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/0210057 | 1 | 0210 | 2002-10-02T08:07:16 | Binary evolution and neutron stars in globular clusters | [
"astro-ph"
] | Improved observations of globular clusters are uncovering a large number of radio pulsars and of X-ray sources. The latter include binaries in which a neutron star or a white dwarf accretes matter from a companion, recycled pulsars, and magnetically active binaries. Most of these sources originate from close encounters between stars in the cluster core; magnetically active binaries and some cataclysmic variables have evolved from primordial binaries. The formation rate through close encounters scales differently with the central density and core radius of the cluster than the probability for a single binary to be perturbed by an encounter. This is exploited in some preliminary observational tests of the close encounter hypothesis.
Accreting black holes have been found in globular clusters with other galaxies; the absence of such holes in the Milky Way clusters is compatible with the small number expected. | astro-ph | astro-ph |
New horizons in globular cluster astronomy
ASP Conference Series, Vol. **, 2002
Eds. G. Piotto, G. Meylan, G. Djorgovski, M. Riello
Binary evolution and neutron stars in globular clusters
Frank Verbunt
Astronomical Institute, Utrecht University, Postbox 80.000, 3508 TA
Utrecht, The Netherlands; email [email protected]
Improved observations of globular clusters are uncovering a
Abstract.
large number of radio pulsars and of X-ray sources. The latter include
binaries in which a neutron star or a white dwarf accretes matter from
a companion, recycled pulsars, and magnetically active binaries. Most of
these sources originate from close encounters between stars in the cluster
core; magnetically active binaries and some cataclysmic variables have
evolved from primordial binaries. The formation rate through close en-
counters scales differently with the central density and core radius of the
cluster than the probability for a single binary to be perturbed by an
encounter. This is exploited in some preliminary observational tests of
the close encounter hypothesis.
Accreting black holes have been found in globular clusters with other
galaxies; the absence of such holes in the Milky Way clusters is compati-
ble with the small number expected.
1.
Introduction
Interest in binaries in globular clusters revived when the first X-ray maps of
>
the whole sky showed an overabundance in globular clusters of bright (Lx ∼
1036 erg/s) sources: whereas globular clusters contain ∼0.1% of the stars of
our galaxy, they contain about 10% of the bright X-ray sources. It was soon
suggested that close stellar encounters in globular cluster cores were responsible
< 1035 erg/s) sources
(Clark 1975). With more sensitive instruments faint (Lx ∼
were detected in globular clusters: 8 in 8 clusters with the Einstein satellite in
the 1980s (Hertz & Grindlay 1983), and 57 in 23 clusters with ROSAT in the
1990s (Verbunt 2001). Of the latter 57, 17 are more than two core radii from
the cluster core. In the new millenium, Chandra detected about 100 faint X-
ray sources in 47 Tuc alone, and 25-40 sources in each of NGC6752, NGC6397,
and NGC6440 (Grindlay et al. 2001a,b, Pooley et al. 2002a,b; Fig. 1). First
XMM results are now being published (Webb et al. 2002; Webb, Gendre in
these proceedings).
Radio pulsars have also been detected in large numbers: 20 in 47 Tuc
alone, 8 in M15, 5 in NGC6752, and some 20 in a dozen other clusters (Freire
et al. 2001, D'Amico et al. 2002, and lists in e.g. Phinney 1992, Lyne 1994, and
on the Web: Freire 2002). Almost all of these are recycled radio pulsars, i.e.
have obtained their rapid rotation (and probably also their low magnetic field)
through accretion of matter and angular momentum from a companion star.
1
2
Verbunt
10 arcsec
12
21
2
4
22
17
23
10
3
18
1
6
11
9
14
15
24
19
716
5
20
13
8
N
E
Figure 1.
X-ray sources
in NGC6440.
Sources 1
certainly and 2, 3 and 5
probably are neutron stars
accreting from a compan-
ion star at a low rate; most
of the other sources are
probably white dwarfs ac-
creting from a companion,
i.e. cataclysmic variables.
Note that the source distri-
bution extends beyond the
core radius (indicated with
a dashed circle), but re-
mains well within the half-
mass radius (solid-line cir-
cle). From Pooley et al.
(2002b).
At the moment 13 bright X-ray sources are known in a total of 12 globular
clusters. The only binaries producing such luminosities are low-mass X-ray
binaries, in which a neutron star or black hole accretes mass from a companion.
Because of the occurrence of X-ray bursts, 12 of the 13 bright sources in clusters
are known to be neutron stars; the 13th one is -- on the basis of its X-ray spectrum
-- probably a neutron star as well (in 't Zand et al. 1999, White & Angelini 2001).
Thus there is no evidence for accreting black holes in any globular cluster with
< 1035 are most likely
our galaxy. The dim sources with 1032.5
∼
< 1032.5) are
neutron stars accreting at a low rate; most of the dimmer sources (
∼
probably cataclysmic variables, but some of them are recycled radio pulsars, or
magnetically active close binaries ('RS CVns') -- see e.g. Fig.8 in Verbunt et al.
(1997); and the articles by Cool and Edmonds in these proceedings.
< Lx(erg/s)
∼
In this review I discuss the formation and evolution of these X-ray sources
and recycled radio pulsars. Formation through evolution from a primordial bi-
nary is compared with formation via close stellar encounters in Section 2. The
evolution of binaries with a compact star (neutron star or white dwarf) is de-
scribed in Section 3. Various tests of our theoretical ideas are made by com-
parison with observation in Section 4. A brief discussion in Section 5 of X-ray
sources in globular clusters of other galaxies, precedes the Summary.
2. Formation: evolution versus encounters
All binaries that we observe in the disk of the galaxy have evolved from pri-
mordial binaries. Binaries which are very common in the disk, apparently are
commonly formed via ordinary binary evolution. If we find such binaries in a
globular cluster, it is likely that they have evolved from primordial binaries as
well. Examples of such binaries are RS CVn's and contact binaries. On the
other hand, binaries with neutron stars (or black holes) are extremely rare in
Neutron stars in globular clusters
3
the disk; such a binary in a globular cluster is formed almost certainly from a
close encounter of a neutron star with a single star or with a binary. Cataclysmic
variables are in between these extremes: in globular clusters with dense cores,
such as 47 Tuc, most may have formed via stellar encounters; but in more open
clusters, as ω Cen, they may have evolved from primordial binaries (Verbunt &
Meylan 1988).
We consider tidal capture first. The encounter rate is proportional to the
encounter cross section A, to the relative velocity v, and to the numbers nc and n
per unit volume, of compact stars and ordinary stars respectively. To obtain the
encounter rate Γ for the cluster as a whole, one integrates over the cluster volume;
because of the high densities in the core this roughly corresponds to multiplying
the central rate with the core volume. Due to gravitational focussing, the cross
section A of close encounters is proportional to the radius of the star R and
inversely proportional to the square of the velocities v (see also Davies, these
proceedings). Thus (Hut & Verbunt 1983):
Γ ∝ Z ncnAvdV ∝ Z ncnR
v
dV ∝
or3
ρ2
c
v
R
(1)
Where ρo is the central mass density and rc the core radius. An analogous
reasoning gives the exchange encounter rate
Γe ∝ Z ncnbAbvdV ∝ Z ncnba
v
dV ∝
ρ2
or3
c
v
a
(2)
where nb is the number of binaries per unit volume, and a the semi-major axis
of the binary. The ratio of tidal capture to exchange encounters is roughly
Γ
Γe ∼
R
a
n
nb
(3)
The susceptibility of large binaries to close encounters with other cluster
stars implies that many such binaries are dissolved by passing stars. As a result,
the formation of cataclysmic variables from evolution of primordial binaries,
which passes through a stage in which the binary is very wide, is suppressed in
globular clusters (Davies 1997).
The importance of tidal capture is under debate, because it leads to an
initially highly eccentric orbit: the circularization of this orbit is accompanied
by dissipation of an amount of energy ∆E which is comparable to the binding
energy Eb of the ordinary star:
∆E
Eb ≃ (cid:18)−GM m
2ac (cid:19) / 3GM 2
5R ! ≃
5
6
m
M
R
ac
(4)
where M and R are the mass and radius of the tidally distorted star, m the
mass of the other star and ac the radius of the circularized orbit. If the energy is
dissipated more rapidly than it can be radiated or convected away, the ordinary
star is destroyed, and no binary remains (Ray et al. 1987).
Tidal capture, if successful, tends to lead to orbits with a ≃ 3R, i.e. to orbits
with short periods, less than 1 day, say. In contrast, exchange encounters favour
4
Verbunt
wide orbits, and in addition will cause a recoil velocity of the binary which may
take it out of the cluster core -- perhaps even out of the cluster! Tidal capture
can also occur during an encounter of a single star with a binary, when the
three-body intraction brings two stars close to one another.
A much debated point of uncertainty is the number of neutron stars that
remain in a cluster after their formation: young pulsars obtain a kick velocity
at birth, as witnessed by the observed velocities of radio pulsars. It has been
argued that these velocities are higher than the escape velocities of globular
clusters, implying that no neutron stars would be retained. In my opinion, the
velocities of pulsars have been rather overestimated, due to underestimates of
the errors in proper motions and distances (see Hartman 1997). A recent list of
16 accurate velocities obtained with VLBI (Brisken et al. 2002) indicates that
as much as a third of neutron stars is born with velocities less than ∼ 50 km/s.
Thus as many as 30% of the neutron stars born in globular clusters are retained.
Loss of neutron stars from globular clusters is counteracted by mass segregation,
which concentrates the (relatively heavy) neutron stars in the cores, where the
encounter rates are highest (Verbunt & Meylan 1988).
The presence of pulsars in globular clusters is interesting, as they constrain
the amount of non-luminous matter (to which they themselves do not contribute
significantly, as the mass in white dwarfs is always much higher). A pulsar which
has a radial velocity vr with respect to us, will be observed at a period shifted by
∆P = (vr/c)P . If the pulsar is accelerated, a period derivative ∆ P = (ar/c)P
is measured on top of the intrinsic period derivative. Since the latter is always
positive, measurement of a negative period derivative indicates that acceleration
dominates, and thus provides an estimate of the mass density. Phinney (1992)
has shown that the pulsars in M15 put a tight constraint on the mass of any
black hole in the cluster core -- the most likely mass of the cluster core can
in fact be explained by the sum of visible stars and white dwarfs, if the mass
distribution initially followed the Salpeter function, i.e. no black hole is required.
3. Binary evolution
The evolution of a binary with a neutron star (or black hole, or white dwarf) is
a complex topic, which is described only briefly here. For more detail, see e.g.
Verbunt (1993). The mass transfer in a close binary in which a companion to
a neutron star fills its Roche lobe, can be driven by loss of angular momentum
from the orbit.
If a compact star of mass m receives matter from a Roche-
M/M ≃
filling star with mass M , the mass transfer rate is roughly given by −
J/Jb, where Jb is the binary angular momentum. For a main sequence
m/m ∼
star donor, the orbital period Pb is roughly given by Pb(hr) ≃ 8M (M⊙), and
gravitational radiation leads to a mass transfer rate of order 10−10M⊙ yr−1.
When accreted onto a neutron star this leads to an X-ray luminosity ∼ 1036
erg s−1. The observation of higher X-ray luminosities in such binaries have lead
to the suggestion of additional mechanisms of loss of angular momentum, such
as 'magnetic braking', in which a stellar wind carries away angular momentum
efficiently because it is tied to the stellar surface out to large radii by magnetic
field lines. The efficiency of magnetic braking is unknown; recent studies of
stellar rotation in young star clusters suggests that its importance may have
Neutron stars in globular clusters
5
been over-estimated (e.g. Tinker et al. 2002). The mass transfer rate may also
be affected if the donor is irradiated by X-rays from the accreting star.
If
this irradiation causes the star to expand, it may enhance the mass transfer
rate; if it causes an enhanced stellar wind and mass loss from the binary, it may
reduce the mass transfer rate. Whether irradition is important is not known; the
absence of strong heating effects in most well-observed low-mass X-ray binaries
suggests that the accretion disk shield the secondary from X-rays emitted near
the accreting star.
If the mass transfer stops, the neutron star may spin rapidly enough to
switch on as a radio pulsar. The standard description of evolution through loss
of angular momentum does not foresee such an end however, but rather predicts
an everlasting, albeit ever lower, mass transfer. The observation of close binaries
with a pulsar shows that the mass transfer does, in fact, stop. Perhaps if the
mass transfer is variable, it may reach a state sufficiently low that the pulsar
switches on, after which the pulsar wind itself can prevent further mass transfer.
Once more, the details of such a process are not understood.
M /M ≃ m/m ∼
The star that donates matter to a neutron star can also be a white dwarf;
in that case the orbital period is roughly Pb(s) ≃ 50/M (M⊙). In such a binary,
stable mass tranfer can only occur if the white dwarf has a mass less than
∼ 0.66M⊙; a more massive white dwarf will expand so rapidly with mass loss
that it is disrupted within one binary revolution (Verbunt & Rappaport 1988).
If the initial binary is too wide, loss of angular momentum is not important,
and mass transfer will only start when the donor expands as it ascends the
R/R, where R is the stellar
(sub)giant branch. In that case −
R increases with R, and thus the mass
radius. According to stellar evolution,
transfer rate is expected to be higher in wide binaries. Conservation of angular
momentum widens the orbit as mass is transferred from the (sub)giant to the
neutron star; typically, the orbital period has increased by a factor ∼ 7 by the
time that the whole envelope has been transferred. From that point on, the
binary will consist of a neutron star and a white dwarf. The mass transfer
puts an early end to the growth of the giant core, and the emerging white
dwarf is undermassive; typically Mwd ∼ 0.25M⊙. Tidal interaction during the
mass transfer leads to an orbit which is almost but not quite circular. The
observed white dwarf masses and the relation between orbital period and (small)
eccentricity provide strong support for this evolution scenario for recycled pulsars
with white dwarf companions (Phinney 1992).
It is observed that rather more millisecond pulsars are found in globular
clusters than can be formed from the currently observed X-ray binaries. A
solution could be that most recycled pulsars were spun up in binaries with
donors of intermediate mass, 1-3 M⊙ (Davies & Hansen 1998).
4. Formation through close encounters: testing the theory
With a growing number of binaries with neutron stars observed in globular
clusters, we can start comparing their observed properties with the theory of
formation. To do so, we refer to Eqs. 1 and 2, and note that the virial theo-
rem relates the velocity dispersion to the central density and core radius of the
6
Verbunt
Figure 2.
Central density as a function of core radius for globular
clusters in our Galaxy. Data from Harris (1996, revision of June 22
1999). Dotted lines indicate loci of constant formation rate Γ in the
cluster, according to Eq. 6; approximate numerical factors are indicated
on the right for R = R⊙ (applicable to tidal capture) and for a =
100R⊙, nb ≃ n (exchange encounters) where x indicates one encounter
per 10x yr. Dashed lines indicate lines of constant encounter rate γ for
one single system, according to Eq. 7; approximate numerical factors
are indicated on top.
cluster:
Therefore
v ∝ √ρo rc
1.5rc
2R
and
Γ ∝ ρo
1.5rc
2a
(5)
(6)
Γe ∝ ρo
In Figure 2 we plot lines of constant Γ in a graph showing the central density as
function of core radius for the globular clusters of our Milky Way. The numbers
indicated use the King model for the numerical factor in Eq. 5. The formation
rate of binaries with neutron stars can be high because the central density is
high, or because the cluster has a large core.
The situation is different for the probability that a binary, once formed, will
undergo a subsequent encounter which may change or destroy it. The rate at
which a binary undergoes close encounters is given by
γ = nAv ∝
ρo
v
a ∝
0.5
ρo
rc
a
(7)
Neutron stars in globular clusters
7
Figure 3.
Various -- preliminary -- tests on neutron stars in binaries
in globular clusters. a) Number of binaries with neutron stars, b) Slope
of the luminosity function of the X-ray sources, c) Orbital periods of
pulsar binaries, d) Pulsar pulse periods. For details see text.
Lines of constant γ are also shown in Figure 2. We see that in the clusters with
the highest central density, both close binaries (formed from tidal capture) and
wider binaries (formed from an exchange encounter) are affected by subsequent
encounters.
The exact encounter rates in a cluster depend on the mass function in the
core, hence on the mass segregation, on the fraction of stars in binaries, and
on the period distribution of the binaries. For encounters involving neutron
stars, the rates depend on the retention fraction of neutron stars. Also, it is not
clear how one should treat a collapsed cluster. All these factors are different in
different clusters. In the absence of detailed information on most clusters, we
can only perform preliminary tests, in which we ignore these factors.
One such test is that the probability that a cluster contains a bright X-ray
source scales with Γ, i.e. with the collision number Σ ≡ ρ2
c /v. The bright X-ray
sources pass this test (Verbunt & Hut 1987). Johnston et al. (1992) claim to find
that the probability that a cluster contains a recycled pulsar is less dependent
on ρc, viz. ∝ ρ1.5
; however, this lower dependence may be an artefact of their
assumption that v is the same in all clusters (compare eqs.1 and 6).
c r3
c
8
Verbunt
Some further tests are illustrated in Figure 3. Numbers used in these tests
for the X-ray sources / radio pulsars were taken from: ω Cen Rutledge et al. 2002
/ -- ; 47 Tuc Grindlay et al. 2001a / Freire et al. 2001; NGC6440 Pooley et al.
