Coc, A., Nunes, N. J., Olive, K. A., Usan, J.-P. & Vangion, E. 2006, astro-ph/0610733
Cyburt, R. H., Fields, B. D., Olive, K. A., & Skillman, E. 2005, Astroparticle Physics,
23, 313
8
Peimbert et al.
Esteban, C., Peimbert, M., Garc´ıa-Rojas, J., Ruiz, M. T., Peimbert, A., & Rodr´ıguez,
M. 2004, MNRAS, 355, 229
Fukugita, M., & Kawasaki, M. 2006, ApJ, 646, 691
Garc´ıa-Rojas, J., Esteban, C., Peimbert, A., Rodr´ıguez, M., Peimbert, M., & Ruiz, M.
T. 2007, RevMexAA, in press, astro-ph/0610065
Izotov, Y. I., Chaffee, F. H., Foltz, C. B., Green, R. F., Guseva, N. G., & Thuan, T. X.
1999, ApJ, 527, 757
Izotov, Y. I., Schaerer, D., Blecha, A., Royer, F., Guseva, N. G., & North, P. 2006,
A&A, 459, 71
Izotov, V. I., & Thuan, T. X. 2004, ApJ, 602, 200
Luridiana, V. 2003, proceedings of the XXXVIIth Moriond Astrophysics Meeting "The
Cosmological Model" eds. Y. Giraud-H´eraud, C. Magneville, J. Tran Thanh Van,
The Gioi Publishers (Vietnam), 159 (astro-ph/0209177)
Luridiana, V., Peimbert, A., Peimbert, M., & Cervino, M. 2003, ApJ, 592, 846
Olive, K. A., & Skillman, E. D. 2004, ApJ, 617, 29
Peimbert, A. 2003, ApJ, 584, 735
Peimbert, A., Peimbert, M., & Luridiana, V. 2002 ApJ, 565, 668
Peimbert, M. 2003, in The Light Element Abundances, ed. P. Crane (Springer), p. 165,
1995.
Peimbert, M., Luridiana, V., & Peimbert, A. 2007, ApJ, submitted (astro-ph/0701580)
Peimbert, M., Peimbert, A., Luridiana, V. & Ruiz, M. T. 2003, in Star Formation
through Time, ASP Conference Series, 297, 81
Peimbert, M., Peimbert, A., & Ruiz, M. T. 2000, ApJ, 541, 688
Porter, R. L., Bauman, R. P., Ferland, G. J. , & MacAdam, K. B. 2005, ApJ, 622L, 73
Relano, M., Peimbert, M., & Beckman, J. 2002, ApJ, 564, 704
Spergel, D. N. et al. 2006, astro-ph/0603449
Steigman, G. 2006a, Int. J. Mod. Phys. E, 15, 1, astro-ph/0511534
Steigman, G. 2006b, astro-ph/0606206
|
astro-ph/0301261 | 1 | 0301 | 2003-01-14T17:19:23 | Blue Straggler Stars: a direct comparison of Star counts and population ratios in six Galactic Globular Clusters | [
"astro-ph"
] | The central regions of six Galactic Globular Clusters (GGCs) (M3, M80, M10, M13, M92 and NGC 288) have been imaged using HST-WFPC2 and the ultraviolet (UV) filters (F255W, F336W). The selected sample covers a large range in both central density and metallicity ([Fe/H]). In this paper, we present a direct cluster-to-cluster comparison of the Blue Stragglers Stars (BSS) population as selected from (m_{255},m_{255}-m_{336}) Color Magnitude Diagrams (CMDs). We have found:(a) BSS in three of the clusters (M3, M80, M92) are much more concentrated toward the center of the cluster than the red giants; because of the smaller BSS samples for the other clusters we can only note that the BSS radial distributions are consistent with central concentration; (b) the specific frequency of BSS varies greatly from cluster to cluster.
The most interesting result is that the two clusters with largest BSS specific frequency are at the central density extremes of our sample: NGC 288 (lowest central density) and M80 (highest). This evidence together with the comparison with theoretical collisional models suggests that both stellar interactions in high density cluster cores and at least one other alternate channel operating low density GGCs play an important role in the production of BSS. We also note a possible connection between HB morphology and blue straggler luminosity functions in these six clusters. | astro-ph | astro-ph |
Submitted to the Astrophysical Journal
Blue Straggler Stars: a direct comparison of Star counts and
population ratios in six Galactic Globular Clusters1
Francesco R. Ferraro2, Alison Sills3, Robert T. Rood4, Barbara Paltrinieri5, Roberto
Buonanno5,6
ABSTRACT
The central regions of six Galactic Globular Clusters (GGCs) (M3, M80, M10,
M13, M92 and NGC 288) have been imaged using HST-WFPC2 and the ultra-
violet (UV) filters (F255W, F336W). The selected sample covers a large range in
both central density (log ρ0) and metallicity ([Fe/H]). In this paper, we present
a direct cluster-to-cluster comparison of the Blue Stragglers Stars (BSS) popu-
lation as selected from (m255, m255 − m336) Color Magnitude Diagrams (CMDs).
We have found: (a) BSS in three of the clusters (M3, M80, M92) are much more
concentrated toward the center of the cluster than the red giants; because of the
smaller BSS samples for the other clusters we can only note that the BSS radial
distributions are consistent with central concentration; (b) the specific frequency
of BSS varies greatly from cluster to cluster. The most interesting result is that
the two clusters with largest BSS specific frequency are at the central density
extremes of our sample: NGC 288 (lowest central density) and M80 (highest).
This evidence together with the comparison with theoretical collisional models
suggests that both stellar interactions in high density cluster cores and at least
one other alternate channel operating low density GGCs play an important role
in the production of BSS. We also note a possible connection between HB mor-
phology and blue straggler luminosity functions in these six clusters.
2Dipartimento di Astronomia, Universit`a di Bologna, via Ranzani 1, I -- 40126 Bologna, Italy;
fer-
[email protected]
3Department of Physics and Astronomy, McMaster University, 1280 Main Street West, Hamilton, ON,
L8S 4M1, Canada; [email protected]
4Astronomy Dept., University of Virginia, Charlottesville, VA 22903-0818, USA; [email protected]
5Osservatorio Astronomico di Roma, Via Frascati 33, 00040 Monte Porzio Catone, Italy
6Universit`a di Roma "Tor Vergata", Dip di Fisica, Via della Ricerca Scientifica, I-00133 Roma, Italy
-- 2 --
Subject headings: Globular clusters:
NGC 288 ); stars: evolution -- blue stragglers -- binaries: close;
individual (M3, M80, M10, M13, M92,
1.