2002b / Lyne et al. 1996; M15 White & Angelini 2001 / Phinney 1992; NGC6752
Pooley et al. 2001a / D'Amico et al. 2002; NGC6397 Grindlay et al. 2002b / -- .
Numbers for other pulsars from: M53 Kulkarni et al. 1991; M5 Anderson et al.
1997; M4 Thorsett et al. 1999; NGC6342 Lyne et al. 1993; NGC6624 Biggs et
al. 1994.
a) ω Cen contains one neutron star X-ray binary and no (known) pulsar,
whereas 47 Tuc, NGC6440 and M 15 contain respectively 2, 4 and 2 neutron star
X-ray binaries and 20, 1 and 8 (known) pulsars. ω Cen indeed has a lower value
of Γ than the other three clusters.
b) The slope p of the X-ray luminosity function dN (Lx) ∝ Lx
−pd ln Lx
(including neutron star binaries, cataclysmic variables, pulsars, and magnetically
active binaries) is steep in 47 Tuc (p = 0.8), intermediate in NGC6752 and
NGC6440 (p = 0.5), and shallow in NGC6397 (p = 0.3). This may be related to
the high value of γ in NGC6397, which prevents a binary from evolving without
being interfered with; thus ordinary binaries in this cluster have largely been
destroyed (Pooley et al. 2002b).
c) The clusters M4 and M53 contain pulsars in binaries with long orbital
periods of 191 and 256 days; these clusters indeed have a low value of γ, i.e.
wide binaries are not affected by encounters. 47 Tuc, with a higher value for γ,
does not contain pulsars in binaries with periods longer than 2.3 d. (The binary
in M4 in fact has a companion third star at about 6000 R⊙; the outer binary
has only an expected life time of 108 yr.)
d) It has been an outstanding puzzle that all pulsars in 47 Tuc and NGC6752
have very short pulse periods (< 7.6 ms and < 9.0 ms), whereas those in M15
have pulse periods ranging up to 111 ms. Long periods also are found in NGC6342,
NGC6440 and NGC6624 (1.0 s, 289 ms and 379 ms). There is no obvious corre-
lation of the pulse period range with Γ or γ.
The encounter hypothesis for the formation of binaries with a neutron star
survives the preliminary tests.
5. Globular clusters in other galaxies
With the Einstein and Rosat satellites, X-ray sources in globular clusters of
M31 have been discovered. On the whole it appears that the properties of these
sources are not significantly different from those in our galaxy (Supper et al.
1997). Thanks to Chandra it has become possible to observe X-ray sources in
other galaxies, with fairly accurate positions. As a result, it has been found
that globular clusters in other galaxies contain X-ray sources with luminosities
above 1039 erg/s (Angelini et al. 2001, Sarazin et al. 2001, Kundu et al. 2002)
At such luminosities, the X-ray sources are probably black holes accreting from
a binary companion; this is confirmed by the soft X-ray spectrum when it can
be measured. Apparently the clusters in those galaxies do contain black holes in
binaries. This is not necessarily a significant difference with the clusters in our
galaxy:
if 20 % of the bright X-ray sources in globular clusters would contain
black holes, the expectation value for such binaries in our globular cluster system
Neutron stars in globular clusters
9
would be 2 or 3. For this expectation value the probability of finding zero is
appreciable.
A first contribution by XMM is the detection of a faint X-ray source in
Mayall II, the largest cluster of M31 (Verbunt, Meylan & Mendez, in prepara-
tion).
6. Summary
The overabundance in globular clusters of binaries with neutron stars is due to
formation of such binaries in close encounters. When the neutron star accretes
matter from its companion, it is an X-ray source. Chandra observations show
that there are 5 to 10 times as many neutron stars accreting at low rates than
at high rates. Whether this is related to their formation mechanism is not clear,
since a similar overabundance is found in the galactic disk where binaries with
neutron stars evolve from primordial binaries (Cornelisse et al. 2002). When the
neutron star stops accreting, it can become a radio pulsar. Most cataclysmic
variables in globular clusters are also formed in close encounters.
Tidal capture leads to a short orbital period, and has no associated recoil
velocity; exchange encounters favour wide binaries and do involve recoil. Most
neutron star binaries are found in or close to the cores of globular clusters. Most
binaries with a pulsar have relatively short orbital periods, and mass functions
that indicate a mass of ≃ 0.25M⊙ for their white dwarf companion.
In my
view, this indicates that these binaries were formed via tidal capture. Some
binaries have rather longer orbital periods, and some are far from the cluster
core. These binaries were probably formed via exchange encounters. From the
X-ray luminosity function it appears that magnetically active binaries -- that
make up the low -- luminosity end of the distribution -- are destroyed in dense
clusters.
Observations of globular clusters in other galaxies show that black holes
are present in them. It would be interesting to investigate whether any of the
low -- luminosity X-ray binaries in the globular clusters of our galaxy contains a
black hole. The small number of X-ray sources in clusters of our Galaxy hampers
comparison of their properties with those in globular clusters systems of other
galaxies.
References
Anderson, S., Wolszczan, A., Kulkarni, S., & Prince, T. 1997, ApJ, 482, 870
Angelini, L., Loewenstein, M., & Mushotzky, R. 2001, ApJ, 557, L35
Biggs, J., Bailes, M., Lyne, A., et al. 1994, MNRAS, 267, 125
Brisken, W., Benson, J., Goss, W., & Thorsett, S. 2002, ApJ, 571, 906
Clark, G. 1975, ApJ, 199, L143
Cornelisse, R., Verbunt, F., in 't Zand, J., et al. 2002, A&A, 392, 931
D'Amico, N., Possenti, A., Fici, L., et al. 2002, ApJ, 570, L89
Davies, M. 1997, MNRAS, 288, 117
Davies, M. & Hansen, B. 1998, MNRAS, 301, 15
10
Verbunt
Freire, P. 2002 http://www.naic.edu/∼pfreire/GCpsr.html
Freire, P., Camilo, F., Lorimer, D., et al. 2001, MNRAS, 326, 901
Grindlay, J., Heinke, C., Edmonds, P., , & Murray, S. 2001a, Science, 292, 2290
Grindlay, J., Heinke, C., Edmonds, P., et al. 2001b, ApJ, 563, L53
Harris, W. 1996, AJ, 112, 1487
Hartman, J. 1997, A&A, 322, 127
Hertz, P. & Grindlay, J. 1983, ApJ, 275, 105
Hut, P. & Verbunt, F. 1983, Nat, 301, 587
in 't Zand, J., Verbunt, F., Strohmayer, T., & et al. 1999, A&A, 345, 100
Johnston, H., Kulkarni, S., & Phinney, E. 1992, in X-ray binaries and the forma-
tion of binary and millisecond pulsars, ed. E. van den Heuvel & S. Rap-
paport (Dordrecht: Kluwer), 349 -- 364
Kulkarni, S., Anderson, S., Prince, T., & Wolszczan, A. 1991, Nat, 349, 47
Kundu, A., Maccarone, Th., & Zepf, S. 2002, ApJ, 574, L5
Lyne, A. 1994, in The evolution of X-ray binaries, ed. S. Holt & C. Day (New
York: AIP), 331 -- 338
Lyne, A., Biggs, J., Harrison, P., & Bailes, M. 1993, Nat, 361, 47
Lyne, A., Manchester, R., & D'Amico, N. 1996, ApJ (Letters), 460, L41
Phinney, E. 1992, Phil. Trans. R. Soc. London A, 341, 39
Pooley, D., Lewin, W., Homer, L., et al. 2002a, ApJ, 569, 405
Pooley, D., Lewin, W., Verbunt, F., et al. 2002b, ApJ, 573, 184
Ray, A., Kembhavi, A., & Antia, H. 1987, A&A, 184, 164
Rutledge, R., Bildsten, L., Brown, E., et al. 2002, ApJ in press
Sarazin, C., Irwin, J., & Bregman, J. 2001, ApJ, 556, 533
Supper, R., Hasinger, G., Pietsch, W., & et al. 1997, ApJ, 317, 328
Thorsett, S., Arzoumanian, Z., Camilo, F., & Lyne, A. 1999, ApJ, 523, 763
Tinker, J., Pinsonneault, M., & Terndrup, D. 2002, ApJ, 564, 877
Verbunt, F. 1993, ARA&A, 31, 93
-- . 2001, A&A, 368, 137
Verbunt, F., Bunk, W., Ritter, H., & Pfeffermann, E. 1997, A&A, 327, 602
Verbunt, F. & Hut, P. 1987, in The Origin and Evolution of Neutron Stars, IAU
Symposium No. 125, ed. D. Helfand & J.-H. Huang (Dordrecht: Reidel),
187 -- 197
Verbunt, F. & Meylan, G. 1988, A&A, 203, 297
Verbunt, F. & Rappaport, S. 1988, ApJ, 332, 193
Webb, N., Gendre, B. & Barret, D. 2002, A&A, 381, 481
White, N. & Angelini, L. 2001, ApJ, 561, L101
|
astro-ph/0611475 | 1 | 0611 | 2006-11-15T11:21:52 | Evidence for bipolar jets from the optical spectra of the prototypical symbiotic star Z Andromedae | [
"astro-ph"
] | We have studied optical spectra of the symbiotic star Z And, obtained during its latest outburst started in April 2006, with the aim of finding changes in the spectrum yielding clues to the nature of the hot component and its outbursts. The spectroscopic observations of Z And have been made using the 1.5-meter telescope at the Tartu Observatory, Estonia, and processed in a standard way. We have found high velocity satellites to the hydrogen Balmer emission lines. Starting from July 30, 2006, weak additional emission components at velocities of about +/-1150 km/s were detected. Their appearance near the outburst maximum and similarity to the emission features in another symbiotic star Hen 3-1341 imply fast collimated outflows from the hot component of Z And. This finding is consistent with the earlier results by several authors that symbiotic stars can emit bipolar jets at certain stages of their outbursts. A significant decrease in the temperature of the hot component in initial stages of the outburst was detected by the disappearance of the high excitation emission line from the spectrum. | astro-ph | astro-ph |
Astronomy&Astrophysicsmanuscript no. 6630
October 22, 2018
c(cid:13) ESO 2018
Letter to the Editor
Evidence for bipolar jets from the optical spectra of the
prototypical symbiotic star Z Andromedae ⋆
M. Burmeister1,2 and L. Leedjarv1
1 Tartu Observatory, 61602 Toravere, Estonia
2 Institute of Theoretical Physics, University of Tartu, Tahe 4, 51010 Tartu, Estonia
Received 24 October 2006 / Accepted 3 November 2006
ABSTRACT
Aims. We have studied optical spectra of the symbiotic star Z And, obtained during its latest outburst started in April 2006, with the
aim of finding changes in the spectrum yielding clues to the nature of the hot component and its outbursts.
Methods. The spectroscopic observations of Z And have been made using the 1.5-meter telescope at the Tartu Observatory, Estonia,
and processed in a standard way.
Results. We have found high velocity satellites to the hydrogen Balmer emission lines. Starting from July 30, 2006, weak additional
emission components at velocities of about ± 1150 km s−1 were detected. Their appearance near the outburst maximum and similarity
to the emission features in another symbiotic star Hen 3-1341 imply fast collimated outflows from the hot component of Z And.
This finding is consistent with the earlier results by several authors that symbiotic stars can emit bipolar jets at certain stages of
their outbursts. A significant decrease in the temperature of the hot component in initial stages of the outburst was detected by the
disappearance of the high excitation emission lines from the spectrum.
Key words. stars: outflows -- binaries: symbiotic -- stars: individual: Z And
1. Introduction
Z Andromedae is considered to be the prototype for the class of
symbiotic stars (Kenyon 1986; Corradi et al. 2003) -- interacting
binary stars consisting of a red giant and a hot compact com-
panion, mostly a white dwarf. The hot component ionises part
of the wind of the red giant, thus giving rise to the characteris-
tic "combination" spectrum where high excitation emission lines
are superimposed on the cool giant's spectrum. Z And has an or-
bital period of 757.5 days (Mikołajewska & Kenyon 1996) and
an orbital inclination of 47◦ ± 12◦ (Schmid & Schild 1997). Its
light curve represents classical symbiotic star outbursts, demon-
strating brightening of the star by 2 -- 3 magnitudes in the visual
region in every few years. Z And is also the only symbiotic star
for which coherent optical oscillations with the period of 28 min-
utes have been detected, most likely indicating the presence of
a strongly magnetic WD (Sokoloski & Bildsten 1999). Finally,
Z And is among about 10 known symbiotic stars, producing col-
limated outflows or jets (Brocksopp et al. 2004; Leedjarv 2004).
Extended collimated outflows perpendicular to the orbital plane
were detected from radio images during the 2000 -- 2002 outburst
(Brocksopp et al. 2004).
Behaviour of Z And during the outburst and the following
quiescence in 2000 -- 2003 was extensively studied and analysed
by Sokoloski et al. (2006). They proposed to call Z And and
possibly other classical symbiotic stars 'combination novae', in-
dicating that both dwarf nova-like accretion disk instability and
nova-like nuclear shell burning are involved in the outbursts
of classical symbiotic stars. According to the visual photomet-
ric observations from the AAVSO database, a new outburst of
Send offprint requests to: M. Burmeister, e-mail: [email protected]
⋆ Based on observations collected at Tartu Observatory, Estonia
Z And began in early June 2004. Having hardly recovered from
this outburst, the star started a new activity cycle in April 2006.
In the present paper we describe and analyse the optical spectra
of Z And, obtained mostly during decline from the optical maxi-
mum in 2006. Section 2 describes the observations and in Sect. 3
we present and analyse evidence for fast bipolar outflows during
the outburst of Z And. Section 4 contains conclusions.
2. Observations
Spectroscopic observations of Z And have been carried out at
the Tartu Observatory, Estonia, using the 1.5-meter telescope
equipped with the Cassegrain grating spectrograph. Two dif-
ferent CCD cameras from the SpectraSource Instruments were
used until March 2006. Later on the spectra are registered with
the Andor Technologies CCD camera Newton USB-207 with
400 × 1600 chip, pixel size 16 × 16 µm, Peltier cooled. A few oc-
casional spectra of Z And have been obtained since September
1997, mostly in the region of Hα. A more systematic monitoring
started on July 30, 2006, after detecting the signs for bipolar jets.
Most of the red spectra cover the wavelength region
6480−6730 Å, with linear dispersion about 0.2 Å pixel−1. Blue
spectra are mostly taken in the 4600−5060 Å region with
∼ 0.3 Å pixel−1. Some of the lower dispersion spectra (about
0.5 Å pixel−1 and 0.6 Å pixel−1, respectively) are also used. The
spectra were reduced using the software package MIDAS pro-
vided by ESO. After standard procedures of subtracting the bias
and sky background, the spectra were reduced to the wavelength
scale using the Th-Ar hollow cathode lamp spectrum for cali-
bration. Thereafter, the spectra were normalised to the contin-
uum, and positions, peak intensities, and equivalent widths of
the emission lines were measured.
2
M. Burmeister and L. Leedjarv: Bipolar jets from Z And
y
t
i
s
n
e
t
n
i
e
v
i
t
a
l
e
R
25
20
15
10
5
0
-1000
0
Radial velocity (km s-1)
1000
Fig. 1. Hα line profiles from August 7 (solid line) and September
10 (dotted line). The former shows an absorption component.
Table 1. Radial velocities of the jet components of Hα and Hβ
line in km s−1
Date
2006
Jul 30
Aug 7
Aug 9
Aug 10
Sep 1
Sep 10
Sep 12
Sep 13
Sep 15
Sep 21
Sep 24
Sep 27
Oct 6
Oct 12
Mean
Hα
JD -
2 400 000 Blue
-1114
53947.5
-1115
53955.5
53957.4
-1106
-1134
53958.5
-1243
53980.4
-1150
53989.4
-1193
53991.4
53992.5
-1180
-1156
53994.4
-1081
54000.4
-1114
54003.5
-1115
54006.4
54015.3
-1106
-1124
54021.4
-1138±41
Red
1391
1271
1275
1280
1165
1157
1145
1067
1059
1100
1087
1107
1110
1108
1166±96
Blue
-969
-1288
-1180
-1217
-1191
-1176
-1070
-1162
-1150
Hβ
Red
1376
1160
1084
1127
1061
1039
1098
1137
1134
-1196
-1160±82
1134
1135±88
3. Results and discussion
Both Hα and Hβ lines have a central emission component at
radial velocities of about 30 km s−1. The shape of the profiles
is asymmetric with the blue wing distorted by absorption. On
August 7 and 9, the blue step of the Hα line deepens into an ab-
sorption component (Fig. 1). A similar absorption is seen in Hβ
on September 1.
The most striking features of these Balmer lines, however,
are the satellite emission components (see Fig. 2), detected for
the first time on July 30, 2006 in the Hα line. We are not aware
of observations of similar features in the spectrum of Z And be-
fore. On September 14, 2006, a telegram on the detection of jets
in Z And was published by Skopal & Pribulla (2006). Very sim-
ilar additional emission features were discovered in the spectra
of the symbiotic stars Hen 3-1341 by Tomov et al. (2000), and
StHα 190 by Munari et al. (2001), who interpreted them to be
emitted by the collimated bipolar jets.