Introduction
Stellar collisions are thought to be the dominant channels for the formation and evo-
lution of various types of unusual stellar objects and binary systems in the dense stellar
environments, such as the central regions of Galactic Globular Clusters (GGCs). Among the
possible collisionally induced populations, Blue Stragglers Stars (BSS) are surely the most
studied. BSS were first discovered by Sandage (1953) in M3. In recent years the realization
that BSS are the ideal diagnostic for a quantitative evaluation of the effects of dynamical
interactions inside star clusters has led to a remarkable burst of activity in the search for and
the systematic study of BSS in clusters. In addition, the Hubble Space Telescope (HST),
with its high angular resolution and UV imaging capabilities, has made it possible to search
for candidates BSS in the cores of highly concentrated GGCs.
Within this framework we are performing an extensive HST-UV survey in the cores of a
selected sample of GGCs searching for various products of binary evolution. In particular, we
are constructing an homogeneous database of BSS in the UV-bands. Some results have been
already published in a series of papers presenting the BSS content in individual clusters
(see Ferraro et al. 1997a, 1999a, 2001 and Paltrinieri et al. 1998, Bellazzini et al. 2002,
Ferraro, Paltrinieri & Cacciari, 1999, for a review). In this series of papers we empirically
demonstrated that UV-CMDs are the most appropriate planes for the study of BSS.
In this paper, we present a direct comparison of the BSS content in six intermediate to
high density (log ρ0 ∼ 2 -- 5.4) GGCs with intermediate to low metallicity: M13 (NGC 6205),
M10 (NGC 6254), M80 (NGC 6093), M3 (NGC 5272), M92(NGC 6341) and NGC 288. In
§3 we present the BSS samples discovered in each cluster on the basis of the UV-CMDs, and
we discuss the properties and the radial distributions of BSS with respect to normal cluster
stars. In §4 we present the comparison of the BSS distribution in the UV-CMDs with that
expected from theoretical collisional models.
1Based on observations with the NASA/ESA Hubble Space Telescope, obtained at the Space Telescope
Science Institute, which is operated by AURA, Inc., under NASA contract NAS5-26555
-- 3 --
2. Observations and data analysis
The data-set used in this work consist of a series of HST-WFPC2 exposures covering
the central regions of six GGCs: M3, M10, M13, M80, M92 and NGC 288.
In Table 1
we summarized the main characteristics of each cluster: metallicity ([Fe/H]) from Zinn &
West (1984), central density (ρ0 and mass (from Trager & Meylan 1993), distance (d) (from
Ferraro et al. 1999b), and central dispersion velocity (σ0) (from Trager & Meylan 1993). For
each cluster, the Planetary Camera (PC) was roughly centered on the cluster center (in the
case of NGC 288 the cluster center was outside the field of view of the PC - see Figure 1
by Bellazzini et al. 2002). By adopting this observational strategy a significant fraction of
the cluster luminosity has been sampled with a single HST pointing, In Table 3 (column
2) the fraction of the cluster luminosity sampled (LS/LT ) in each cluster is reported. Here
we present the results obtained using the mid-UV (F255W) and near-UV (F336W) filters,
those best suited for picking out BSS. Table 2 lists the total duration of the exposures in
each filter for the six clusters.
All the photometric reductions (with the exception of NGC 288; see Bellazzini et
al. 2002) have been carried out using ROMAFOT (Buonanno et al. 1983), a package specif-
ically developed to perform accurate photometry in crowded fields. The standard procedure
described in Ferraro et al. (1997a) was adopted. More details on the reduction procedure
and the entire dataset secured in these clusters can be found in some specific papers already
published (see for example Ferraro et al. 1997a for M3 and Ferraro et al. 1999a for M80).
Briefly, the images in the same filter were aligned and combined in order to obtain a median
image in each filter. We used the combined F255W image as reference frame for searching
the objects in each field. Then the PSF fitting procedure was separately performed on each
individual frame, and an average magnitude was computed for each star. The instrumental
magnitudes were finally transformed into the STMAG system, using Table 9 of Holtzmann
et al. (1995).
3. Results
In the UV cluster light is dominated by hot stars, specifically the blue HB stars and the
BSS. (see for example Figure 1 by Ferraro, Paltrinieri & Cacciari 1999). Since the BSS are
one of the hottest sub-populations in the cluster, they are easily separated from the cooler
Turn-Off and SGB stars. Indeed, in previous papers (Ferraro et al. 1997a, 1999a, 2001 and
Paltrinieri et al. 1998, Ferraro, Paltrinieri & Cacciari, 1999) we have shown that UV-CMDs,
in particular the (m255, m255 − m336) plane, are ideal for selecting BSS.
-- 4 --
Figure 1 shows the (m255, m255 − m336) CMDs for the six clusters. More than 50,000
stars are plotted in the six panels of Figure 1. As can be seen, in this plane, the BSS
populations define a clean and well-defined sequence spanning ∼ 3 mag in m255. Further,
they are clearly distinguishable from SGB-TO stars. In order to select the BSS samples we
followed the same criteria that we adopted for M3.
In Ferraro et al. (1997a), the global
population of BSS was divided in two sub-samples: (a) bright BSS with m255 < 19 and
(b) faint BSS with 19.0 < m255 < 19.4. To apply the same criteria to the other clusters
shown in Figure 1, the CMD of each cluster has been shifted to match that of M3 by using
the brightest portion of the HB as the normalization region. Figure 1 shows the UV-CMD
after the alignment. Thus, the solid horizontal line (at m255 = 19) in the figure shows the
cut-off magnitude for the bBSS sample. In this paper we are comparing the BSS content in
clusters with quite different structural parameters. The level of difficulty in the detecting
and the measuring of (especially faint) BSS varies from cluster to cluster. Thus, we decided
to compare the samples by using only the bright BSS samples (hereafter bBSS). The number
of bright BSS (NbBSS) identified in each cluster is listed in Table 3.
3.1. bBSS radial distribution
As pointed out by many authors most BSS in GGCs are found to be centrally concen-
trated with respect to normal stars. Since the central relaxation time in these systems is
much less than the cluster age, this result is generally ascribed to dynamical mass segrega-
tion and can be interpreted as an evidence that BSS are more massive than the comparison
stars.
We have determined the radial distributions of the bBSS of the six GGCs using the
RGB stars (from the tip down to the SGB ) as a "reference" population. The RGB stars
have been selected in the (m555, m336 − m555) CMD, i.e. roughly a (V, U − V ) CMD in order
to reduce any bias which might be introduced by the poor photometry for the very red stars
in the UV bands. The cumulative radial distributions for the bBSS and the RGB stars are
plotted in Figure 2 as a function of the projected distance (r) from the cluster center. It is
evident from the plots that the BSS (solid line) are significantly more centrally concentrated
than RGB stars (dotted line) in M3, M92 and M80. A Kolmogorov-Smirnov test has been
applied to the two distributions, in each cluster, to check the statistical significance of the
detected differences. The test yields the probability that the bBSS population and the RGB
stars are extracted from the same parent distribution. The probability values obtained in
each cluster are reported in Figure 2. The effect is strongly statistically significant in the
case of M3, M92 and M80. For NGC 288 and M10 the BSS are more concentrated that the
-- 5 --
RGB but the results are significant at only roughly the 2σ and 1σ levels respectively.
At first glance the radial distribution of the BSS in M13 appears to be indistinguishable
from the other cluster stars. Since it is difficult to imagine any BSS formation scenario
in which the BSS are not at least somewhat centrally condensed, we have more carefully
examined the role of the small M13 bBSS sample. A series of Monte Carlo simulations have
been performed in which 103 subsamples of the same number of the bBSS population found
in M13 (16) have been randomly extracted from the M3 bBSS population (72). Then, the
radial distribution of each subsample was compared with that of the RGB stars of M3. A
KS test has been applied to quantify the significance of the differences. In these simulations
drawn from a sample in which the central concentration of the BSS has been convinvingly
shown, 40% of the M13-like samples have p > 20% that the BSS and RGB have the same
distribution. Less than 10% have as a convincing a case for different distributions as in
M3. About 6% of the M13-like samples have p as large as for the true M13 comparison.
In another test, two M13 BSS were arbitrary moved from the middle of the distribution to
the inner part and p dropped from 0.77 to 0.34. We conclude that the M13 bBSS sample is
too small to make a definitive statement about its central concentration. Our M13 results
are consistent with no central concentration, but they are not inconsistent with a central
concentration even as large as that of M3.
The case of NGC 288 deserves a short comment since the location of the center in this
cluster is quite uncertain. The radial distribution plotted above could be affected by large
uncertanties. In accordance with Bellazzini et al. (2002) we adopted the coordinates listed by
Webbink (1985), and the cluster center turns out to be located just outside the field-of-view
of the PC (see their Figure 1).
1/2 ). Values for rBSS
We can make an additional quantitative comparison by using the parameters rBSS
1/2 (the
radius containing half the bBSS sample) and fRGB(the fraction of the total RGB sample
1/2 and fRGB are listed in Table 3 along with the core
contained within rBSS
radius (rc). With the exception of NGC 288 (for which we took the value from Trager,
Djorgovski & King, 1993), the rc was independently determined in each cluster by fitting a
King model to the observed density profile (see for example Figure 4 in Ferraro et al. 1999a).
These values confirm the general impression from Figure 2,that the BSS are much more
concentrated than the RGB in M80, M3 and M92. Thus, even considering the brightest part
of the BSS distribution in M80, we have fully confirmed the previous results by Ferraro et
al. 1999a, who suggested that the BSS in M80 are the most concentrated population observed
in a GGCs.
In Figure 3 the magnitude distributions (equivalent to a luminosity function -- LF) of
bBSS for the six clusters are compared. In doing this we use the parameter δm255 defined
-- 6 --
as the magnitude of each bBSS (after the alignment showed in Figure 1) with respect to the
magnitude threshold (assumed at m255 = 19 - see Figure 1). Then δm255 = mbBSS
255 − 19.0.
From the comparison shown in Figure 3 (panel(a)) the bBSS magnitude distributions for M3
and M92 appear to be quite similar and both significantly different from those obtained in
the other clusters. This is essentially because in both these two clusters the bBSS magnitude
distribution seems to have a tail extending to brighter magnitudes (the bBSS magnitude tip
reaches δm255 ∼ −2.5). A KS test applied to these two distributions yields a probability of
93% that they are extracted from the same distribution. In panel(b) we see that the bBSS
magnitude distribution of M13, M10 and M80 are essentially indistinguishable from each
other and significantly different from M3 and M92. A KS test applied to the three LFs
confirms that they are extracted from the same parent distribution. Moreover, a KS test
applied to the total LFs obtained by combining the data for the two groups: M3 and M92
(group(a)), and M13, M80 and M10 (group(b)) shows that the the bBSS-LFs of group(a)
and group(b) are not compatible (at 3σ level).
It is interesting to note that the clusters grouped on the basis of bBSS-LFs have some
similarities in their HB morphology. The three clusters of group(b) have an extended HB
blue tail; the two clusters of group(a) have no HB extention. Could there be a connection
between the bBSS photometric properties and the HB morphology? This possibility needs
to be further investigated.
Finally we note that the magnitude function for the bBSS in NGC 288 (solid line in
panel (b)) is significantly different than that of all the other clusters.
3.2. Specific Frequency
In order to properly compare the BSS populations in different clusters, the BSS number
must be normalized to account for the size of the total cluster population. For this reason
we have defined various specific frequencies.
In Ferraro, Fusi Pecci & Bellazzini (1995)
we used S4BSS (the BSS number normalized to the integrated bolometric luminosity of the
surveyed region in units of 104L⊙). Analogously, in Ferraro et al. (1999a) we defined a more
appropriate specific frequency:
F BSS
HB =
NBSS
NHB
where NBSS is the number of BSS and NHB is the number of HB stars in the same area. The
two ratios are similar, since in absence of any special segregation of HB stars with respect
-- 7 --
to the "normal" cluster-star, the observed number of HB stars NHB is an excellent indicator
of the sampled cluster luminosity. In fact, the simple relation by Renzini & Buzzoni (1986):
Nj = B(t)LStj
easily links the number of stars sampled in any post main sequence stage (Nj) with the sam-
pled cluster luminosity (LS) and the duration of the phase (tj). B(t) is the specific flux of the
population. Following Renzini & Buzzoni (1986) we assume B(t) ∼ 2 × 10−11 stars L−1
⊙ yr−1
for an old population of 1010 yr.
Thus, assuming tHB ∼ 108 yr, we can expect to sample 2 × 10−3 HB stars for each solar
luminosity sampled in a given stellar population.
The F BSS
HB ratio has the advantage that it is a purely observative quantity and it can be
easily computed in the UV-CMDs since the HB population is quite bright in these planes and
the HB sequence is well separated from the other branches. However, while in the following
discussion we will mainly use the F BSS
HB ratio, the S4BSS is also computed for each cluster and
the values are listed in Table 3 for sake of comparison.