We fitted these emission components with a Gaussian to find
their radial velocities. The results are given in Table 1. On av-
erage the RVs are about ±1150 km s−1. Given the asymmetric
Hα
Feb 26, 2006
Jul 30, 2006
Aug 7, 2006
Aug 9, 2006
Sep 1, 2006
Sep 15, 2006
Sep 27, 2006
Oct 12, 2006
-1000
0
y
t
i
s
n
e
t
n
i
e
v
i
t
a
l
e
R
1.4
1.2
1
1.4
1.2
1
1.4
1.2
1
1.4
1.2
1
1.4
1.2
1
1.4
1.2
1
1.4
1.2
1
1.4
1.2
1
Hβ
1.4
1.2
1
1.4
1.2
1
1.4
1.2
1
1.4
1.2
1
1.4
1.2
1
1000
-1000
1000
Radial velocity (km s-1)
0
Fig. 2. Time evolution of the jet components around the Hα and
Hβ lines. Our last spectrum before the outburst from February
26, although more noisy, does not show additional emission
components.
shape of the additional components, the accuracy of those ve-
locities is no better than about 40−50 km s−1. If we assume that
the jets are perpendicular to the orbital plane and the inclina-
tion of the orbit is 47◦, the true velocities of the jets are close to
±1700 km s−1.
The equivalent widths of the supposed jet components form
about 2% of that of the central emission, but these estimates are
uncertain, as the small emissions are partly blended in the central
component. The measured values of EWs are plotted in Fig. 3.
At least in some of the red spectra, weak [N II] 6548 and
6584 Å emission lines can be distinguished at the low-velocity
sides of Hα jet components. Where measurable, the velocities of
the [N II] lines occur at about −45 km s−1.
According to the AAVSO light curve, the optical maximum
of Z And took place in early July 2006 (around JD 2 453 920 −
930). Our last optical spectrum before the outburst was made on
February 26, 2006. No signs of jets are seen in this spectrum, so
the jets must have been formed at some time between February
and July. For most of the time of our observations, the jet features
in the spectrum have been rather stable. Smaller EWs on July 30
possibly indicate the early phase of jet development.
Such an appearance of the jets at the time of the optical max-
imum fits the behaviour of another symbiotic star Hen 3-1341
as described by Munari et al. (2005) well. In general, collimated
bipolar jets are observed from several types of stars, such as, e.g.,
X-ray binaries, supersoft X-ray sources, young stellar objects,
etc., and of course, from active galactic nuclei. The standard sce-
M. Burmeister and L. Leedjarv: Bipolar jets from Z And
3
He I 4713
He II 4686
3
2.5
2
W
E
1.5
1
0.5
0
53960
53980
54000
JD - 2 400 000
54020
Fig. 3. EWs of the blue (filled circles) and red (empty circles) jet
component of the Hα line and of the blue (filled triangles) and
red (empty triangles) jet component of the Hβ line.
nario for the production of jets (e.g., Livio 1997; Lynden-Bell
2003, and many later references) includes an accretion disc that
is threaded by a vertical magnetic field. In addition, an energy
or wind source associated with the central accreting object is
needed. This requirement explains, for instance, the absence of
jets from magnetic cataclysmic variables, in spite of the presence
of an accreting white dwarf and magnetic field. The behaviour
of Hen 3-1341 well demonstrates that the jets are emitted only
in the early and the brightest phase of the outburst, and that they
are fed by the wind from the outbursting component (Munari et
al. 2005). The outburst of Hen 3-1341, however, lasted longer
(from 1998 to 2004) than those of Z And usually do.
The wings of the Hα main emission component extend to
about ±1000 km s−1, which may indicate fast wind from the out-
bursting hot component. On the other hand, indications of low-
velocity outflow can be seen in the lines of He I 4471, 4713, and
5016, which present P Cygni profiles with deep absorption com-
ponents. The development of the P Cygni profile of He I 4713
is seen in Fig. 4. Velocity of the outflow measured from these
lines is between −50 and −100 km s−1. This is comparable to
the velocity −90 kms−1 found by Skopal et al. (2006) during the
2000 -- 2003 outburst. At the same time, the He I 4922 and 6678
lines are fully in emission or have only a very weak absorption
component. Figure 5 presents an example.
The present outburst of Z And is similar to that of Hen 3-
1341 also by the behaviour of other emission lines. It is known
that high excitation emission lines tend to weaken during the
outbursts of symbiotic stars. In Hen 3-1341, the He II 4686 and
the Raman scattered O VI line at 6825 disappeared completely.
Similar changes in the spectrum of Z And during the 2000 -- 2003
outburst were reported by Skopal et al. (2006) and Sokoloski
et al. (2006). We also an detected almost complete disappear-
ance of the Raman scattered line at 6825 and of the He II 4686
line in the spectrum of Z And during the present outburst in
2006. Figure 6 presents the comparison of those spectral lines
from the quiescence in September 1997 and from the outburst in
September 2006. In Fig. 4, the gradual reappearance of the He II
4686 line is seen while the brightness of the star decreases.
We made an attempt to estimate the effective temperature of
the hot component during our observations by using the method
Aug 9, 2006
Sep 1, 2006
Sep 15, 2006
Sep 27, 2006
Oct 12, 2006
y
t
i
s
n
e
t
n
i
e
v
i
t
a
l
e
R
2
1.6
1.2
0.8
2
1.6
1.2
0.8
2
1.6
1.2
0.8
2
1.6
1.2
0.8
2
1.6
1.2
0.8
2
1.6
1.2
0.8
2
1.6
1.2
0.8
2
1.6
1.2
0.8
2
1.6
1.2
0.8
2
1.6
1.2
0.8
-200
0
200
-200
0
200
Radial velocity (km s-1)
Fig. 4. Time evolution of the He I 4713 and He II 4686 lines,
representing matter outflow and outburst state, respectively. The
profiles are given on the same scale for comparison.
He I 4713
He I 6678
3
2.5
2
1.5
1
y
t
i
s
n
e
t
n
i
e
v
i
t
a
l
e
R
-300
-200
-100
100
Radial velocity (km s-1)
0
200
300
Fig. 5. Example of an He I 4713 line and He I 6678 line from
September 27, 2006.
proposed by Iijima (1981). Like Sokoloski et al. (2006) we used
equivalent widths of Hβ and He II 4686 instead of fluxes and
neglected the He I 4471 line. The temperatures in August and
September tend to be low, falling in the range 75 000 -- 90 000 K,
and begin to rise at the end of September. The results are given in
Table 2. A similar decline of Thot at the brightness maximum and
its gradual increase after the maximum was found by Sokoloski
4
M. Burmeister and L. Leedjarv: Bipolar jets from Z And
Table 2. Effective temperatures of the hot star estimated by the
Iijima method
Date
2006
Aug 9
Sep 10
Sep 12
Sep 13
Sep 15
Sep 21
Sep 24
Sep 27
Oct 12
JD -
2 400 000
53957.438
53989.398
53991.346
53992.549
53994.394
54000.363
54003.519
54006.388
54021.368
Temperature
(K)
76 000
91 000
80 000
75 000
81 000
87 000
91 000
100 000
114 000
He II 4686
O VI 6825
20
15
10
5
y
t
i
s
n
e
t
n
i
e
v
i
t
a
l
e
R
2.5
2
1.5
1
0
4670
4680
4690
0.5
4700
6800
6820
6840
6860
Wavelength (A)
Fig. 6. Left: He II 4686 line of Z And from September 27, 1997
(solid line) and September 27, 2006 (dotted line). Right: Raman
scattered line O VI 6825 from September 2, 1997 (solid line)
and September 27, 2006 (dotted line). Disappearance of the high
ionization emission lines refers to an outburst state.
et al. (2006) during the 2000 -- 2003 outburst. On the contrary,
Thot increased from the quiescence value of <∼ 150 000 K to
about 180 000 K during the smaller scale short outburst in 1997.
Those facts imply similarity with another classical symbiotic
star AG Dra in which Gonz´alez-Riestra et al. (1999) have dis-
tinguished between cool and hot outbursts, based on the changes
in the He II Zanstra temperature. The outbursts of Z And in 2000
and 2006 can be considered as cool ones and that in 1997 as hot
outburst. It might be of interest to note that AG Dra has also been
in outburst since July 2006. We will describe spectroscopic be-
haviour of AG Dra in another forthcoming paper. Here we only
note that during the present outburst the He II 4686 line and the
Raman scattered O VI line 6825 also became very weak in the
spectrum of AG Dra. This is different from smaller scale out-
bursts of AG Dra in the late 1990s -- early 2000s when all the
emission lines became stronger (e.g., Tomov & Tomova 2002;
Leedjarv et al. 2004).
4. Conclusions
From our observations of the prototypical symbiotic star Z And
we can conclude the following:
(1) The additional emission components of the hydrogen Hα
and Hβ lines at ±1150 km s−1 indicate ejection of bipolar jets
starting from late July 2006, and persisting for at about three
months, at least. Taking into account the orbital inclination 47◦,
ejection velocity of the jets would be about 1700 km s−1.
(2) Although Z And has been extensively studied over the
past decades, no such jets were detected from the optical spectra
before. We have some spectra in the Hα region near the bright-
ness maximum of the 2004 outburst, but no additional emission
components can be found from these. According to Skopal et al.
(2006), no signs of jets could be seen during the maximum of the
2000 -- 2003 outburst. Spectra by Sokoloski et al. (2006) from the
same time have too low resolution to confirm the existence or ab-
sence of additional emissions. We also have one spectrum from
August 25, 2001, which was taken almost at the same time when
Brocksopp et al. (2004) discovered the radio jets from Z And.
In our spectrum, the jets are not seen. This most likely means
that the jets are transient phenomenon, emitted only for a short
time in the brightest phase of the outbursts, thus confirming the
standard scenario for jet ejection in which an additional source
of energy or wind is needed besides the accreting central body
and magnetic field.
(3) The high excitation lines He II 4686 and Raman scattered
O VI 6825 almost disappeared from the spectrum of Z And in the
early phase of the outburst, confirming that the temperature of
the hot component decreased to about 75 000 K, as was found by
the Iijima method. In September 2006, about 3 -- 4 months after
the start of the outburst, the He II 4686 line started to become
stronger again.
The ongoing outburst of Z And has shown that new aspects
can be found in the behaviour of such well studied symbiotic
stars. Together with the combination nova model proposed by
Sokoloski et al. (2006), new clues for understanding the out-
bursts of AG Dra, Hen 3-1341, and other classical symbiotic
stars can be obtained. Continuing monitoring of the ongoing out-
burst of Z And is strongly encouraged.
Acknowledgements. The authors thank Dr. Kalju Annuk for taking some of the
spectra of Z And and for useful comments on the manuscript of the paper. We
also thank our referee, Dr. Michael Bode, for his suggestions. The present study
was supported by the Estonian Ministry of Education and Research under the
target financed project 0062464S03 "Structure, chemical composition, and evo-
lution of stars" and by the Estonian Science Foundation grant No 6810.
References
Brocksopp, C., Sokoloski, J.L., Kaiser, C., et al. 2004, MNRAS, 347, 430
Corradi, R.L.M., Mikołajewska, J., & Mahoney, T.J. (eds.) 2003, Symbiotic Stars
Probing Stellar Evolution, ASP Conf. Series, Vol. 303
Gonz´alez-Riestra, R., Viotti, R., Iijima, T., & Greiner, J. 1999, A&A, 347, 478
Iijima, T. 1981, In: Photometric and Spectroscopic Binary Systems, eds. E.B.
Carling & Z. Kopal, Dordrecht, Kluwer, 517
Kenyon, S.J. 1986, The Symbiotic Stars, Cambridge University Press
Leedjarv, L. 2004, Baltic Astronomy, 13, 109
Leedjarv, L., Burmeister, M., Mikołajewski, M., et al. 2004, A&A, 415, 273
Livio, M. 1997, In: Wickramasinghe, D.T., Bicknell, G.V., Ferrario, L. (eds.),
Accretion Phenomena and Related Outflows, ASP Conf. Series, Vol. 121,
845
Lynden-Bell, D. 2003, MNRAS, 341, 1360
Mikołajewska, J., & Kenyon, S.J. 1996, AJ, 112, 1659
Munari, U., Tomov, T., Yudin, B. F., et al. 2001, A&A, 369, L1
Munari, U., Siviero, A., & Henden, A. 2005, MNRAS, 360, 1257
Schmid, H.M., & Schild, H. 1997, A&A, 327, 219
Skopal, A., & Pribulla, T. 2006, Astronomers Telegram, 882
Skopal, A., Vittone, A.A., Errico, L. et al. 2006, A&A, 453, 279
Sokoloski, J.L., & Bildsten, L. 1999, ApJ, 517, 919
Sokoloski, J.L., Kenyon, S.J., Espey, B.R., et al. 2006, ApJ, 636, 1002
Tomov, N. A., & Tomova, M. T. 2002, A&A, 388, 202
Tomov, T., Munari, U., & Marrese, P. M. 2000, A&A, 354, L25
List of Objects
'Z And' on page 1
M. Burmeister and L. Leedjarv: Bipolar jets from Z And
5
'Hen 3-1341' on page 2
'StHα 190' on page 2
'Hen 3-1341' on page 2
'AG Dra' on page 4
|
astro-ph/0010145 | 1 | 0010 | 2000-10-07T04:02:46 | Dynamics of Stellar Collisions | [
"astro-ph"
] | I compare gas-dynamical and stellar-dynamical models of collisions. These two models have distinctly different physics; for example, shocks introduce irreversibility in gas systems, while stellar systems evolve in a completely reversible fashion. Nonetheless, both models yield broadly similar results, suggesting that analogies between gas and stellar dynamics have some heuristic validity even applied to collisions. | astro-ph | astro-ph |
Stellar Collisions, Mergers, and Their Consequences
ASP Conference Series
M. Shara, ed.
Dynamics of Stellar Collisions
J. E. Barnes
Institute for Astronomy, University of Hawaii, 2680 Woodlawn Drive,
Honolulu HI, 96822
I compare gas-dynamical and stellar-dynamical models of
Abstract.
collisions. These two models have distinctly different physics; for exam-
ple, shocks introduce irreversibility in gas systems, while stellar systems
evolve in a completely reversible fashion. Nonetheless, both models yield
broadly similar results, suggesting that analogies between gas and stellar
dynamics have some heuristic validity even applied to collisions.
1.
Introduction
Stars and galaxies are prototypical examples of self-gravitating systems. While
both are held together by gravity, they differ in size, makeup, and structure.
From a dynamical point of view, the basic difference is that stars are made up of
particles which often undergo collisions, while galaxies are made up of particles
which almost never collide (this conference notwithstanding).
Analogies between stellar and gaseous systems run deep. For example, there
is a close relationship between spherical stellar systems with isotropic distribu-
tion functions f = f (E) and gaseous systems with barotropic equations of state
P = P (ρ). The stability properties of spherical stellar and gaseous systems
provide further analogies; an isotropic stellar system with df /dE < 0 is stable if
the barotropic gas-sphere with the same density profile is stable (Antonov 1962;
Lynden-Bell 1962). This is a sufficient but not necessary condition; on the whole,
spherical stellar systems seem to be more stable than their gaseous counterparts.
For flattened systems the situation is more complex. Infinite, uniformly rotating
sheets of gas and stars have very similar stability criteria (Toomre 1964). On
the other hand, certain finite, uniformly-rotating stellar disks (Kalnajs 1972) are
much less stable than gaseous disks with the same density profile.
Questions of existence and stability exploit analogies between stellar and
gaseous systems at or near equilibrium. In this paper I use numerical simulations
to examine how such analogies work in situations which are far from equilibrium:
collisions. Colliding stars and colliding spherical galaxies evolve in superficially
similar ways, as seen in Fig. 1. Here the top row of frames shows a collision
of two gas-spheres, while the bottom row shows the analogous collision of two
spherical systems of collisionless particles, hereafter labeled "grit"1. A more
1In my original presentation I used "dust" instead of "grit", borrowing this term from cosmology.
However, S. Shapro pointed out that "dust" is only applicable to a cold, pressureless medium.
Casting about for a suitable term, I chose "grit" since this conveys some resilience to pressure.
1
2
Barnes
Figure 1.
Gas (top) and grit (bottom) versions of the same collision. Ini-
tially the two bodies approach on a parabolic relative orbit, passing each
other in a counter-clockwise direction. The mass ratio is M1/M2 = 2 and the
targeted pericentric separation rp = 0.5 is half the radius of the larger body.
detailed comparison of the two collisions in Fig. 1 reveals significant differences.
For example, it's evident that orbit decay doesn't work in quite the same way;
contrasting the third pair of frames, the two bodies are much closer in the gas
model than they are in the grit model. Moreover, grit evolves reversibly, while
gas evolves irreversibly; by integrating the merged grit model backward in time,
one can unscramble the wreckage and recover the initial conditions (van Albada
& van Gorkom 1977), but the presence of shocks in the gas model precludes such
reversals.
1.1. Gas model
Technically, the interior of a normal star is a partly-ionized plasma in local
equilibrium with a black-body radiation field; support against gravity is provided
by some combination of thermal and radiation pressure. But for low-mass stars
the latter is insignificant, and to a good approximation the pressure is given by
the equation of state for an ideal gas.
Gas dynamics unfolds in a space of three dimensions; the dynamical vari-
ables are functions of position r and time t. Two such variables are the density
ρ = ρ(r, t) and the velocity field v = v(r, t). One more variable is needed to
represent the thermodynamic state of the gas; in these calculations I use the
entropy function a = a(S), which enters directly in the equation of state:
P = a(S)ργ ,
(1)
where P is the pressure, and γ is the ratio of specific heats, here assigned the
value γ = 5/3 appropriate for a monatomic gas.
Dynamics of Stellar Collisions
I adopt the following dynamical equations:
∂ρ
∂t
+
∂
∂r
· (ρv) = 0 ,
∂v
∂t
∂a
∂t
+ (cid:18)v ·
+ (cid:18)v ·
∂
∂
1
ρ
∂P
∂r
−
∂Φ
∂r
∂r(cid:19) v = −
∂r(cid:19) a = (γ − 1)ρ1−γ uv .
,
3
(2)
(3)
(4)
Here eq. 2 represents conservation of mass and eq. 3 represents conservation of
momentum; the gravitational potential Φ is calculated from Poisson's equation.
Eq. 4 describes the evolution of the entropy function; the flow is adiabatic,
conserving the entropy of each fluid element, except where the viscous heating
function uv > 0. A standard form of artificial viscosity (Monaghan & Gingold
1983) is used to implement the increase in gas entropy due to shocks.
I solve these dynamical equations using Smoothed Particle Hydrodynamics
or SPH (eg. Monaghan 1992). The code is similar to "TREESPH" (Hernquist &
Katz 1989); it includes a hierarchical algorithm to compute gravitational forces,
individual smoothing radii hi set so that that each particle i interacts with a fixed
number of neighbors, and individual time-steps adjusted to satisfy a Courant
condition. The code doesn't include terms proportional to ∇hi which arise
when the SPH equations are derived from a Hamiltonian (Nelson & Papaloizou
1993, 1994). Since the code integrates Eq. 4 instead of the equivalent energy
equation, the neglect of these ∇hi terms leads to imperfect energy conservation
(Hernquist 1993). Most of the calculations presented below conserve energy to
∼ 1% or better, so the neglect of these terms is probably not critical in this
application.
1.2. Grit model
The dominant mass components in typical galaxies are stars and dark matter. In
the absence of conflicting evidence the latter is often assumed to be composed
of particles with masses much less than 106M⊙ (cf. Lacey & Ostriker 1985).
The gravitational potential of a galaxy is quite smooth and individual stars or
dark matter particles follow smooth trajectories, undisturbed by collisions, close
encounters, or any effect due to the discrete nature of the mass.
Stellar dynamics unfolds in a space of six dimensions; the phase-space dis-
tribution f = f (r, v, t) is a function of position r, velocity v, and time t. In the
limit appropriate for galaxies, f evolves according to the Collisionless Boltzmann
Equation or CBE:
∂f
∂t
+ v ·
∂f
∂r
−
∂Φ
∂r
·
∂f
∂v
= 0
(5)
Like eq. 2, eq. 5 is a continuity equation, conserving the mass of the system.
But eq. 5 describes an incompressible flow in six dimensions; dynamics moves
elements of phase-fluid around but conserves the value of f associated with each.
I solve the CBE using standard N-body techniques (eg. Barnes 1998). The
code uses a hierarchical algorithm to compute gravitational forces and a simple
leap-frog integrator to advance particle coordinates.
4
Barnes
1.3. Polytropes
I chose to collide simple polytropes, instead of using stellar models. While ac-
curate simulations of merging stars require realistic stellar models (Sills & Lom-
bardi 1997), polytropes are still appropriate for comparing the basic physics of
stellar and galactic collisions. Polytropes are easily constructed, and their prop-
erties are well-understood (eg. Chandrasekhar 1939). Moreover, grit models of
polytropes have simple distribution functions (Eddington 1916) and the stabil-
ity of these systems has been studied using N-body simulations (H´enon 1973;
Barnes, Goodman, & Hut 1986).
Both gas and grit polytropes obey the following relationship between mass
density ρ and gravitational potential Φ:
ρ = ρ(Φ) = ρ1(1 − Φ/Φ1)n .
(6)
Here Φ1 is the value of the potential on the surface of the polytrope and ρ1 is a
constant with units of density. By construction, gas polytropes are marginally
stable to convection, and this implies that the index n is related to the ratio of
specific heats by n = 1/(γ − 1). For grit polytropes, n is a free parameter, and
the distribution function takes the form
f = f (E) = (cid:26) f1(1 − E/Φ1)n−3/2
0
if E < Φ1,
otherwise,
(7)
where f1 is a constant with units of phase-space density.
For a monatomic gas, γ = 5/3, and the appropriate index n = 3/2. The cor-
responding grit polytrope has an unusually simple structure; f has the constant
value f1 within the six-dimensional volume defined by E = 1
2 v2 + Φ(r) < Φ1,
and vanishes everywhere else.
To construct realizations of polytropes I first tabulated the density profile
ρ(r), potential Φ(r), and pressure P (r). Grit models were generated by chosing
initial positions ri and velocities vi for each particle according to the distribu-
tion function f (r, v) = f ( 1
2 v2 + Φ(r)) (H´enon 1973). Likewise, gas models were
generated by chosing initial positions ri according to the density profile ρ(r),
setting initial velocities vi = 0, and assigning initial entropy function values
ai = P (ri)/ργ (ri). In both cases all particles had equal masses. Since particle
coordinates were assigned independently, these initial configuration have Pois-
sonian fluctuations. Such fluctuations are inevitable in N-body simulations, but
they are much larger than the fluctuations normally present in SPH calculations.
Therefore, the gas realizations were relaxed by evolving them with the SPH code
using a velocity-damping term. This has the effect of "ironing out" the initial
density fluctuations without disturbing the overall density profile.
2. Collision Sample
The dynamics of a collision depend on the relative orbit and mass ratio of the
participants as well as their internal structures. The collisions reported here were
restricted to parabolic (zero-energy) orbits, which are appropriate for collisions
in globular clusters and similar environments. I used n = 3/2 polytropes and
assumed a linear relationship between mass and radius; these choices roughly
Dynamics of Stellar Collisions
5
Figure 2.
Energies vs. time for all collisions, each labeled by (M1/M2, rp).
The three solid curves in each panel are from the gas models; from top to
bottom, they show the thermal energy T (t), kinetic energy K(t), and potential
energy U (t). Likewise, the two dashed curves are from the grit models; they
show the kinetic and potential energies.
caricature the properties of low-mass main-sequence stars. It may ultimately be
worth relaxing these restrictions, but even with them in place, two parameters
must still be chosen for each collision: the mass ratio and the pericentric sepa-
ration of the initial relative orbit. I adopted mass ratios M1/M2 = 1 and 2, and
studied collisions with pericentric separations rp ranging from zero (head on)
to R1 + R2 (grazing). To simplify comparisons between collisions with different
mass ratios, I stipulated that all cases have the same total mass.
Since G is the only dimensional constant in these calculations, it's con-
venient to adopt units with G = 1. The polytropes used for the equal-mass
collisions (M1/M2 = 1) have masses M1 = M2 = 0.75 mass units and radii
R1 = R2 = 0.75 length units; for the gas models, the initial entropy func-
tion value is a(S1) = a(S2) = 0.2889. The polytropes used for the unequal-
mass collisions (M1/M2 = 2) have masses M1 = 1.0 and M2 = 0.5 and radii
R1 = 1.0 and R2 = 0.5; the initial entropy function values are a(S1) = 0.4240
and a(S2) = 0.1682.
For both mass ratios I considered four collisions with pericentric separations
rp = 0, 0.5, 1, and 1.5 (= R1 + R2); each collision was run with both gas and grit
models, using 24576 particles for each calculation. I ran all collisions with rp ≤ 1
until the participants merged and relaxed to near-equilibrium configurations.
The grazing collisions were not run to completion; orbit decay is very gradual
in such wide passages and the time required to reach merger seemed excessive.
Fig. 2 summarizes the evolution of the collisions. In each case the partici-
pants have a first pericentric passage at t ≃ 2.5, marked by sharp minima in the
6
Barnes
Figure 3.
First passages for equal-mass collisions. Each row shows a dif-
ferent collision at times equally spaced between t = 2 and t = 3; rp increases
from top to bottom. Contours indicate density on the orbital plane in steps
of a factor of 4; dashed lines show grit, solid lines show gas, with a heavier
contour for ρ = 1. Half-tones indicate shocks.
curves of potential energies U (t). In the grit models (dashed lines) this potential
energy is entirely invested in kinetic energy K(t), which has a maxima opposing
the minima of U (t). In the gas models (solid lines) the gravitational energy is
shared between K(t) and the thermal energy T (t), with the relative apportion-
ment depending on the collision parameters. Head-on collisions produce sharp
peaks in T (t) as gravitational energy heats the gas; the stored thermal energy
drives a subsequent re-expansion and bounce before the system settles down.
On the other hand, the collisions with rp ≥ 1 put most of the available gravi-
tational energy into bulk motion, producing modest peaks in K(t) while barely
disturbing T (t).
Dynamics of Stellar Collisions
7
Figure 4.
First passages for unequal-mass collisions. Each row shows a
different collision at times equally spaced between t = 2 and t = 3; rp increases
from top to bottom. Contours indicate density on the orbital plane in steps
of a factor of 4; dashed lines show grit, solid lines show gas, with a heavier
contour for ρ = 1. Half-tones indicate shocks.
2.1. First passages
Figs. 3 and 4 report on the first passages of all collisions. Here the contours
represent density in the orbital plane, evaluated using SPH-style interpolation;
dashed contours show grit density, while solid contours show gas density. Shaded
areas show where uv > 0; these track the shocks propagating through the gas.
The most obvious differences between the gas and grit models occur in the
head-on collisions (top rows of Figs. 3 and 4). In the equal-mass head-on collision
the two grit polytropes pass directly through each other. As they do so they
are briefly compressed by their mutual gravity; at later times they rebound,
spraying grit particles onto loosely-bound orbits, and fall back to form a single
merged object (van Albada & van Gorkom 1977). The gas polytropes obviously
8
Barnes
can't interpenetrate; a shock develops when they first touch. Subsequently the
shocked gas forms a disk perpendicular to the collision axis and bounded on
the left and right by strong shocks. In the final image of this collision (upper
right in Fig. 3) no trace of the two gas polytropes remains; the shocked gas
forms a single object which expands due to its large thermal energy content. On
the other hand, the grit polytropes, having passed through each other, are still
distinct.
The unequal-mass head-on collision produces a more complex morphology.
As the small grit polytrope plunges through its larger partner it generates a
"wake" -- a region of higher density where gravitational focusing concentrates
grit from the large polytrope towards the collision axis. This wake is similar to
the mutual compression of the two grit polytropes in the equal-mass collision;
in both cases the energy needed is extracted from orbital motion, resulting in
rapid orbit decay. The gas model produces a very different result; as the small
polytrope plows into its partner it creates a strong bow shock. This shock pushes
gas away from the collision axis, evacuating a cavity in the wake of the small
gas polytrope. Meanwhile, a weaker shock travels backward through the small
polytrope. By the last image (upper right in Fig. 4) the bow shock has reached
the surface of the large polytrope, while the cavity has been filled in; the small
gas polytrope has lost most of its forward momentum and lies, still largely intact,
near center of the combined object. In contrast, at this time the grit polytropes
have temporarily separated, though they will soon fall back together.
Many of the dynamical effects just described also play a role in the off-
axis collisions. In the equal-mass rp = 0.5 collision the bodies of the two gas
polytropes are traversed by parallel shocks which slow their relative motion and
increase their entropy. The unequal-mass version of this collision produces an
asymmetric bow shock as the small polytrope moves through the outer layers
of its companion, leaving a temporarily furrow behind.
In contrast, the grit
versions of these collisions again show regions of enhanced density in the wakes
of the interpenetrating polytropes. Less dramatic effects are seen in the wide
collisions; the rp = 1 collisions are basically "kinder and gentler" versions of
their closer counterparts, with weaker shocks and more subtle wakes. Finally,
the grazing collisions (rp = 1.5), shown at the bottom of Figs. 3 and 4, don't
develop significant shocks; these passages involve only gravitational interactions.
2.2. Orbit decay
Although their initial orbits are parabolic, all of the off-center collisions leave
the participants on bound orbits after their first passage. This orbit decay is
generally due to the transfer of orbital energy to internal degrees of freedom, but
the way this transfer takes place may be quite different in gas and grit models.
From Fig. 2 it's already clear that the interplay between gravity and gas-
dynamics in orbit decay is complex. The plots of U (t) for the collisions with rp =
0.5 and 1.0 show two or more minima, each representing a different pericentric
passage. In three of these collisions the grit models return to pericenter before the
corresponding gas models, but the opposite order is seen in the (M1/M2, rp) =
(2, 0.5) collision. Thus compared to the grit models, gas can either accelerate or
delay orbit decay.
Dynamics of Stellar Collisions
9
Figure 5.
Orbital trajectories for all off-center collisions. Solid lines are
gas results, while dashed lines are grit results. Each is labeled by (M1/M2, rp);
the scale bar under each label is 1 unit long.
Fig. 5 shows orbital trajectories for all off-center collisions. These trajec-
tories confirm that in the (M1/M2, rp) = (1, 0.5), (1, 1.0), and (2, 1.0) collisions
the grit models undergo more rapid orbit decay than their gas counterparts, and
that the opposite is true for the (2, 0.5) collision. Moreover, they reveal a further
puzzle; in grazing collisions the outgoing orbits of the gas polytropes are more
tightly bound than those of their grit doppelgangers.
Closer inspection of Fig. 5 suggests a mechanism which might delay orbit
decay in some gas models. Collision (M1/M2, rp) = (1, 0.5), in which the differ-
ence between the gas and grit models is quite dramatic, shows it most clearly:
on their first passage the gas polytropes "bounce" off each other as if elastic. Of
course, this is not too surprising; an ideal gas is perfectly elastic in the absence
of shocks. Moreover, the corresponding panel of Fig. 2 shows a sharp peak in
the thermal energy T (t) during this passage, and Fig. 3 shows that the gas is
compressed to roughly twice its initial density at the moment of closest approach
(t ≃ 2.5). Similar bouncing trajectories are seen in the other off-axis collisions.
On the other hand, the rapid orbit decay of the (M1/M2, rp) = (2, 0.5)
collision points to a different effect. Here as in the head-on cases the orbits
of the gas polytropes decay faster than the orbits of their grit counterparts.
For the head-on collisions this is no great mystery; ram pressure brings the
10
Barnes
gas polytropes to a screeching halt on their first and only passage. A similar
explanation probably applies in this deeply-penetrating off-axis collision; the
small polytrope does so much work plowing through the body of its large partner
that it loses most of its orbital momentum.
The puzzle of the grazing collisions can't be explained in this way as the
gas polytropes barely touch each other. These decays are governed by tides,
and it may seem paradoxical that the gas and grit models with rp = 1.5 di-
verge since the same tidal gradients exist in both cases. However, the bottom
rows of Figs. 3 and 4 show that gas polytropes suffer stronger tidal distortions
in grazing collisions than do their grit counterparts. This make sense; in the
initial polytropes, gas particles are nearly stationary, while grit particles are in
constant motion. Thus during a tidal encounter, a well-placed gas particle can
accumulate more momentum than a wandering grit particle. Now the stronger
tidal response of the gas polytropes is the key; recall that there would be no
orbit decay whatsoever if both participants remained exactly spherical. By be-
ing more responsive, gas polytropes couple orbital motion to internal degrees of
freedom more effectively, and hence in grazing collisions their orbits decay faster
than those of grit polytropes.
2.3. Later passages
Thanks to orbit decay, later passages are slower and closer than first ones. Closer
passages are always more violent, as Figs. 3 and 4 amply illustrate. On the other
hand, the speed of passage affects gas and grit models in different ways. For grit
models, slower passages are generally more disruptive; galaxies collide in exactly
the way that cars don't, because low speeds give tides more time to act. Gas
models, though also sensitive to tides, suffer less from shocks in slow collisions;
the coalescence of two gas polytropes is more like a gentle swirling together than
a head-long plunge.
Every passage pumps orbital energy into internal degrees of freedom, dis-
tending both gas and grit polytropes. After each passages, gas systems relax
into convectively stable configurations; low-entropy material remains centrally
concentrated, while high-entropy material forms an extended envelope. Grit sys-
tems also acquire distended outer envelopes; eq. 5 conserves phase-space density,
so in phase-space such an envelope is really a thin "hyper-ribbon", tidally ex-
tracted from the main body, which still has the high phase-space density of
its source. Nonetheless, this ribbon comes to resemble a smooth envelope as
phase-mixing winds it into a tight spiral.
In basic outline, final encounters resemble mild versions of head-on colli-
sions; either both partners are subsumed into a single object, or one partner
burrows into the center of the other.