HB
HB
are listed in Table 3. As can be seen F bBSS
As discussed in Section 3 we consider only the bright BSS, hence NbBSS is the number
of bright bBSS (according to the original definition by Ferraro et al. (1997a)). The values
of F bBSS
varies by a factor 13 from M13 to NGC
288. Interesting enough NGC 288, the cluster with the lowest central density, shows the
largest bBSS specific frequency. As already noted by Bellazzini et al. (2002) this cluster
turns out to have as many bBSS as HB stars over the sampled area. Though the number
of bBSS detected in the cluster is intrinsically small (∼ 30) the result is surprising: F bBSS
is comparable with what found in the central region of the core-collapsing cluster M80 by
Ferraro et al. (1999a). If we consider only the PC field of view, the F bBSS
in M80 rises up to
0.7.
HB
HB
The central density of a cluster is not the only factor which determines the number
of collision products we expect to see. Collisions involving binary star systems are more
likely than collisions between single stars and have a significant probability of producing a
stellar merger. We used equation 14 from Leonard (1989) in order to estimate the number
of binary-binary (bb) encounters occuring per Gyr (Nbb) in the core of each cluster listed in
Table 1. Adopting the distance modulus listed in Table 1, the rc listed in Table 3, σ0 and the
central density Log(ρ0) listed in Table 1 (from Pryor & Meylan (1993)), assuming an average
mass of 0.2M⊙ (Kroupa 2001), we computed tbb (the mean interval between collisions) from
equation (14) by Leonard (1989) and derived the expected number of encounters, per Gyr,
as a function of the binary fraction (fb) in the core.
-- 8 --
In doing this, the semimajor axis which separates hard from soft binaries (ahs) has been
computed for each cluster according to the following relation:
ahs =
GM
9σ2
0
derived by combining equation 4 of Leonard & Linnell (1992) with the Kepler's third law.
The semimajor axes computed for each cluster and expressed in AU are listed in Table 4.
Following Leonard (1989) we multiplied tbb by a factor of two in order to take into
account the fact that not all encounters lead to a physical collision. The rate of single-binary
collisions can also be estimated from Leonard's results.
In his equation 8 if one assumes
that the pericenter distance is proportional to binary semimajor axis, like in binary-binary
collisions, and that M1 = M⋆, M2(binary) = 1.5 M⋆, one finds a factor of 5 instead of his
6 in the cross section. Thus the number of single-binary encounters is roughly the 5/6 the
number of binary-binary encounters.
In Table 4 we list the results with three different fb values, respectively 0.2 and 1.00.
These values should be considered as indications of the effectiveness of physical collisions in
the cluster cores. It is clear that the bb (or sb) collision channel is more than one order of
magnitude more efficient in a dense cluster like M80 than in a low density cluster like NGC
288. Therefore, the high specific frequency of blue stragglers in NGC 288 suggests that the
binary fraction in NGC 288 is much higher than the binary fraction in M80. Only then would
one expect a similar encounter frequency in the two clusters. Note that a cluster like M80
may have originally had a higher binary fraction but because of the efficiency of encounters,
those primordial binaries were "used up" early in the history of the cluster, producing some
collisional BSS which have evolved away from the MS (see for example the evolved E-BSS
population found in M3 and M80 by Ferraro 1997a, 1999a).
We should note that on the basis of the simulation results shown in Figure 8 and in
Table 4 the binary-binary channel can still account for the number of the BSS observed in
NGC288 if a large enough percentage of binary fraction is assumed to reside in the core.
However, without invoking ad hoc binary content, a more natural explanation for
the origin of BSS in NGC 288 (as discussed in Bellazzini et al. (2002)) is the mass transfer
process in primordial binary systems (Carney et al 2001). Here we probably have another
confirmation of the scenario suggested by Fusi Pecci et al. (1992, and references therein):
BSS living in different environments have different origins.
-- 9 --
4. Collisional Models
From our sample of six clusters with rather diverse bBSS populations we can select pairs
with similarities in parameters like density, metallicity, velocity dispersion, or we can search
for trends in bBSS populations as some parameter varies. There are two quite striking and
unsuspected results. First, the two clusters at the high and low density extremes, M80 and
NGC 288, have the largest bBSS specific frequencies. Second, two clusters which are in most
ways quite similar, M3 and M13, have very different bBSS populations.
To aid our understanding of the BSS populations in these clusters, we present a compar-
ison of the photometric characteristics of the BSS observed in the selected clusters with some
collisional models. The models we used in this paper are described in detail in Sills & Bailyn
(1999), and have been applied to 47 Tucanae (Sills et al. 2000). We assume that the blue
stragglers in the central regions of the six clusters are all formed through stellar collisions
between single stars during an encounter between a single star and a binary system. While
binary-binary collisions may well be important, we do not currently have the capability of
modeling the BSS they might produce. The trajectories of the stars during the collision are
modeled using the STARLAB software package (McMillan & Hut 1996). The masses of the
stars involved are chosen randomly from a mass function for the current cluster and a differ-
ent mass function which governs the mass distribution within the binary system. A binary
fraction, and a distribution of semi-major axes must also be assumed. The output of these
simulations is the probability that a collision between stars of specific masses will occur.
We have chosen standard values for the mass functions (dN(M) ∝ m−(1+x)dM) and binary
distribution. The current mass function has an index x = −2, and the mass distribution
within the binary systems are drawn from a Salpeter mass function (x = 1.35). We chose a
binary fraction of 20% and a binary period distribution which is flat in log P . The total stel-
lar density was taken from the central density of each cluster. The effect of changing these
values is explored in Sills & Bailyn (1999). The collision products are modeled by entropy
ordering of gas from colliding stars (Sills & Lombardi 1997) and evolved from these initial
conditions using the Yale stellar evolution code YREC (Guenther et al. 1992). The models
reported here used a metallicity of [Fe/H] = −1.6, which is approximately correct for 4 of
the clusters investigated in this paper. BSS populations should not be strongly dependent
on metallicity. By weighting the resulting evolutionary tracks by the probability that the
specific collision will occur, we obtain a predicted distribution of BSS in the color-magnitude
diagram.
In order to explore the effects of non-constant BSS formation rates, we considered a
series of truncated rates.
In these models we assumed that the BSS formation rate was
constant for some portion of the cluster lifetime, and zero otherwise. This assumption is
-- 10 --
obviously unphysical -- the relevant encounter rates would presumably change smoothly on
timescales comparable to the relaxation time. However these models do demonstrate how the
distribution of BSS in the color-magnitude diagram depends on when the BSS were created,
and thus provide a basis for understanding more complicated and realistic formation rates.