3. Remnant Structure
In the aftermath of a merger the wreckage undergoes a last episode of dynamical
relaxation as it evolves towards an equilibrium configuration. This final convul-
sion generally begins at the center and travels outward. In a grit system po-
tential fluctuations during the final merger scatter particles onto loosely-bound
radial orbits; as these outgoing particles reach apocenter they create a caustic or
Dynamics of Stellar Collisions
11
Figure 6.
Density profiles derived by spherical averaging. Solid lines are
gas results, while dashed lines are grit results. Left: remnants of equal-mass
collisions. Right: remnants of unequal-mass collisions; thin lines show profiles
for particles from the small polytropes. Top to bottom: remnants of collisions
with rp = 0, 0.5, and 1, displaced downward by successive factors of 100 for
clarity.
"shell". In a gas system the final merger may create a weak outgoing shock, but
much of the readjustment takes place at subsonic velocities, with no attendant
increase in entropy.
3.1. Density profiles
Fig. 6 presents density profiles for all six merging collisions (those with rp ≤ 1).
These profiles were calculated by evaluating the mass in nested spherical shells.
Despite the differing dynamics of gas and grit mergers, the resulting remnants
have similar profiles over a range of five decades in density. At large radii both
grit and gas densities fall off roughly as ρ ∝ r−4. For grit models this asymptotic
slope results as phase-mixing spreads out a population of particles with a wide
range of energies including both bound and unbound orbits (Jaffe 1987; White
1987).
The unequal-mass mergers yield density profiles with "shoulders" where the
slope briefly becomes shallower as r increases. These compound profiles arise
because the small polytropes resist disruption during the merger process. The
thin lines in the right-hand panel of Fig. 6 show density profiles derived using
only particles originating in from the small polytropes. In every case, the small
12
Barnes
Figure 7.
Initial vs. final binding energies for particles from remnants of
(M1/M2, rp) = (1, 0.0) collision. Left: gas model; right: grit model. Note
that the two plots have different scales. In both plots, E = 0 is escape energy.
polytropes have settled, virtually intact, in the centers of the merger remnants.
For the gas models, this outcome is explained by the lower entropy which the
small polytrope retains throughout the collision and merger; the remnant can't
be convectively stable unless this low-entropy material winds up at the center.
If both partners have the same entropy, how much rearrangement takes
place during a merger? Fig. 7 presents scatter-plots of initial binding energy Ei
vs. final binding energy Ef for particles in both versions of the (M1/M2, rp) =
(1, 0.0) collision. Here Ei was measured with the polytropes at infinite separa-
tion, while Ef was measured from the merger remnant; internal energies were
included in calculating binding energies of gas particles. These plots show that
initial binding energy is a useful, albeit imperfect, predictor of final binding
energy. For the gas model, the correlation between Ei and Ef is fairly strong;
moreover, there is a one-to-one relationship between binding energy and radius,
so the initial and final radii of gas particles are also correlated.
For the grit model things are more complex. As the right-hand panel of
Fig. 7 shows, grit particles have a bimodal distribution of final binding energies.
This arises because the orbital phases of individual grit particles have a strong
bearing on their response to tidal interactions. For example, consider a particle
which reaches an apocenter of its orbit just before the polytropes interpenetrate,
and the subsequent pericenter just after they separate. Such a particle sees a
deeper potential well while falling in than it does while climbing out, so it gains
energy; conversely, a particle on a similar orbit with a different phase may lose
energy. The resulting bimodal energy distribution explains the shoulder in the
density profile of this remnant (Fig. 6, left-hand panel, upper dashed curve).
Tightly-bound particles form the core, while the rest form the ρ ∝ r−4 envelope;
the deficit of particles with Ef ≃ −1.2 accounts for the dip in ρ(r) at r ≃ 0.8.
Dynamics of Stellar Collisions
13
Figure 8. Merger remnants. Contours indicate densities in steps of a factor
of 4; solid lines show gas, while dashed lines show grit. Each remnant is
shown twice, once on a slice through the orbital plane (above), and once on
a perpendicular slice (below).
3.2. Shapes
Fig. 8 presents density contours for the full sample of gas and grit merger rem-
nants. Two slices through each remnant are shown; one slice in the orbital
plane, and one slice at right angles to this plane. As in Figs. 3 and 4, here too
the head-on collisions produce the most striking discrepancies between gas and
grit models. The head-on gas remnants are nearly spherical, as expected for a
self-gravitating gas configuration with no angular momentum. The head-on grit
remnants, on the other hand, are prolate bars aligned with the collision axis.
This is a "long-term" memory of the original collision velocities. Without the
14
Barnes
Figure 9.
Rotation velocities vφ vs. cylindrical radii R for remnants of
(M1/M2, rp) = (1, 1.0) collision. Left: gas model; middle: gas atoms; right:
grit model.
short-range interactions which randomize gas motions, grit systems maintain
anisotropic velocity dispersions which can support such aspherical shapes.
The remnants produced in off-axis collisions have significant amounts of
angular momentum. Most such collisions yield fairly oblate remnants; sliced
perpendicular to the orbital plane, the gas remnants are sometimes "disky"
while the grit remnants seem more "boxy", but in general the density contours
are very similar. The remnants formed by the (M1/M2, rp) = (1, 0.5) collisions
are somewhat unusual in this regard. Both are distinctly non-axisymmetric;
sliced along the orbital plane, the grit model has elliptical density contours at
all radii, while the gas model seems to have a doubled core and a slightly elon-
gated envelope. The grit remnant may be stable, since anisotropy and rotation
together can support a tumbling triaxial structure in equilibrium. The same
may not be true of the non-axisymmetric gas remnant; more experiments could
help settle this issue.
3.3. Kinematics
Fig. 9 compares the kinematics of gas and grit versions of remnants from collision
(M1/M2, rp) = (1, 1.0). This remnant has a relatively large amount of angular
momentum and a major-to-minor axis ratio of roughly 2:1. Here the velocity in
the direction of rotation, vφ, has been plotted against cylindrical radius in the
rotation plane, R. Between the plots of particle velocities for gas (left) and grit
(right), an additional plot shows velocities of gas atoms taking thermal motion
into account (middle). The gas has a very regular velocity field, as reported
in previous studies (eg. Lombardi, Rasio, & Shapiro 1996).
In contrast, the
grit has a broad distribution of velocities at each point; while the overall sense
of rotation is the same as in the gas case, some grit particles even counter-
rotate. But when the grit is compared to the gas atoms, it's evident that these
velocity distributions are very much alike. The similarity of these distributions
is consistent with the good match between the density contours of the gas and
grit versions of this remnant (Fig. 8, upper-right).
Kinematics for the remnants of the head-on collision (M1/M2, rp) = (1, 0.0)
are presented in Fig. 10. Here the radial velocity, vr, has been plotted against the
Dynamics of Stellar Collisions
15
Figure 10.
(1, 0.0) collision. Left: gas model; middle: gas atoms; right: grit model.
Radial velocities vr vs. radii r for remnants of (M1/M2, rp) =
spherical radius, r. This remnant has no angular momentum, so gas velocities
should vanish; the small velocities seen in the left-hand panel indicate that the
remnant is not yet fully relaxed. The grit version also is not yet completely
relaxed, since in perfect equilibrium the distribution should be symmetric with
respect to the line vr = 0. Moreover, the grit plot shows that in phase-space
the envelope of this remnant is not smooth, but consists of narrow ribbons of
particles. These particles populate the upper peak in the final binding energy
distribution for this remnant (Fig. 7, right-hand panel); launched as the two
polytropes interpenetrated, they fell back in coherent streams and have since
been wound up by ongoing phase-mixing. With time this fine structure will
become harder to see, and eventually this plot will look something like the plot
of atomic velocities shown in the middle. But even in the limit t → ∞ some
differences will remain; these include the radial anisotropy of the grit velocity
distribution and bimodal nature of its binding energy distribution.
4. Conclusions
Comparison of gas-dynamical and stellar-dynamical models of colliding systems
reveals many specific differences and hints at some underlying similarities. While
stellar systems interact via gravity alone, gas systems interact via a complex
mixture of gravity, pressure forces, and shocks. This mix greatly increases the
range of dynamical behavior in stellar collisions. Even when interactions between
gas systems are limited to tides, the resulting deformations are stronger than
those seen in equivalent encounters of stellar systems.
Remnants of gas-dynamical and stellar-dynamical mergers also exhibit sig-
nificant differences. Stellar systems have anisotropic velocity distributions, sup-
porting a wide range of remnant morphologies, while gas systems settle into a
small range of equilibria. But at the same time, the stellar and gaseous remnants
studied here have very similar density profiles, and often similar shapes as well.
This seems remarkable, since the physics behind these density profiles is rather
different. The extended envelopes of gaseous remnants contain high-entropy ma-
terial produced by shocks, which are violent, well-localized events resulting from
physical encounters. On the other hand, the envelopes of stellar remnants con-
16
Barnes
tain material extracted from the participants by tides; this material retains the
high fine-grain phase-space density it starts with, and only when coarse-grained
does it approximate the atomic velocity distribution of a gaseous envelope. Two
puzzles remain: first, why do such disparate mechanisms -- shocks and tides --
produce such similar remnant envelopes? And second, how is coarse-graining
-- an operation applied to a stellar system by an observer -- analogous to the
irreversible physical processes which occur in gaseous systems?
Acknowledgments.
I thank Piet Hut, James Lombardi, Frederic Rasio,
and Stuart Shapiro for helpful suggestions. Support for this work was provided
by NASA grant NAG 5-8393 and by Space Telescope Science Institute grant
GO-06430.03-95A.
References
Antonov, V. A. 1962, Vestn. Leningrad Univ., 19, 96 [translated in de Zeeuw, T., ed.
1987, Structure and Dynamics of Elliptical Galaxies (Dordrecht: Reidel), p. 531]
Barnes, J. E. 1998, Galaxies: Interactions and Induced Star Formation, eds. D. Friedli,
L. Martinet, D. Pfenniger (Berlin: Springer-Verlag), 275
Barnes, J., Goodman, J., Hut, P. 1986, ApJ, 300, 112
Chandrasekhar, S. 1939, An Introduction to the Study of Stellar Structure (Chicago:
Univ. of Chicago Press)
Eddington, A. S. 1916, MNRAS, 76, 572
H´enon, M. 1973, A&A, 24, 229
Hernquist, L. 1993, ApJ, 404, 717
Hernquist, L. & Katz, N. 1989, ApJS, 70, 419
Jaffe, W. 1987, Structure and Dynamics of Elliptical Galaxies, ed. T. de Zeeuw (Dor-
drecht: Reidel), p. 511
Kalnajs, A. 1972, ApJ, 175, 63
Lacey, C. G., Ostriker, J. P. 1985, ApJ, 299, 633
Lombardi, J. C., Rasio, F. A., Shapiro, S. L. 1996, ApJ, 468, 797
Lynden-Bell, D. 1962, MNRAS, 124, 279
Monaghan, J. J. 1992, ARA&A, 30, 543
Monaghan, J. J., Gingold, R. A. 1983, J. Comp. Phys., 52, 374
Nelson, R. P., Papaloizou, J. C. B. 1993, MNRAS, 265, 905
Nelson, R. P., Papaloizou, J. C. B. 1994, MNRAS, 270, 1
Sills, A., Lombardi, J. C. 1997, ApJ, 484, L51
Toomre, A. 1964, ApJ, 139, 1217
van Albada, T. S., van Gorkom, J. H. 1977, A&A, 54, 121
White, S. D. M. 1987, Structure and Dynamics of Elliptical Galaxies, ed. T. de Zeeuw
(Dordrecht: Reidel), p. 339
|
0709.1216 | 1 | 0709 | 2007-09-08T15:10:35 | The Persian-Toledan Astronomical Connection and the European Renaissance | [
"astro-ph"
] | This paper aims at presenting a brief overview of astronomical exchanges between the Eastern and Western parts of the Islamic world from the 8th to 14th century. These cultural interactions were in fact vaster involving Persian, Indian, Greek, and Chinese traditions. I will particularly focus on some interesting relations between the Persian astronomical heritage and the Andalusian (Spanish) achievements in that period. After a brief introduction dealing mainly with a couple of terminological remarks, I will present a glimpse of the historical context in which Muslim science developed. In Section 3, the origins of Muslim astronomy will be briefly examined. Section 4 will be concerned with Khwarizmi, the Persian astronomer/mathematician who wrote the first major astronomical work in the Muslim world. His influence on later Andalusian astronomy will be looked into in Section 5. Andalusian astronomy flourished in the 11th century, as will be studied in Section 6. Among its major achievements were the Toledan Tables and the Alfonsine Tables, which will be presented in Section 7. The Tables had a major position in European astronomy until the advent of Copernicus in the 16th century. Since Ptolemy's models were not satisfactory, Muslim astronomers tried to improve them, as we will see in Section 8. This Section also shows how Andalusian astronomers took part in this effort, which was necessary in the path to the Scientific Revolution. Finally, Section 9 presents the Spanish influence on the eve of the Renaissance. | astro-ph | astro-ph | Academia Europaea 19th Annual Conference
in cooperation with:
Sociedad Estatal de Conmemoraciones Culturales,
Ministerio de Cultura (Spain)
“The Dialogue of Three Cultures and our European Heritage”
(Toledo Crucible of the Culture and the Dawn of the Renaissance)
2 - 5 September 2007, Toledo, Spain
Chair, Organizing Committee: Prof. Manuel G. Velarde
The Persian-Toledan Astronomical Connection and the European Renaissance
M. Heydari-Malayeri
Paris Observatory
Summary
This paper aims at presenting a brief overview of astronomical exchanges between the
Eastern and Western parts of the Islamic world from the 8th to 14th century. These cultural
interactions were in fact vaster involving Persian, Indian, Greek, and Chinese traditions. I will
particularly focus on some interesting relations between the Persian astronomical heritage and
the Andalusian (Spanish) achievements in that period. After a brief introduction dealing
mainly with a couple of terminological remarks, I will present a glimpse of the historical
context in which Muslim science developed. In Section 3, the origins of Muslim astronomy
will be briefly examined. Section 4 will be concerned with Khwârizmi, the Persian
astronomer/mathematician who wrote the first major astronomical work in the Muslim world.
His influence on later Andalusian astronomy will be looked into in Section 5. Andalusian
astronomy flourished in the 11th century, as will be studied in Section 6. Among its major
achievements were the Toledan Tables and the Alfonsine Tables, which will be presented in
Section 7. The Tables had a major position in European astronomy until the advent of
Copernicus in the 16th century. Since Ptolemy’s models were not satisfactory, Muslim
astronomers tried to improve them, as we will see in Section 8. This Section also shows how
Andalusian astronomers took part in this effort, which was necessary in the path to the
Scientific Revolution. Finally, Section 9 presents the Spanish influence on the eve of the
Renaissance.
1. Introduction
Before dealing with the main topics of this paper, it seems necessary to comment first on
three widely used terms in this field: Arab/Arabic astronomy, Islamic astronomy, and zij.
Arab/Arabic is not meant as an ethnic but rather a linguistic term. In fact a large
number of Non-Arab scholars, mainly Persians, Turks, and Spanish people, wrote their works
in Arabic. Even so, many astronomical works were also produced in other languages of this
civilization, especially Persian and in the later centuries Turkish. For example, the main zijs
were originally written in Persian, a notable example being the Ulugh Beg’s (c. A.D. 1394-
1449) zij, a landmark in precise observations before the Renaissance. We also note a disparity
with respect to Western scholars who wrote in Latin. As far as these scholars are concerned,
the Latin adjective is not specified (e.g., the expressions like “the Latin astronomer
Copernicus”, “the Latin physicist Newton”, or “the Latin philosopher Leibnitz” are not used).
As for the term Islamic, it should be taken in the sense of the civilization rather than
the religion, because much of the astronomy was secular. Moreover, many non-Muslims
within the Islamic civilization contributed to this science and must be acknowledged. Once
again, we find the above-mentioned disparity, since the term Christian, which refers also to a
civilization, is not used either (e.g. Galileo and Newton are not usually referred to as
“Christian scientists”).
Zij is the generic name applied to books in Arabic and Persian that tabulate parameters
used for astronomical calculations of positions of the Sun, the Moon, and the five planets of
antiquity. The word is derived from Middle Persian zig, variant zih, meaning “cord, string”
(Modern Persian zeh “cord, string”), from Avestan jiiā- “bow-string”, cognate with Sanskrit
jiyā- “bow-string”, Proto-Indo-European base *gwhi- “thread, tendon” (from which derive
also Greek bios “bow”, Latin filum “thread”, Russian žica “thread”). The term zig originally
referred to the threads in weaving, but because of the similarity between the rows and
columns of astronomical tables and the parallel threads, it came to be used for an
astronomical table, and subsequently a set of tables.
2. A glimpse of the historical background
Cultural developments in the course of history are not detached from underlying
social/political events. In order to better understand the advent of Islamic science, it would be
interesting to have a fast glimpse of the historical background.
The first Islamic state, established by the Umayyad dynasty (661-750), lasted for 89
years. There were social turmoil in various parts of the vast conquered territories and in
particular Iranian resistance movements opposed the Arab domination, especially since the
Umayyads considered non-Arabs as mawali, people of lowly status.