4.1. Comparisons with data
Some comparisons between the models and the HST observations are shown in figures
4 -- 8. In figures 4, 5 and 6, we compare our theoretical predictions to the BSS population
of M80. The theoretical predictions have been normalized by the total number of predicted
blue stragglers for each cluster; this allows us to more easily compare the shapes of the distri-
butions. Figure 4 is an example of the predicted distribution of BSS in the m255, m255 −m336
color-magnitude diagram for M80. The theoretical predictions are shown as greyscale con-
tours, with the darker colors indicating more BSS in that region of the CMD. The observed
blue stragglers are shown as crosses. Each of the panels is a different assumption for the BSS
formation rate, as indicated on the top of the panel. The differences between the models can
be understood in terms of lifetimes of the individual collision products which make up the
distributions. For example, if BSS production stopped 5 Gyr ago (center panel), we predict
that there should be no observed bBSS at present because all the massive BSS would have
had time to evolve off to the giant branch. At the other extreme, if all the BSS were formed
in the last Gyr (panel marked "1 Gyr to now"), only a few of the very brightest (i.e. most
massive) would have even started moving towards the subgiant branch. It is worth pointing
out that the brightest BSS shown in these predictions are not on the standard zero age
main sequence, but begin their lives to the red. These are stars which were formed from the
merger of two fairly evolved main sequence stars, and so they do not have enough hydrogen
fuel left at their cores after the merger to return to the ZAMS (Sills et al. 1997).
Figure 5 gives the luminosity function for each of the model distributions for M80. The
data are shown as the solid line, while the dotted line is the theoretical prediction for the
formation times shown on the top of each panel. The luminosity functions are good probes
of the mass distribution of blue stragglers, since the correlation between mass and luminosity
is quite tight. The luminosity functions show, as do the CMD distributions, that bright BSS
require recent formation. They also show that the relative numbers of brighter and fainter
bBSS is not correct in our models. If we model the fainter bBSS very well (as in the panel
marked "stopped 2 Gyr ago"), then our models under-predict the number of blue stragglers
between m255 = 17 and 18.
Figure 6 gives the equivalent "temperature function" for M80. This is the cumulative
-- 11 --
number of stars in each color (i.e. temperature) bin, in direct analogy to the luminosity
function. The distribution of blue stragglers in color can be used to study the age or age
spread of the BSS. Evolutionary tracks for stars of these masses are confined to a fairly
narrow range in luminosity between the main sequence and the base of the giant branch,
so the luminosity function does not give the entire picture. The temperature function is
useful for distinguishing between populations which have lots of main sequence stars and
populations with many subgiants. As in figure 5, the data are shown as the solid line, while
the dotted line is the theoretical prediction for the formation times shown on the top of each
box.
We performed KS tests on the luminosity and temperature functions for each cluster
and each theoretical distribution. Figures 7 and 8 gives the best fit model distribution for
each cluster, its luminosity function, and its temperature function. We will comment on
each cluster individually here, and then draw some general conclusions in the next section.
M80: The fainter bBSS are very well fit by a distribution in which formation was trun-
cated about 2 Gyr ago. However, there is a significant population of BSS which extends up
to 2.5 magnitudes above the turnoff, which is not predicted by such a model and which sug-
gests that some BSS formation continues until the present. Ferraro et al. (1999a) suggested
that the large BSS population in M80 is due to the stellar interactions which are delaying
the core collapse process. Now we add the evidence that many of these BSS are relatively old
(t > 2 Gyr). These results together suggest that the process of holding off the core collapse
can be rather extended.
M10: The BSS in this cluster are fit by different formation models in the temperature
and luminosity planes. However, the difference between the KS probability for the best fit
to luminosity function (BSS formation ended 2 Gyr ago) and the best fit for temperature
function (BSS formation ended 1 Gyr ago) is small. The luminosity function has a 16%
probability of being drawn from the first distribution, and a 10% probability of being drawn
from the second distribution.
M13: Of all the clusters in this paper, this one is the most believable for showing a
truncated BSS distribution. There is no evidence for very bright BSSs close to the main
sequence. The only BSS brighter than δm255 = 2.5 is very red, consistent with an evolved
object.
M3: None of the collisional models fit this cluster very well. The best fit for the
luminosity function, as shown in figure 8, is a model in which the BSS stopped forming
about 1 Gyr ago (3% KS probability). However, the temperature function for the same
model has almost zero probability of being drawn from the same distribution as the data,
-- 12 --
and the best fit temperature function is the scenario in which the BSS stopped forming 2
Gyr ago. In all cases, including the models which predict BSS formation right up to the
present day, we under-predict the number of very bright blue stragglers.
M92: The best fit to both the luminosity and temperature functions is a model in
which the BSS formation was truncated 1 Gyr ago. However, there are also very bright
blue stragglers in this cluster, suggesting that blue straggler formation has continued to the
present day or that the blue stragglers produced in this cluster are not well modeled by
collisions between two stars.
NGC 288: This cluster is significantly different from the previous 5, in that we do not
expect that the BSS formed in this cluster are the products of stellar collisions. The density
is simply too low. We also know that the binary fraction in the core is measured to be
∼ 10 to 38% (Bellazzini et al. 2002), so BSS formed in this low density cluster should be
the result of binary evolution rather than stellar collisions. By assuming that binary merger
BSS are similar to collisional BSS we can still expore the BSS formation history in NGC 288.
The best fit distribution for this cluster, at the 65% level, is achieved when the BSS were
formed at a constant rate over the entire lifetime of the cluster, right up to the present day.
This result is significantly different from the other five clusters investigated in this paper.
Perhaps, it could be an indication of the different formation histories of binary mergers and
collisions. It is also possible that the difference between this cluster and the other 5 could
be a result of different evolutionary tracks for binary merger and collision products.
4.2.
Indications and Trends
The first conclusion that we draw from the models is that since we are limiting our
study to BSS brighter than the limit of m255 = 19.0, we have no information about BSS
formation earlier than about 5 Gyr ago. On the other hand, we can state with certainty
that BSS have been formed in all these clusters during the last 5 Gyr. In all clusters, this
formation is not confined to the very recent past; rather, it has lasted over the entire time
that we can probe, and in fact seems to be concentrated at earlier epochs.
The KS tests on the luminosity and temperature functions give the best results for
distributions in which BSS stopped forming 1 -- 2 Gyr ago in all clusters except NGC 288.
These "best" fits are clearly not good fits, however.
In most of the clusters, there is a
population of bright BSS that was formed more recently. Our models consistently under-
predict the number of bright BSS compared to the number of fainter, redder BSS. There are
a number of possible reasons.
-- 13 --
The first, and most likely, discrepancy is in the formation rate of blue stragglers. We
have assumed a constant formation rate over the some times, and zero otherwise. If, instead,
blue stragglers were formed at some (more or less) constant rate between 5 and 2 Gyr ago,
and then experienced a reduced (but not zero) blue straggler formation rate to the present
day, the predicted distribution would match the observations better. It is very unlikely that
the blue straggler formation rate changes sharply between zero and a constant value. Rather,
the formation rate is going to be smoother function of central density, velocity dispersion,
single star mass function and the number and nature of binary systems in the cluster. The
simple models used in this paper should be taken as indications of formation eras only.
It is also possible that the stellar evolutionary models which go into these blue straggler
distributions are not accurate.
If we were overestimating the lifetimes of the faint blue
stragglers or underestimating the lifetimes of the bright blue stragglers, we would see this
kind of mismatch between observations and data.
It is difficult to understand how the
evolutionary tracks would be overestimating the lifetimes of the faint blue stragglers. The
lower the mass of the collision product, in general, the more similar it is to a normal star of
the same mass, and so its evolutionary track is very much like normal stars -- something we
understand quite well. In order to underestimate the lifetimes of the brighter blue stragglers,
the collision products would have to mix more hydrogen into their cores than is currently
predicted. Rapid rotation is a plausible mechanism for this mixing, although we would have
to have some process by which the lower mass BSS were not rotating as rapidly. It is also
possible that the bright blue stragglers could actually be the merger of three stars rather
than two (as is expected to happen in binary-mediated collisions), and so our evolutionary
models will not be applicable. The merger product of three 0.4 M⊙ stars will have a higher
percentage of hydrogen in its core than the product of two 0.6 M⊙ stars, but approximately
the same mass.
The difference between M3 and M92 on one hand, and M10, M13 and M80 on the other
is primarily the existence of a few very bright blue stragglers in the first group of clusters. In
all 5 clusters, the blue stragglers fainter than about m255 = 18.5 are very well modeled by the
same distribution (a truncated formation scenario, suggesting that blue straggler formation
rates should be weighted towards 3 -- 5 Gyr ago). The fact that M3 and M92 are better fit
by a formation scenario in which BSS formation ended more recently is simply a reflection
of the number of very bright BSS in these clusters. These bright blue stragglers could be an
indication of continued BSS formation, or of a different binary distribution (which produced
more triple collisions). M3 does have a significant population of BSS in the outer part of
the cluster which are thought to be the result of binary mergers (Ferraro et al. 1997a). This
fact does set M3 apart from M13, M10, & M80, which have few exterior BSS. Unfortunately
M92 also has few exterior BSS so they cannot be invoked to explain the M3/M92 similarity.
-- 14 --
As noted in Section 3.1, another very interesting point is that M10, M13 and M80
(which share a common BSS Luminosity function) all have blue tail components to their
horizontal branches, while M3, M92 and NGC 288 do not. M3 and M13 are also a classic
second-parameter pair (see Ferraro et al. 1997c), referring to the difference in their horizontal
branch morphology without an obvious difference in any other cluster parameter: they have
the same metallicity, central density, total mass, and very similar velocity dispersions. (The
two clusters may have slightly different ages, which may or may not affect horizontal branch
morphology -- see Rey et al. 2001, Davidge & Courteau 1999, Ferraro et al. 1997c). From
these data, there is certainly a suggestion that BSS population could be connected to the
second parameter. From the comparisons in this paper, it seems that it is not the total
number of BSS or their specific frequency, but rather the kind of BSS which are produced in
"blue tail" clusters which is different from those produced in non-blue tail clusters. If this
link between BSS and horizontal branch morphology is real, then it suggests that the second
parameter could be related to either stellar collisions and encounters, or to binary evolution.
5. Conclusions
In this paper, we have compared the blue straggler distribution in six galactic globular
clusters: M3, M10, M13, M80, M92 and NGC 288. The blue stragglers were observed in
the HST UV filters F255W and F336W. The clusters have similar total masses, and veloc-
ity dispersions; have intermediate to low metallicities and span a range in central density.
We studied the properties of the blue stragglers relative to other cluster populations; and
we compared the observed blue straggler distributions with theoretical models of collision
products in globular clusters.
Whether they formed via a collisional channel or a merged primoridial binary channel,
BSS are the most massive stars in GGCs other than neutron stars or possibly some white
dwarfs. Since dynamical relaxation times are typically shorter than BSS lifetimes they should
settle toward the cluster center. BSS formed from merged primordial binaries have the entire
cluster lifetime to settle. Collisions occur most often in high density regions and most often
involve binaries providing yet another reason to expect BSS to be found near cluster centers.
Indeed, it is difficult to image a BSS formation scenario which would not lead to a centrally
condensed BSS distribution. Our results are consistent with this expectation: M3, M92,
& M80 show strong central concentrations; the BSS in M10 and NGC 288 are most likely
centrally condensed; while the BSS in M13 could have the same distribution as the RGB,
they could also have a central concentration similar to that of M3. The small sample size is
not adequate to definitively determine the radial distribution.
-- 15 --
The specific frequency of blue stragglers compared to the number of horizontal branch
stars varies from 0.07 to 0.92 for these six clusters, and does not seem to be correlated with
central density, total mass, velocity dispersion, or any other obvious cluster property. We do
not have measurements of the binary fraction in most of these clusters. Binary stars impact
the blue straggler populations in a number of ways. BSS can be formed from binary evolution
in any environment, although binary systems in dense environments will be different from
those in sparse environments -- close binaries will be hardened by encounters (increasing the
number of "binary merger" BSS), but wide binaries are also disrupted (decreasing the number
of "binary merger" BSS). The relative rates of these events is an important factor in looking
at blue stragglers in clusters. Binary systems also facilitate close interactions between stars
because they have a large cross section compared to single stars, so we expect that clusters
with a large binary fraction to have both more collisional and binary merger blue stragglers.
Fusi Pecci et al. (1992, and references therein) presented a qualitative interpretative
scenario concerning the possible origin of BSS in GGCs. They suggest that BSS in loose
clusters might be produced from coalescence of primordial binaries. In high density GGCs
(depending on survival-destruction rate of primordial binaries) BSS might mostly arise from
stellar interactions, particularly those which involve binaries (Ferraro, Fusi Pecci & Bellazz-
ini, 1995). In this scenario one could find clusters where more than one mechanism is at work
in generating BSS. In fact Ferraro et al. (1993, 1997a) found that the radial distribution and
the luminosity functions of BSS in the center of M3 are consistent with a collisional origin,
while in the outer regions they are consistent with a primordial binary origin. The paucity
of BSS in M13 suggests that either the primordial population of binaries in M13 was poor or
that most of them were destroyed. Alternatively, as suggested by Ferraro et al. (1997b), the
mechanism producing BSS in the central region of M3 is more efficient than M13 because
M3 and M13 are in different dynamical evolutionary phases.