The Umayyads were overthrown by the leader of a revolutionary movement Abu
Moslem Khorâsâni (Persian name Behzâdân), who enthroned Abu al-Abbâs as-Saffâh, a
member of the prophet’s lineage. This was the beginning of the Abbasid dynasty (750-1258),
which lasted over five centuries, although Saffâh himself reigned for only four years.
The only Umayyad survivor, Abd ar-Rahmân I, escaped to Andalusia where he
established himself as an independent Emir (756-788). More than a century later one of his
successors, Abd ar-Rahmân III (912-961), assumed the title of Caliph, establishing Cordoba
as a rival to the Abbasid capital.
The Abbasid era was substantially influenced by the Persian culture and tradition of
government. A new position, that of vizier (Arabic from Middle Persian vicīr “decree,
decision”, from vicīrītan “to decide, to judge, to distinguish”; vicīrtār “judge, arbitrator,
decider”), was created and many Abbasid caliphs were eventually relegated to a more
ceremonial role than under the Umayyads. Several prominent viziers, serving Hârun ar-
2
Rashid (786-809) and his son al-Ma’mun (813-833), were members of the Persian Barmakid
family of Buddhist faith. Moreover, al-Ma’mun’s mother was Iranian, and he himself had
grown up in Khorâsân, the Eastern province of Iran.
During the second caliph al-Mansur (754-775) the capital was moved from Damascus
to Baghdad, not far from Ctesiphon, the ancient capital of Iranian Sassanids. The designers
hired by al-Mansur to lay the city’s plan were two Persians: Naubakht, a former Zoroastrian,
and Mâshâ’allah, a Jew from Khorâsân. The two men also determined an astrologically
auspicious date for the foundation of the city: 30 July 762.
It is also notable that the city name Baghdad is Persian, meaning “god-given” or
“God’s gift”, from bagh “god, lord” + dâd “given”. The first component derives from Old
Persian baga-, Avestan baγa- “lord, divider” (from bag- “to allot, share”), cognate with
Sanskrit bhága- “part, portion”, Proto-Indo-European base *bhag- “to divide”; cf.
Slavic/Russian bog “god”, Greek phagein “to eat” (originally “to have a share of food”). The
second component dâd, from dâdan “to give”, Old Persian/Avestan dā- “to give, grant”,
Proto-Indo-European base *do- “to give”; cf. Sanskrit dadáti “he gives”, Greek didonai
“to give”, Latin datum “given”, Russian dat’ “to give”.
The reign of Hârun ar-Rashid and his successors fostered an age of intellectual
activities. A cultural center, the House of Wisdom (Bait al-Hikmah), was set up, which was
reminiscent of the Persian Sassanid academy of Gundishapur. An intense translation activity
was undertaken and all sorts of books were translated from Middle Persian, Sanskrit, Syriac,
and Greek into Arabic. In just a few decades the major scientific works of antiquity, including
those of Galen, Aristotle, Euclid, Ptolemy, Archemides, and Apollonius, were translated into
Arabic. The main translators were Hunayn ibn Ishâq (c. 809-873), a Nestorian Christian with
an excellent command of Greek, and Thâbit ibn Qurra (c. 836-901), a Helenized pagan from
Harran, a town in northern Mesopotamia (today Turkey), that was the center of a cult of “star
worshippers”.
The paper needed for books was abundant, since Muslims had learned the techniques
of paper making from the Chinese. In fact the Chinese prisoners of war captured in the battle
of Talas (731) were ordered to produce paper in Samarkand and by the year 794 a paper mill
was installed in Baghdad.
3. Origins of Muslim astronomy
At the advent of Islam Arabs did not have an elaborated/documented astronomy. The first
astronomical documents translated into Arabic were of Indian and Persian origins. These
translations set the basis for the first Muslim astronomical works. Greek astronomy,
represented by Ptolemy, was introduced later, but it gained a fully dominant position owing to
its predictive capacity. Here are the main founding sources:
The Persian work was Zij-e shâh or Zij-e shahryâr, originally composed for the
Sassanid emperor Khosrow I (Anushirwân) about the year 550. It was translated into Arabic
by Abu al-Hasan al-Tamimi and commented on by Abu Ma’shar (Albumasar) of Balkh. It
3
became the basis for example for the work of the previously mentioned astronomers
Naubakht and Masha’allah. Zij-e shâh contained some elements of Indian and Greek
traditions. It also had its specific Persian particularities, mainly the basic year for the tables,
which was the coronation date of the last Sassanid emperor Yazdegerd III (16 June 632), and
the Solar year based on Nowruz, or spring equinox. The Yazdegerd III’s era was in use in
Muslim astronomy during several centuries, before being replaced with the Hijra. As an
interesting particularity of the Zij-e Shâh, the day started from midnight.
As to the Indian sources, several works have been mentioned in early Muslim
astronomy, the main one Siddhānta (Sanskrit meaning “established end, final aim, doctrine,
concept”) attributed to Brahmagupta (598-670). This work was brought to Baghdad sometime
around 770 by an Indian political delegation, which had an astronomer named Kanka. The
book was translated into Arabic by al-Fazâri and Ya’qub ibn Târiq, who were assisted by the
Indian astronomer. The Sanskrit term was later corrupted to Sindhind in Arabic.
The Greek source was Ptolemy’s Almagest, dating from about A.D. 150. The
Ptolemy’s Mathematike Syntaxis “Mathematical Compilation”, in later antiquity known
informally as Megale Syntaxis or Megiste Syntaxis “The Great Compilation”, was translated
from Greek into Arabic in the 9th century during the translation campaign launched by al-
Ma’mun. It was translated several times, during which the title word Megiste was transformed
into al-majisti. The earliest translators were the above-mentioned Hunayn ibn Ishâq and
Thâbit ibn Qurra. By this time the work was lost in Europe.
The Persian astronomer Ahmad Farghâni (Alfraganus) presented a thorough
descriptive summary of Almagest in his textbook Jawâmi’, known as Elements, written
between 833 and 857. It was translated into Latin in the 12th century and was widely studied
in Europe until the time of Regiomontanus (1436-1476). Farghâni also composed a very
important treatise on the astrolabe around 856. Although the astrolabe was a Greek invention,
the earliest dated instrument that has been preserved comes from the Islamic period. The only
extant ancient treatise on the astrolabe is due to Johannes Philoponos, written in the first part
of the 6th century.
The first Muslim astronomer who based his work principally on Ptolemy was al-
Battâni (c. 853-929, born in Harran), who made his observations at al-Raqqa in Syria.
Ptolemy’s work re-entered Europe from its Arabic versions with the transformed name
Almagest in the medieval Latin translations.
4. Khwârizmi’s zij
Zij of Sindhind, written about 820 by Khwârizmi, was the first major astronomical work in the
Muslim world. It was mainly based on Indian/Persian astronomy. Interestingly, Khwârizmi’s
zij had its greatest long-term influence in Muslim Spain and Western Europe through the
incorporation of some of its material in the Toledan Tables.
Abu Ja’far Muhammad ibn Musâ Khwârizmi (c. 780- c. 850) was a key figure in the
history of algebra, an astronomer and geographer. The epithet Khwârizmi indicates that he or
4
his forebears came from the Persian region of Khwârizm, the present-day Khiva in
Uzbakistan. The historian Tabari (c. 838-923) gives him the additional epithet Majusi (related
to magus), meaning Zoroastrian. This would have been possible at that time for a man of
Persian origin. However, the preface of his Algebra (if effectively written by himself) shows
him a pious Muslim. Anyhow, Tabari’s designation could mean also that his ancestors, and
perhaps himself in his youth, had been Zoroastrian.
Zij of Sindhind was in particular based on the Iranian solar year with the starting era
that of Yazdegerd III, as previously indicated. The Sun, the Moon, and each of the five
planets known in antiquity had a table of mean motion and a table of equations (variations
with respect to mean values). In addition, there were tables for computing eclipses, solar
declination and right ascension, and various trigonometric tables. The form of a set of tables
closely resembled that made standard by Ptolemy. But most of the basic parameters in the Zij
(the mean motions, the mean positions at epoch, positions of apogee and the node) were
derived from Indian astronomy. The maximum equations were taken from Zij of Shâh. The
fundamental meridian was that of Arin, lying 70° east of Baghdad. Arin was a corruption of
Ujjayni (present-day Ujjain), a city situated in central India, which was the “Greenwich” of
the ancient Indian astronomy.
The original Arabic version is lost. A Latin translation exists carried out by the
English scholar Adelard of Bath (c. 1080-1152) in the early 12th century. This translation was
made not from the original, but from a revision executed by a Spanish astronomer, al-Majriti
(c. 950- c. 1007).
The Zij continued to be used in classrooms and commented on even after al-Battâni,
the aforementioned Mesopotamian astronomer produced his great work (Zij al-Sâbi), based
principally on the Almagest and his own observations. Battâni is considered as the first
Muslim astronomer to carry out new, systematic observations since the time of Ptolemy.
Khwârizmi is also recognized for his book on algebra: al-kitâb al-mukhtasar fi hisâb
al-jabr w’al-muqâbala “the Compendious Book on Calculation by Completion and
Balancing”. The term algebra in the European languages derives from al-jabr “completion,
restoration” used by Khwârizmi. It should be underlined that at the early days of Algebra
mathematical symbols were not used and one had to resort to phrases instead.
He is also remembered for being at the origin of the transfer of Hindu numerals
(commonly called Arabic numerals) to Europe. Likewise, the term algorithm derives from his
name, from French algorithme, refashioned under mistaken connection with Greek arithmos
“number”, from Old French algorisme, from Middle Latin algorismus “the Arabic system of
arithmetical notation”, a distorted transliteration of al-Khwârizmi.
Judging from what has remained from Khwârizmi, his importance is mainly due to the
fact that he was the first to scholarly introduce the Indian/Persian traditions in
astronomy/mathematics into the Muslim world.
5. Khwârizmi’s influence on Andalusian astronomy
Abd ar-Rahman II, the Umayyad Emir of Cordoba (822-852), was an amateur of books on
philosophy, medicine, and music. He also loved astrology and his trustworthy astrologer
5
occupied a high rank in the court. It is possible that his liking for astrology stemmed from two
notable astronomical events. The almost total solar eclipse of 17 September 833 drove the
striken mob to the Cordoba’s grand mosque to carry out the ritual fright prayer. The other
event was an intense meteor shower between 2 April and 18 May 839. It was during Abd ar-
Rahman II’s reign that a version of Khwârizmi’s zij was introduced.
The real development of the Andalusian science took place in the second half of the
10th century. The first prominent Andalusian astronomer was Maslama ibn Ahmad al-Majriti
(c. 950- c. 1007) of Madrid. He is particularly renowned for his version of Khwârizmi’s zij, in
which he changed the chronology from the Persian epoch (as explained before) to the Muslim
Hijri calendar. Further, Majriti transferred the standard meridian from Arin (explained earlier)
to Cordoba. Majriti’s version of the Zij was translated into Latin by Adelard of Bath and was
commented upon by Ibn al-Muthannâ, a work being extant in a Hebrew version. It seems
moreover that Majriti’s work improved the calculation methods of Khwârizmi.
Although Andalusian astronomers knew Ptolemy’s work at least from the 10th
century, they never fully abandoned the Sindhind tradition. Even about a century after Majriti,
the Zij of Jayyân, set up by Ibn Mu’âdh al-Jayyani (d. 1093), was an adaptation of
Khwârizmi’s zij to the coordinates of the city of Jaén in south-central Spain.
6. Flourishing of Andalusian astronomy
Andalusian astronomy flourished in the 11th century with Abu Ishâq Ibrâhim an-Naqqâsh (d.
1100), surnamed az-Zarqâli (Spanish transcription Azarquiel), from zarqâ’ (“blue”; “the blue-
eyed one”). He was the first in the history of astronomy to unambiguously state the proper
motion of the solar apogee with respect to stars and distinguish it from the precession of the
equinoxes. The solar apogee was the farthest distance of the Sun from the Earth, when the
Sun had its smallest apparent size. According to Zarqâli’s observations, the motion of the
solar apogee amounted to 12″.04 per year. Resulting from 25 years of solar observations, first
at Toledo and then at Cordoba, this measurement is highly remarkable compared to modern
observations, which yield a displacement of 11″.6 annually. Today we know that the motion
of the solar apogee results from the rotation of the line of apsides, or major axis, of the
Earth’s orbit due to gravitational perturbation by other planets.
Zarqâli’s great manual skill allowed him to construct precision instruments for
astronomical use. Perhaps inspired by the Toledan astronomer Ali ibn Khalaf, he constructed
a “universal astrolabe” that could be used at any latitude. This innovation by Andalusian
astronomers removed the inconvenience of having to change the plate (safiha) for each
latitude. He also built the water clocks of Toledo, which were capable of determining the
hours of the day and night and indicating the days of the lunar month. The clocks were in use
until 1133, when Hamis ibn Zabara (during Alfonso VII) tried to discover how they worked;
they were dismounted but could not be reassembled.
Zarqâli’s results had far-reaching consequences. Up till now Andalusian astronomy
was subject to the pre-eminence of Eastern astronomy. Zaqâli starts a new period in which the
Andalusian astronomers confirm the excellence of their research. The achievements of the
11th century Andalusian astronomers were transmitted and praised in the Middle East,
announcing the beginning of a historical shift in the geographical focus.
6
It would be interesting to recall other notable features of Zarqâli’s epoch in a broader
scope. Zarqâli was contemporary with the Persian mathematician, astronomer, and poet Omar
Khayyâm (1048-1131). He was the first in the history of mathematics to undertake a
systematic study and classifiaction of equations of degree ≤ 3 and to elaborate a geometrical
solution for them. He is also known for his outstanding reform of the Iranian solar calendar,
and his poems (rubaiyat). Interestingly, when Zarqâli was 25 years old one of the rarest
astronomical events happened: the explosion of the famous supernova. On July 4, 1054,
Chinese astronomers noted a "guest star" in the constellation Taurus. This star became about
four times brighter than Venus in its brightest light, or about mag –6; it was visible in daylight
for 23 days and in the night sky for 653 days. It is unlikely that Muslim astronomers did not
see this phenomenon, but the fact that they did not care about this extraordinary event
suggests that they were exclusively concerned with planetary motions.
7. Toledan Tables, Alfonsine Tables
Zarqâli is above all renowned for his contribution to the Toledan Tables which were probably
compiled after 1068. This work, which represents the first original development of
Andalusian astronomy, was extremely influential in Europe for three centuries until the
advent of the Alfonsine Tables. It was adapted to local meridians almost all-over Europe (e.g.
Pisa, London, Toulouse, Paris, Marseille) until the 13th century. The fact that a large number
of copies were made of the Toledan Tables in the 14th and even 15th century implies that
they were in use even after the Alfonsine Tables were introduced.
The main sources for the bulk of the table collections were those of Khwârizmi
(mainly planetary latitudes), Battâni (planetary equations), and Ptolemy. In fact the oldest
version of the Toledan Tables was mainly modeled on Khwârizmi’s Sindhind, but had
admixture from Battâni. In addition, the oldest versions of the Toledan Tables preserve some
tables of Khwârizmi that are rare or absent elsewhere.
The Toledan Tables also incorporated the thory of trepidation attributed to Thâbit ibn
Qurra (c. 830-901). Trepidation was a spurious oscillation of the equinoxes thought to have a
period of about 7000 years. In order to explain trepidation, Thâbit was said to have added a
new sphere to the eight Ptolemaic spheres beyond the sphere of fixed stars. However, Al-
Battâni rejected trepidation. In fact trepidation dominated the medieval astronomy. The
original Arabic version of the Toledan Tables has been lost, but two Latin versions have
survived, one by Gerard of Cremona (12th century) and one by an unknown author.
The Alfonsine Tables were drawn up in Toledo to correct the anomalies in the
Toledan Tables, which they rapidly superceded sometime after 1320. Alfonso X, el sabio,
king of Castile and Léon from 1252 to 1284, encouraged his astronomers and translators to
prepare a major corpus of astronomical text in Castilian. The starting point of the Alfonsine
Tables is January 1, 1252, the year of king’s coronation (1 June). The original Spanish
version of the tables is lost, but a set of canons (introductory instructions) for planetary tables
are extant. They are written by Isaac ben Sid and Judah ben Moses ha-Cohen, two of the most
active collaborators of Alfonso X. The Alfonsine Tables were the most widely used
astronomical tables in the Middle Ages and had an enormous impact on the development of
European astronomy from the 13th to 16th century. They were replaced by Erasmus
Reinhold’s Prutenic Tables, based on Copernican models, that were first published in 1551.
7
The Latin version of the Alfonsine Tables first appeared in Paris around 1320, where a
revision was undertaken by John of Lignères and John of Murs, accompanied by a number of
canons for their use written by John of Saxony. There is a controversy as to the exact
relationship of these tables with the work commissioned by the Spanish king. Surprisingly,
the earliest evidence of their use in Spain dates back to about 1460 in Salamanca when a
Polish astronomer, Nicholaus Polonius, arrived and brought them with him. It is quite
possible that the astronomers of the early 14th century working in Paris did a fine job in
adapting previous astronomical material. However, if they created something quite new, why
did they keep calling it Alfonsine Tables? The fact that the authors of the Parisian tables did
not choose a new title for their work suggests that they did not consider it to be fundamentally
different from the original Alfonsine Tables. In 1483 the Parisian Alfonsine Tables were
printed in Venice for the first time, followed shortly afterwards by other editions, all of them
by printers in that city.