The evidence shown above suggests that different channels of BSS formation should
be at work and/or a large difference in the binary fraction should be present in NGC 288
in order to produce such a large BSS frequency in a cluster with such different structural
parameters. Bellazzini et al. (2002) measured the binary fraction in the core of NGC 288 to
be between 10% and 38%. Similar measurements for M3, M13, and M80 will be extremely
helpful in solving this mystery.
According to simple models, BSS formation occured in all clusters over last 5 Gyr at
least, and more were formed 2 -- 5 Gyr ago than in the recent past. This result needs to
be tested using models of globular cluster evolution in which the feedback between stellar
collisions and cluster evolution is modeled explicitly. Our assumption of a BSS formation
rate which is either constant or zero is unphysical, and more complicated models are clearly
-- 16 --
required.
Finally, we note the possible connection between horizontal branch morphology and blue
straggler population characteristics. Of our six clusters, three have HB blue tails (M10, M13,
M80), and three do not. The three with blue tails have the same blue straggler luminosity
function, which spans about 1.5 magnitudes compared to 2.5 magnitudes in M3 and M92.
This suggests that blue straggler populations and the horizontal branch second parameter
may be linked in some way, perhaps through the complicated world of binary systems and
their effect on cluster evolution and populations.
We warmly thank Michele Bellazzini for useful discussions and the referee, Peter Leonard,
for the many helpful suggestions which significantly improved the presentation of the pa-
per. The financial support of the Agenzia Spaziale Italiana (ASI) and of the Ministero
dell'Istruzione, dell'Universit`a e della Ricerca (MIUR) is kindly acknowledged. RTR is par-
tially supported by grant GO-8709 from STScI.
Bailyn, C.D. 1995, ARA&A, 33, 133
REFERENCES
Bellazzini, M., Fusi Pecci, F., Messineo, M., Monaco, L., Rood, R.T. 2002, AJ, 123, 1509
Buonanno, R., Buscema, G., Corsi, C.E., Ferraro, I., & Iannicola, G. 1983, A&A, 126, 278
Carney, B.W., Latham, D.W., Laird, J.B., Grant, C.E., & Morse, J.A. 2001, AJ, 122, 3419
Castellani, M., & Castellani, V.,1993, ApJ 407, 649
Cool, A.M., Grindlay, J.E., Cohn, H.N., Lugger, P.J., & Slavin, S.D. 1995, ApJ, 439, 695
Cool, A.M., Grindlay, J.E., Cohn, H.N., Lugger, P.J., & Bailyn, C.D. 1998, ApJ, 492, L75
De Marchi, G., Paresce, F., Ferraro, F.R. 1993, ApJSS, 85, 293 (DPF93)
Djorgovski, S., Meylan, G., 1993, in Structure and Dynamics of Globular Clusters, ed. S. G.
Djorgovski & G. Meylan (ASP: San Francisco), 325
Ferraro, F.R., Fusi Pecci, F. & Bellazzini, M., 1995, A&A, 294, 80.
Ferraro, F.R.,Paltrinieri, B., Fusi Pecci, F., Cacciari, C., Dorman, B., Rood, R.T., Buonanno,
R., Corsi, C.E., Burgarella, D., Laget, M. 1997a, A&A, 324, 915
-- 17 --
Ferraro, F.R., Paltrinieri, B., Fusi Pecci, F., Dorman, B., & Rood, R.T. 1997b, MNRAS,
292, L45
Ferraro, F.R., Paltrinieri, B., Fusi Pecci, F., Cacciari, C., Dorman, B., Rood, R. T. 1997c,
ApJ, 484, L145.
Ferraro, F.R., Paltrinieri, B., Fusi Pecci, F., Rood, R.T., & Dorman, B. 1998, in Ultra-
violet Astrophysics -- Beyond the IUE Final Archive, eds. R. Gonz´alez-Riestra, W.
Wamsteker, & R. A. Harris (ESA: Noordwijk), 561
Ferraro, F.R., Paltrinieri, B., Rood, R.T., Dorman, B., 1999a, ApJ, 522, 983
Ferraro, F.R., Messineo, M., Fusi Pecci, F., De Palo, A., Straniero, O., Chieffi, A., Limongi,
M. 1999b, AJ, 118, 1738
Ferraro, F.R., Paltrinieri, B., Cacciari, C. 1999, Mem SAIt, 70, 599
Ferraro, F.R., Paltrinieri, B., Rood, R.T., Fusi Pecci, F., Buonanno, R. 2000a, ApJ, 537,
312
Ferraro, F.R., Paltrinieri, B., Paresce, F., De Marchi, G. 2000b, ApJ 542, L29
Ferraro, F.R., D'Amico, N., Possenti, A., Mignani, R.P., Paltrinieri, B. 2001, ApJ, 561, 337
Fusi Pecci, F., Ferraro, F.R., Corsi, C.E., Cacciari, C., & Buonanno, R. 1992, AJ, 104, 1831
Holtzmann, J.A., Burrows, C.J., Casertano, S., Hester, J.J., Trauger, J.T., Watson, A.M.,
& Worthey, G. 1995, PASP, 107, 1065
Guenther, D. B., Demarque, P., Kim, Y.-C., & Pinsonneault, M. H. 1992 ApJ, 387, 372
Kroupa, P. 2001, MNRAS, 322, 231
Leonard, P.J.T. 1989, AJ, 98, 217
Leonard, P.J.T., Linnell, A.P. 1992, AJ, 103, 1928
McMillan, S. L. W., & Hut,P. 1996, ApJ, 467, 348
Paltrinieri, B., Ferraro, F.R., Fusi Pecci, F., Rood, R.T., Dorman, B. 1998,
in Ultra-
violet Astrophysics -- Beyond the IUE Final Archive, eds. R. Gonz´alez-Riestra, W.
Wamsteker, & R. A. Harris (ESA: Noordwijk), 565
Paresce, F., et al. 1991, Nature, 352, 297 (P91)
-- 18 --
Pryor, C., & Meylan, G. 1993, in ASP Conf. Ser. 50, Structure and Dynamics of Globular
Clusters, ed., S.G. Djorgovski & G. Meylan, (San Francisco: ASP), 357
Renzini, A., Buzzoni, A. 1986 in Spectral Evolution of Galaxies, C. Chiosi& A. Renzini (eds),
Dordrecht: Reidel, p.135.