8. Andalusian criticism of Ptolemy
Ptolemy’s Almagest, with its 13 chapters, dominated medieval astronomy. Notwithstanding,
Ptolemy’s model was an inadequate representation of planetary motions. As an extreme
example, according to Ptolemy’s model for the Moon, our satellite should appear to be almost
twice as large when it is full than it is at quadrature, which is an absurdity since it is not seen
as such.
The Muslim scholars did not follow Ptolemy blindly but disagreed on a number of
points. For instance, the Egyptian Ibn al-Haytham (965-1040), the greatest authority on optics
in the Middle Ages, wrote a treatise, Al-shukuk ‘alâ Batlamyus (Doubts about Ptolemy),
criticizing the Ptolemaic system for its complexity and incompatibility with the Aristotelian
physics. Similarly, the Persian polymath Abu Rayhân Biruni (973-1048), author of Qânun al-
Mas’udi, considered as the most important Muslim astronomical text, rejected Ptolemy’s
opinion about the immobility of the solar apogee on the basis of many astronomical
observations. Mostly, the criticisms of Ptolemy were mainly philosophically motivated.
Aristotle was a great authority who dominated every field of knowledge, including physics,
whereas Ptolemy did not have such a standing. The objections had three main reasons:
1) In the Aristotelian cosmology the Earth was situated at the center of the Universe. In
contrast, for Ptolemy the Earth did not occupy the central place. The main sphere (orb), called
deferent, was centered on a point halfway between the Earth and another point invented by
Ptolemy himself, termed equant. This point was meant to bring Ptolemy’s planetary theory
closer to observations. The epicycle rotated on the deferent with uniform motion, not with
respect to the Earth, nor around the center of the deferent, but with respect to the off-center
point equant. However, this does not work since one might think in terms of a sphere moving
at a uniform speed around an axis that does not pass through its center.
2) Aristotle believed that circular motion with uniform speed was the natural property of
celestial objects. The equant point introduced by Ptolemy caused the epicycle center to move
with variable speed on the deferent circle.
8
3) The Ptolemaic epicycles and deferents were abstract, geometrical concepts, while in the
Aristotelian cosmology the spheres were physical entities to which celestial bodies were
attached.
In Andalusia, after Zarqâli the Ptolemaic model was much debated and its criticism
gained momentum, mainly by philosophers. Ibn Rushd (Averroes, 1126-1198) and Ibn Bâjja
observed spots on the solar disk which they ascribed to the transits of Mercury and Venus.
The idea of spots on the Sun was not admissible since Aristotle’s doctrine maintained that the
heavens were incorruptible. Anyhow, putting forward the transit argument meant a criticism
of Ptolemy, since he had argued that the lines of sight joining these planets to Earth never
cross the Sun. These transits were a matter of debate among Andalusian thinkers. Further,
Averroes rejected Ptolemy’s eccentric deferents and argued for a strictly concentric model of
the Universe.
The 12th century Andalusian astronomers were divided between those who followed
Zarqâli’s tradition, like Abu as-Salt, Ibn al-Kammâd (c. 1100), Ibn al-Haïm (c. 1205), and
Jâber ibn Aflah (1100-1150) and tried to improve the “orthodox” astronomy, and those who
were critical of Ptolemy. Ibn Rushd, Maïmonid, Ibn Bâjja, and Ibn Tufayl dreamed of an
astronomy that would be in agreement with the Aristotelian physics, based on three kinds of
motions: centrifugal, centripetal, and circular about a center that should be identified with
Earth.
Al-Bitruji (Alpetragius, d. c. 1190) is the only Aristotelian philosopher to formulate an
embryo of a strictly geocentric system. However, there was nothing really new in this since
Democritus had allegedly stated similar ideas. Anyhow, Bitruji’s system was quite qualitative
and he was unable to calculate the planetary tables with his model. For example, in Bitruji’s
system Saturn could on occasion deviate from the ecliptic by as much as 26 degrees (instead
of the required 3 degrees).
The most significant improvement of Ptolemy’s model to comply with absolute
circularity was obtained in the 13th century by the Persian astronomer, mathematician,
founder of the Marâgha observatory, and political figure Nasireddin Tusi (1201-1274). He
devised a theorem on the combination of regular circular motions (now called the Tusi
couple) that generated linear motion. A circle of radius R rotates inside a circle of radius 2R.
The smaller circle rotates at twice the speed of the larger one and in opposite direction. The
initial tangent point will travel in linear motion back and forth along the diameter of the larger
circle. Today we know that the linear motion is a particular case of the family of hypocycloid
curves. A change in relative sizes and speeds of the circles will produce multi-cusped curves.
This innovation allowed Tusi to account for non-uniform motions of the planets. It explained
how the epicycle could move uniformly around the equant, and still oscillate back and forth
toward the center of the deferent. In fact the Tusi couple was the only new mathematical
model for planetary motions from the time of Ptolemy. Tusi used the theorem successfully in
his Tadhkira to reconstruct models for the Moon and the upper planets. However, as he
himself points out, he could not generalize the solution to Mercury. Tusi’s work was
elaborated by other astronomers at the Maragha observatory, Mu’ayyad ad-Din al-‘Urdi (d.
1266) and Qutbeddin Shirâzi (d. 1311). Finally, following Tusi’s work, a completely
concentric rearrangement of the planetary mechanisms was achieved by the Syrian
9
astronomer Ibn ash-Shâtir (1304-1376), who succeeded in eliminating not only the equant but
also certain other objectionable circles from Ptolemy’s models and thus placing the Earth at
the center of the Universe.
9. The eve of Scientific Revolution
On the eve of the Renaissance Spain was the major interface between Muslim science
and the rising European scientific activities. Many of the treatises written by Muslim scholars
reached Europe via Spain, where they were translated into Latin. The astrolabe was re-
introduced in the West through Spain in the 11th century. In about 1390 the English poet
Geoffrey Chaucer (c. 1340-1400), inspired by such translations, wrote an essay of the
astrolabe. It is possible that scientific activity centered at Oxford at the time contributed to the
surge of interest in the device. On the astrolabes of that period one finds typical sets of Arabic
star names written in Gothic Latin letters, for example Vega, Altair, Rigel, Alpheratz, and so
on. Thus, as a result of the astrolabe tradition of Andalusian astronomy, most navigational
stars today have Arabic names.
The Muslim attempts to criticize Ptolemy could not spare the Ptolemaic model.
Nonetheless, they were necessary steps in the path leading to the Scientific Revolution. They
certainly helped Nicolaus Copernicus (1473-1543) to come up with his revolutionary
heliocentric model, more especially since he included the concepts of epicycle and deferent to
explain the planetary motions. He was courageous and smart enough to replace the Earth with
the Sun and thus open a new era in the history of human thought. Nonetheless, he preserved
the geometrical/mathematical tools used by Ptolemy, since he still viewed celestial motions as
circular rather than elliptical and so still required the equant to describe elliptical motions.
The three Keplerian laws of planetary motions involving elliptical orbits (around 1610) and
Newton’s universal theory of gravitation (Principia 1687) were not yet discovered.
Copernicus knew of the results by Muslim scholars, and cites Battâni as well as the
Andalusians Zarqâli and Ibn Rushd. He made use of the Tusi couple in his models in De
revolutionibus orbium coelestium (On the Revolutions of the Celestial Spheres). There is
growing consensus that Copernicus, living some three centuries after Tusi, became aware of
Tusi’s result in some way, probably through Byzantine intermediaries, although an exact
chain of transmission has not yet been identified. In brief, Copernicus put an end to a long
period in the Middle Ages in which Muslim astronomers prevailed. Once the human thought
was freed from the shackles of geocentric model, unprecedented discoveries and revolutions
occurred owing to great scientific figures like Kepler, Galileo, Newton, and others.
Copernicus closed a parenthesis between the Golden Greek/Roman epoch and the
Renaissance. Europe started to excel in other fields than art and architecture.
The Alfonsine Tables created by Toledan astronomers in the 13th century were the
last major astronomical work by Spanish astronomers before the Renaissance. After the fall of
the Almohad dynasty (1121–1269), the Muslim Spain was reduced to the Nasrid kingdom of
Grenada (1232-1492). In spite of Alfonso X’s policy to attract Muslim scholars, they
preferred to settle in Grenada or emigrate to North Africa or the Middle East. The Christian
Spain gradually established itself as a European power. The discovery of America under
10
Spanish patronage was also a serious strike on Ptolemy’s worldview, since his Geographia
ignored the new continent. The discovery proved Ptolemy wrong as far as the form and
contents of the terrestrial surface were concerned, and this likewise had repercussions for his
celestial ideas. Further, the discovery boosted navigation and the need for exact time
determinations at different longitudes. More generally, it opened a new world with a wealth
of plants, animals, commodities, and peoples previously unknown to Europeans.
Another parallel which can be drawn between Persia and Andalusia is of linguistic
nature. In the wind of the Renaissance “national” languages started to step out of the
occultation by the lingua franca. In particular, the creation of a standardized Spanish
language based on the Castilian dialect began in the 13th century with King Alfonso X, who
encouraged scholars to write original works in Castilian and translate histories, chronicles,
scientific, legal, administrative, and literary works from other languages (principally Latin,
Greek, and Arabic). This was a tremendous job since Castilian had never been used
previously for writing on technical matters in particular astronomy. The Alfonsine scholars
had to exert much effort to build a new specialized language and create the necessary
terminology. As for Persian, although it had resisted Arabic and had created immense literary
masterpieces like Shâh-nâmeh by Ferdowsi (935-1020) as well as scientific and philosophical
works by scholars like Ibn Sinâ (Avicenna, c. 980-1037) and Biruni, the 13th century
witnessed a kind of surge in producing science and philosophy in Persian. This was mainly
encouraged by the above-mentioned astronomer/mathematician/politician Tusi, who himself
wrote several works in Persian: Zij-e Ilkhâni (based on observations at the Marâgha
Observatory), an astronomical treatise entitled Moiniyyé, and a work on ethics, Akhlâq-e
Nâseri. He even found time to translate into Persian the Book of Fixed Stars, by the Persian
astronomer Abd ar-Rahmân Sufi (903-986), originally written in Arabic.
References
Bagheri, M. 1998, The Persian Version of Zij-i Jâmi’ by Kušyâr Gilâni, in « La Science dans le Monde Iranien à
l’Epoque Islamique », Institut Français de Recherche en Iran, Téhéran
Breggren, L. 1997, Islamic Astronomers and the Equant, JHA, xxviii, 270
Calvo, E. 2005, Science in al-Andalus in the Lifetime of Khayyâm, Farhang (Institute for Humanities and
Cultural Studies, Tehran), Vol. 18, No 53-54, p. 329
Chabás, J.; Goldstein, B.R. 2000, Astronomy in the Iberian Peninsula: Abraham Zacut and the Transition from
Manuscript to Print, American Philosophical Society
Chabás, J.; Goldstein, B.R. 2003, The Alfonsine Tables of Toledo, Kluwer, Dordrecht
Copernic, N. Des révolutions des orbes célestes, traduction by Alexandre Koyré, 1998, Diderot Editeur
Dictionary of Scientific Biographies [DSB]. 1978, New York
Gingerich, O. 1986, Islamic Astronomy, Scientific American, 256, 74
Gingerich, O. 2002, An Annotated Census of Copernicus’ De Revolutionibus (Nuremberg, 1543 and Basel,
1566), Brille
Hartner, W. 1977, The Role of Observations in Ancient and Medieval Astronomy, JHA, viii, p. 1
Heydari-Malayeri, M., The Iranian Calendar, http://aramis.obspm.fr/~heydari/divers/calendar.html
Hugonnard-Roche, H. 1997, Influence de l’astronomie arabe en Occident médiéval, in Rashed 1997, p. 309
Kennedy, E.S. 1994, Tusi’s Cosmology, JHA, xxv, p. 321
King, D.A. 1987, Ninth-Century Islamic Astronomy, JHA, xviii, p. 284
Kremer, R.L. 2007, “Abbreviating” the Alfonsine Tables in Cracow, JHA, xxxviii, p. 283
Kokowski, M. 2004, Copernicus's Originality: Towards Integration of Contemporary Copernican Studies,
Polish Academy of Sciences, Warsaw
Morelon, R. (Edited and translated by). 1987, Thâbit ibn Qurra: OEuvres d’Astronomie, Les Belles Lettres, Paris
11
Neugebauer, O. 1957, The Exact Sciences in Antiquity, Dover Publications, Inc., New York
Pedersen, F. 2002, The Toledan Tables (A review of the manuscripts and textual versions with an edition), Royal
Danish Academy of Sciences and Letters
Ragep, F.J. 1992, Thâbit’s Astronomical Works, JHA, xxiii, p. 61
Ragep, F.J. 1993, Nasir al-Din al-Tusi’s Memoir on Astronomy, Springer Verlag New York (in two volumes)
Rashed, R. (sous la direction de). 1997, Histoire des sciences arabes, 1. Astronomie, théorie et appliquée.
Editions du Seuil, Paris
Rashed, R.; Vahabzadeh, B. 1999, Al-Khayyâm mathématicien, Librairie Scientifique et Technique Albert
Blanchard, Paris
Rosenfeld, B. 1974, Biruni, JHA, v, p. 134
Saliba, G. 1987, Theory and Observation in Islamic Astronomy: The Work of Ibn al-Shâter of Damascus, JHA,
xiii, p.35
Saliba, G. 1985, Solar Observations at the Maraghah Observatory before 1275: A New Set of Parameters, JHA,
xvi, p.113
Saliba, G. 1994, A Sixteenth-Century Arabic Critique of Ptolemaic Astronomy: The Work of Shams al-Din al-
Khafri, JHA, xxxv, p. 15
Saliba, G. 1997, Les théories planétaires en astronomie arabe après le XIe siècle, in Rashed 1997, p. 71
Sigismondi,, C.; Fraschetti, F. 2001, Measurement of the Solar Diameter in Kepler’s Time, The Observatory
121, 380
Stephenson, F.R.; Said, D.D. 1991, Precision of Medieval Islamic Eclipse Measurements, JHA, xxii, p. 197
Thurston, H. 1994, Early Astronomy, Springer Verlag
Veselovsky, I.N. 1973, Copernicus and Nasir al-Din al-Tusi, JHA, iv, p. 128
Vernet, J.; Samsó, J. 1997, Les développements de la science arabe en Andalousie, in Rashed 1997, p. 271
12
|
astro-ph/0109105 | 2 | 0109 | 2002-08-30T22:35:37 | Overproduction of primordial helium-4 in the presence of neutrino oscillations | [
"astro-ph",
"hep-ph"
] | The maximum overproduction of helium-4 in cosmological nucleosynthesis with active--sterile neutrino oscillations, nu_e -> nu_s, efficient after decoupling of electron neutrino, is analyzed. The kinetic effects on primordial nucleosynthesis due to neutrino spectrum distortion, caused by oscillations, are precisely taken into account.
The maximum overproduction of primordial He-4 as a function of oscillation parameters is obtained from the analysis of the kinetics of the nucleons and the oscillating neutrinos, for the full range of parameters of the discussed oscillation model. A maximum relative increase of He-4, up to 14% for non-resonant oscillations and up to 32% for resonant ones is registered. Cosmological constraints on oscillation parameters are also discussed. | astro-ph | astro-ph | CERN-TH/2002-217
Overproduction of primordial helium-4
in the presence of neutrino oscillations
D. P. Kirilova1
Theory Division, CERN, Geneva, Switzerland and
Institute of Astronomy, Bulgarian Academy of Sciences,
blvd. Tsarigradsko Shosse 72, Sofia, Bulgaria
Abstract
The maximum overproduction of helium-4 in cosmological nucle-
osynthesis with active -- sterile neutrino oscillations, νe ↔ νs, efficient
after decoupling of electron neutrino, is analyzed. The kinetic effects
on primordial nucleosynthesis due to neutrino spectrum distortion,
caused by oscillations, are precisely taken into account.
The maximum overproduction of primordial 4He as a function of oscil-
lation parameters is obtained from the analysis of the kinetics of the
nucleons and the oscillating neutrinos, for the full range of parame-
ters of the discussed oscillation model. A maximum relative increase
of 4He, up to 14% for non-resonant oscillations and up to 32% for
resonant ones is registered. Cosmological constraints on oscillation
parameters are also discussed.
2
0
0
2
g
u
A
0
3
2
v
5
0
1
9
0
1
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
1Regular Associate of the Abdus Salam ICTP, Trieste, Italy
1
1
Introduction
In this work I study the maximum overproduction of 4He in Big Bang Nucle-
osynthesis (BBN) with electron -- sterile neutrino oscillations νe ↔ νs.
The positive indications for oscillations, obtained by the neutrino exper-
iments (SuperKamiokande, SNO, Soudan 2, LSND, etc.) turned the subject
of neutrino oscillations into one of the hottest points of astrophysics and neu-
trino physics. The solar neutrino problem, the atmospheric neutrino anomaly
and the positive results of LSND experiments may be naturally resolved by
the phenomenon of neutrino oscillations, implying nonzero neutrino mass
and mixing. Although sterile neutrino impact in oscillations explaining at-
mospheric and solar neutrino anomalies, is strongly constrained from the
analysis of experimental oscillations data, still some small fraction of νs may
participate in oscillations. Hence, it is interesting to study cosmological ef-
fects of such oscillations.