Sills, A. & Lombardi, J. C., Jr. 1997, ApJ, 489, L51
Sills, A., & Bailyn, C. D. 1999, ApJ, 513, 428
Sills, A., Bailyn, C. D., Edmonds, P. D., Gilliland, R. L. 2000, ApJ, 535, 298
Trager, S.C., Djorgovski, S.G., & King, I.R. 1993, in ASP Conf. Ser. 50, Structure and
Dynamics of Globular Clusters, ed., S.G. Djorgovski & G. Meylan, (San Francisco:
ASP), 347
Webbink, R. 1985, in IAU Symp. 113, Dynamics of Star Clusters, ed. J. Goodman & P. Hut
(Dordrecht:Reidel), 541
This preprint was prepared with the AAS LATEX macros v5.0.
-- 19 --
Fig. 1. -- (m255, m255 − m336) CMDs for the selected clusters. Horizontal and vertical shifts
have been applied to all CMDs in order to match the main sequences of M3. The horizontal
solid line corresponds to m255 = 19 in M3. The bright BSS candidates are marked as large
empty circles.
-- 20 --
Fig. 2. -- Cumulative radial distributions for the bright BSS (solid line) with respect to the
RGB stars (dashed line) as a function of their projected distance (r) from the cluster center
for each of the six clusters. The probability that the two populations are extracted from the
same distribution is also reported in each panel.
-- 21 --
Fig. 3. -- Cumulative magnitude distributions for the bright BSS for each of the six clusters.
The δm255 parameter is the difference in magnitude with respect to the limit threashold
(m255 = 19) for bright-BSS, see text. In Panel (a) the BSS distributions for M3 and M92
(the two clusters for which the BSS distribution extends up to more than two magnitudes
brighter than the threashold) are compared. In Panel (b) the BSS magnitude distributions
for the other 4 clusters are plotted.
-- 22 --
Fig. 4. -- The distribution of bright blue stragglers in the color-magnitude diagram for
M80, compared to theoretical distributions. The observations are plotted as crosses, and
the greyscale contours give the theoretical distributions, with darker colors indicating more
blue stragglers. The different panels correspond to different eras of constant blue straggler
formation, as indicated at the top of each panel.
-- 23 --
Fig. 5. -- Luminosity functions for the bright blue stragglers in M80 (solid line), compared
to theoretical predictions (dotted line). The different panels correspond to different eras of
constant blue straggler formation, as indicated at the top of each panel. The luminosity
functions are good probes of the mass function of blue stragglers. This figure shows that the
theoretical models do not accurately model the relative numbers of fainter and brighter blue
stragglers.
-- 24 --
Fig. 6. -- Temperature functions for the bright blue stragglers in M80 (solid line), compared
to theoretical predictions (dotted line). The different panels correspond to different eras of
constant blue straggler formation, as indicated at the top of each panel. The temperature
functions are good probes of the bulk evolutionary state of the population of blue straggler
stars, indicating whether the stars are all on the main sequence (as in the panel marked "1
Gyr to now") or closer to the subgiant & giant branches (as in the panel marked "stopped
2 Gyr ago"). The BSS in M80 show a significant spread in evolutionary state.
-- 25 --
Fig. 7. -- The observed blue stragglers (crosses in the CMD, solid lines in the luminosity and
temperature functions) and the best fit theoretical distribution, as determined from KS tests
on the luminosity and temperature functions. This figure shows the results for M80, M10
and M13. In all cases, the nominal best-fit distribution is one in which the blue stragglers
stopped forming 2 Gyr ago. However, only for M13 is this actually plausible. M80, and
to a lesser extent M10 were forming blue stragglers more recently than 2 Gyr ago, since
we see a population of BSS brighter than m255 = 18. On the other hand, the temperature
function (i.e. the distribution of BSS in evolutionary state) is very well fit by this truncated
-- 26 --
Fig. 8. -- Same as figure 7, for M3, M92 and NGC 288. M3 and M92 contain the brightest
blue stragglers in our sample, and so the best-fit theoretical distribution is one in which young
blue stragglers (up to 1 Gyr old) are found. However, because of the spread in temperature,
these clusters are best fit by a truncated blue straggler formation model, like M80, M10 and
M13. NGC 288 is different from the previous 5 clusters -- its blue straggler distribution is
extremely well fit by a model which has had essentially constant blue straggler formation
over the lifetime of the cluster right to the present day.
-- 27 --
Table 1. Cluster parameters
Cluster
[Fe/H]
Logρ0
Mass
[M⊙ pc−3]
[Log(M/M⊙)]
NGC 5272(M3) −1.66
NGC 6205(M13) −1.65
NGC 6093(M80) −1.64
NGC 6254(M10) −1.60
−1.40
NGC6341 (M92) −2.24
NGC 288
3.5
3.4
5.4
3.8
2.1
4.4
5.8
5.8
6.0
5.4
4.9
5.3
d
[kpc]
10.1
7.7
9.8
4.7
8.8
9.0
σ0
[km s−1]
5.6
7.1
12.4
5.6
2.9
5.9
Table 2. WFPC2 exposure times
Cluster
F336W-Exp [s]
F255W-Exp [s]
M3
M13
M80
M10
NGC 288
M92
3340
560
2400
1500
3760
3800
1200
200
1160
5200
700
2200
Table 3. Bright BSS populations and distributions
Cluster
LS/LT NbBSS
S4BSS
F bBSS
HB
rBSS
1/2
rc
rBSS
1/2 /rc
fRGB
M3
M13
M80
M10
NGC 288
M92
0.25
0.27
0.48
0.30
0.20
0.45
72
16
129
22
24
53
6.0
1.6
13.4
5.6
21.8
4.8
0.28
0.07
0.44
0.27
0.92
0.33
22′′
46′′
7′′
34′′
60′′
15′′
30′′
40′′
6.5′′
40′′
85′′
14′′
0.73
1.15
1.07
0.85
0.71
1.07
0.31
0.47
0.17
0.35
0.38
0.29
-- 28 --
Table 4. Expected number of binary-binary
encounters per Gyr
Cluster
ahs(AU) Nbb(fb = 0.2) Nbb(fb = 1)
M3
M13
M80
M10
NGC 288
M92
0.63
0.39
0.13
0.63
2.36
0.57
12
4
66
12
2
47
304
99
1640
289
52
1178
|
Subsets and Splits
No community queries yet
The top public SQL queries from the community will appear here once available.