Cosmological nucleosynthesis with neutrino oscillations was studied in
numerous publications, discussed in detail in [1, 2]. In these publications the
central goal was to obtain cosmological constraints on oscillation parameters.
The overproduction of 4He itself was not studied in detail until now. Such
a study may be of interest for constructing of alternative cosmological mod-
els, for constraining galactic chemical evolution, it may be useful for non-
standard models predicting active-sterile oscillations, like models with extra
dimensions, mirror world particles, etc.
Here I address that question of the possible maximal overproduction
of 4He due to oscillations. For that purpose the case of νe ↔ νs, oscilla-
tions, effective after the electron neutrino decoupling from the plasma (i.e
for δm2 sin2 2ϑ ≤ 10−7eV2), is the most suitable. In that case due to the fact
that sterile neutrino state is usually less populated than the active one at
the start of oscillations, the oscillations may cause strong distortion of the
electron neutrino spectrum, which affects the kinetics of the nucleons freezing
before nucleosynthesis, and hence, the primordial production of 4He. And,
as will be shown by the numerical analysis, this kinetic effect of oscillations,
may be much bigger than the one corresponding to an increase in the total
energy density due to an additional neutrino flavor δNρ, mainly considered
in literature.
We analyze 4He primordial production, taking into account all known ef-
fects of νe ↔ νs oscillations on the primordial synthesis of 4He. The produc-
2
tion of 4He was calculated in the non-resonant and resonant oscillation cases.
In both cases strong overproduction of 4He was found possible - up to 14%
and 32%, correspondingly. Thus in the discussed model of non-equilibrium
oscillations the maximum overproduction of 4He corresponds to an increase
of the neutrino effective degrees of freedom δN max
The oscillation effects on BBN and their description are briefly reviewed
in the next section. The kinetic approach, the results on 4He primordially
produced abundance in the presence of oscillations, and the cosmological
constraints on oscillation parameters are discussed in the last section.
kin ∼ 6.
2 Cosmological Nucleosynthesis with Neutrino
Oscillations
2.1 Standard Cosmological Nucleosynthesis
According to the standard BBN, during the early hot and dense epoch of the
Universe, the light elements D, 3He, 4He, 7Li were synthesized successfully.
The most reliable and abundant data are now available for 4He. This fact
and its relatively simple chemical evolution make 4He the preferred element
for the analysis of the oscillations effect on BBN.
The contemporary values for the mass fraction of 4He, Yp, inferred from
observational data, are in the range 0.238 -- 0.245 (the systematic errors are
supposed to be around 0.007) [3].
Primordially produced 4He abundance Yp, is calculated with great preci-
sion within the standard BBN model [4]. According to the standard BBN
4He is a result of a complex network of nuclear reactions, proceeding after
the neutron-to-protons freezing. It essentially depends on the freezing ratio
(n/p)f . The latter is a result of the freezing of the weak processes:
νe + n ↔ p + e−
e+ + n ↔ p + νe,
(1)
which maintained the equilibrium of nucleons at high temperature (T > 1
MeV). Their freeze-out occurred when in the process of Universe cooling
the rates of these weak processes, Γw, became comparable and less than the
expansion rate H(t):
3
Γw ∼ G2
F E 2
ν Nν ≤ H(t) ∼ √gef f T 2
Thus, the produced 4He is a strong function of the number of the effective
degrees of freedom at BBN epoch, gef f = 10.75 + 7/4δNρ.
Yp depends also on the electron neutrino spectrum and number density,
In
and on the neutrino-antineutrino asymmetry, which enter through Γw.
the standard BBN model three neutrino flavors, zero lepton asymmetry and
equilibrium neutrino number densities and spectrum distribution are postu-
lated:
nνe(E) = (1 + exp(E/T ))−1.
Almost all neutrons, present at the beginning of nuclear reactions, are
sucked into 4He. So, the primordially produced mass fraction of 4He can be
approximated by
Yp ∼ 2(n/p)f /(1 + n/p)f exp(−t/τn).
where τn is the neutron mean lifetime.
The theoretical uncertainty of the precisely calculated Yp is less than 0.1%
(δYp < 0.0002) within a wide range of η. So, in the standard BBN scenario,
where τn, and gef f are fixed, Yp, as well as the rest cosmologically produced
light elements are functions of only one parameter - the baryon-to-photon
ratio η.
Deuterium measurements in pristine environments towards low metal-
licity quasar absorption systems at very high z ∼ 3 provide us with the
most precision determination of the baryon density [5], giving the value:
η = (5.6 ± 0.5) × 10−10. Recently, the baryon density was also determined
from observations of the anisotropy of the cosmic microwave background
(CMB) by DASI [6], BOOMERANG [7, 8], CMB [9] and MAXIMA exper-
iments [10]. For the combined analysis of the data see also [11, 12]. The
CMB anisotropy data is in remarkable agreement with the baryon density
determined from deuterium measurements and BBN.
So, the predicted primordial 4He abundance Yp is in accordance with the
observational data and is consistent with deuterium measurements.
4
2.2 Effects of neutrino oscillations on nucleosynthesis
In case neutrino oscillations are present in the Universe primordial plasma,
they lead to changes in the Big Bang Nucleosynthesis, depending on the
oscillation channels and the way they proceed. The effect of flavor neutrino
oscillations on BBN is negligibly small because the temperatures and hence,
the densities of the neutrinos with different flavors are almost equal.
Active-sterile oscillations, however, are capable to shift neutrino num-
ber densities and spectrum from their equilibrium values. Besides, they
may change neutrino-antineutrino asymmetry and excite additional neutrino
types. Thus, the presence of neutrino oscillations invalidates the main BBN
assumptions about three neutrino flavors, zero lepton asymmetry and equi-
librium neutrino energy distribution.
Qualitatively, the oscillations effects on the nucleosynthesis of 4He may
be illustrated as follows:
• excitation of additional degrees of freedom
It is known that active-sterile oscillations may keep sterile neutrinos in ther-
mal equilibrium [13] or bring them into equilibrium in case they have already
decoupled. The presence of light steriles in equilibrium leads to an increase
of the effective degrees of freedom during BBN and to faster Universe ex-
pansion H(t) ∼ g1/2
ef f and earlier n/p-freezing, Tf ∼ (gef f )1/6, at times when
neutrons were more abundant [13, 14]:
n/p ∼ exp [−(mn − mp)/Tf ]
This effect leads to 5% 4He overproduction corresponding to one additional
neutrino type brought into equilibrium by oscillations.
• distortion of the neutrino spectrum
Much stronger effect of oscillations may be achieved in case of oscillations
between initially empty sterile neutrino state and electron neutrino. The
non-equilibrium initial condition leads to spectrum distortion of the active
neutrino due to oscillations.
Since oscillation rate is energy dependent Γ ∼ δm2/E the low energy
neutrinos start to oscillate first, and later the oscillations become noticeable
5
for the more energetic neutrinos. Due to that, the neutrino energy spectrum
nν(E) may strongly deviate from its equilibrium form [15]. This spectrum
distortion affects the kinetics of nucleons freezing - it leads to an earlier
n/p-freezing and an overproduction of 4He yield.
The effect can be easily understood having in mind that the distortion
leads both to a depletion of the active neutrino number densities Nν [16]:
Nν ∼ Z dEE 2nν(E)
and to a decrease of the mean neutrino energy, which reflects into a decrease
ν Nν, and hence, to an overproduction of 4He pri-
of the weak rates Γw ∼ E 2
mordial abundance.
The decrease of the electron neutrino energy due to oscillations into
low temperature sterile neutrinos, has also an additional effect: Due to the
threshold of the reaction converting protons into neutrons, when neutrinos
have lower energy, protons are preferably produced in reactions (1), which
may lead to an underproduction of 4He [17]. However, this turns to be a
minor effect.
The effect of spectrum distortion was numerically analyzed for hundreds
combinations of mass differences and mixing angles in previous studies of
active-sterile oscillations [15, 18, 19, 20]. It was proved important both in
the resonant [19] and in the non-resonant oscillations case [20].
In the discussed here scenario it is the dominant effect and leads to over-
production of the primordial 4He abundance.
• neutrino-antineutrino asymmetry growth
The idea of neutrino-antineutrino asymmetry generation during the resonant
transfer of neutrinos was first proposed in ref. [21]. Dynamically produced
asymmetry exerts back effect to oscillating neutrino and may change its oscil-
lation pattern [22, 23], it may suppresses oscillations at small mixing angles,
leading to weakening the oscillations effect on BBN, i.e. less overproduction
of
4He. 2
The effect of the oscillations generated asymmetry on 4He for the discussed
here model was analyzed for hundreds of δm2 − ϑ combinations in [19].
2There were independent studies of the asymmetry growth in active-sterile oscillations
for the case of large mass differences along the lines of the pioneer work [24].
6
3 Helium-4 overproduction due to νe ↔ νs
neutrino oscillations
We have provided an exact study of all the oscillation effects on the primordial
production of 4He.
3.1 The required kinetic approach
For the analysis of the non-equilibrium picture of active -- sterile neutrino os-
cillations, producing non-equilibrium neutrino number densities, distorting
neutrino spectrum and generating neutrino-antineutrino asymmetry a self-
consistent numerical analysis of the evolution of the nucleons and the oscil-
lating neutrinos in the high temperature Universe was provided.
Exact kinetic equations for the nucleons and for the neutrino density ma-
trix in momentum space [18] were used. This allowed to describe precisely the
kinetic effects of oscillations on helium production due to spectrum distortion
at each neutrino momentum.
The equation for the neutron number densities in momentum space nn
reads:
(∂nn/∂t) = Hpn (∂nn/∂pn) +
+Z dΩ(e−, p, ν)A(e−p → νn)2 [ne−np(1 − ρLL)−
−nnρLL(1 − ne−)]
−Z dΩ(e+, p, ν)A(e+n → pν)2 [ne+nn(1 − ¯ρLL)−
−np ¯ρLL(1 − ne+)] .
(2)
where dΩ(i, j, k) is a phase space factor and A is the amplitude of the cor-
responding process, neutrino ρLL and antineutrino number densities ¯ρLL at
each integration step of eq. (2) are taken from the simultaneously performed
integration of the set of equations for neutrino density matrix (see ref. [18]).
The equation provides a simultaneous account of the different compet-
ing processes, namely: neutrino oscillations (entering through ρLL and ¯ρLL),
Hubble expansion (first term) and weak interaction processes (next terms).
The numerical analysis was performed for the temperature interval [2 MeV,
0.3 MeV].
7
We have found that for a wide range of oscillation parameters the spec-
trum distortion of electron neutrino is considerable during the period of nu-
cleons freezing. Hence, usually the kinetic effect of oscillations due to electron
neutrino energy spectrum distortion plays the dominant role in the overpro-
duction of helium.
In Fig. 1 the evolution of the energy spectrum of the
electron neutrino through the period of nucleons freezing is illustrated. In
the Figs.1a,1b and 1c the energy spectra at different characteristic tempera-
tures, correspondingly T = 1 MeV, T = 0.7 MeV and T = 0.5 are presented.
By the dashed curve the equilibrium spectrum at the given temperature is
given for comparison. The spectra are calculated for oscillation parameters
δm2 = 10−7 eV2 and sin2 2ϑ = 0.1.
It is seen that oscillations affect non trivially the neutrino spectrum and
distort strongly its equilibrium form. So, it is not possible to describe cor-
rectly the spectrum distortion due to oscillations simply by shifting the effec-
tive temperature of the neutrino and accounting only for the depletion of its
total number density, considering its spectrum equilibrium as in refs. [26, 27].
For the proper description of the spectrum distortion in the case of non-
equilibrium electron-sterile oscillations, studied here, the evolution of the
neutrino ensembles should be explored using neutrino density matrix in mo-
mentum space, allowing to describe the evolution of neutrino at each mo-
mentum.
The neutrino-antineutrino asymmetry in the resonant case grows up to
5 orders of magnitude from its initial value (taken to be of the order of the
baryon asymmetry). So, this asymmetry influences BBN only indirectly -
through oscillations. This oscillations generated asymmetry leads to a de-
crease in the produced 4He compared with the case without asymmetry ac-
count. However, its effect comprises only up to about a 10% of the total
effect.
3.2 Maximum helium-4 overproduction
The overproduction of the primordial 4He, δYp = Y osc
p − Yp in the presence of
νe ↔ νs oscillations was calculated for mass differences δm2 ≤ 10−7 eV2 and
0 ≤ ϑ ≤ π/4. Several hundreds of δm2 − ϑ combinations were explored.
We have used the data from the precise calculations of the n/p-freezing
provided for the non-resonant case in [20] and for the resonant case in [19].
As far as it is the essential for the production of 4He. The neutron decay was
8
accounted adiabatically till the beginning of nuclear reactions at about 0.09
MeV.
We have found that the effect of oscillations becomes very small (less
than 1%) for small mixings: as small as sin22ϑ = 0.1 for δm2 = 10−7 eV2,
and for small mass differences: δm2 < 10−10 eV2 at maximal mixing. For
very small mass differences δm2 ≤ 10−11 eV2, or at very small mixing angles
sin2 2ϑ ≤ 10−3, the effect on nucleosynthesis becomes negligible.
In the non-resonant case the oscillation effect increases with the increase
of the oscillation parameters, hence it is maximal at maximal mixing and
greatest mass differences. In Fig. 2 (the lower curve) the maximal relative
increase in the primordial 4He as a function of neutrino mass differences at
maximal mixing: δY max
p /Yp(δm2)θ=π/4 is presented.
/Yp = δY osc
p
It is seen that for maximal mixing, the oscillation effect becomes greater
than 5% (the one corresponding to one additional neutrino type) already at
δm2 ≥ 3 × 10−9 eV2 (in the resonant case) and δm2 ≥ 6 × 10−9 eV2 (in
the non-resonant one). It continues to grow up with the increase of the mass
differences, and at δm2 ∼ 10−7 eV2 is several times bigger: δN max
kin ∼ 3 in
the non-resonant case and δN max
Further increase of the mass differences, however, will lead to oscillations
effective before electron neutrino decoupling, and therefore, to a smaller spec-
trum distortion effect, because the interactions with the plasma will lead to
faster thermalization of the sterile state. Hence, the effect on helium will
decrease with further increase of δm2 and finally reach an overproduction
of 5% again, corresponding to a full thermalization of the sterile state and
its equilibrium spectrum.
kin ∼ 6 for the resonant one.
In Fig. 3 we present a combined plot (for the resonant and the non-
resonant oscillation case) of δYp dependence on the mixing angle for δm2 =
10−7 eV2 and δm2 = 10−8 eV2. While in the non-resonant case the oscillations
effect increases with the increase of the mixing (see l.h.s of Fig.3.), in the
resonant case for a given δm2 there exists some resonant mixing angle, at
which the oscillations are enhanced by the medium (due to the MSW effect),
and hence, the overproduction of 4He is greater than that corresponding to
the vacuum maximal mixing angle. This behavior of the helium production
on the mixing angle is illustrated in the r.h.s. of the figure.
The upper curve in Fig.2 shows the maximal relative increase in the
resonant oscillations case as a function of mass differences. Each maximum
4He value corresponds to the resonant mixing angle for the concrete mass
9
(δm2, ϑres
p
p
/Yp = Y osc
difference: δY max
δm). As can be seen from Figs. 2 and 3,
a considerable overproduction can be achieved:
in the resonant case up to
32% and in the non-resonant one -- up to 14%. So, the net effect of spectrum
distortion of electron neutrino due to oscillations on the production of 4He
may be considerable and several times larger than the effect due to excitation
of one additional neutrino type.
Several words are due to cosmological constraints on neutrino oscillation
parameters, calculated in the discussed model of oscillations. Due to the
strong kinetic effect of neutrino spectrum distortion caused by active-sterile
oscillations discussed, the obtained constraints in that model of oscillations
are more stringent than those obtained in pioneer works, not accounting for
the kinetic effects of oscillations. Hence, according to these more precise
studies of spectrum distortion effects of oscillations, the δYp/Yp < 3% limit
excludes almost completely the LOW electron-sterile solution to the solar
neutrino problem [25], in addition to the excluded sterile LMA solution in
previous investigations.
it is more than an order of magnitude more
restrictive to the mass differences.
I.e.
This study has shown that considerable Yp overproduction may result
from the electron neutrino spectrum distortion due to νe ↔ νs oscillations.
The overproduction is maximal for the case of initially empty sterile neu-
trino state, considered here. The dependence of 4He overproduction on the
degree of population of the sterile neutrino state before νe ↔ νs oscillations
is considered in a separate paper [28]. Bigger initial population of νs leads to
a smaller spectrum distortion of νe, and hence smaller kinetic effect on pri-
mordial nucleosynthesis δNkin and higher energy density due to the increase
of the number of degrees of freedom δNρ. The interplay between the two
effects, however, is such that the overproduction of 4He is smaller than in the
case when initially δNρ = 0, discussed here.
4 Conclusions
The primordial production of 4He in the presence of νe ↔ νs oscillations, effec-
tive after electron neutrino decoupling was analyzed. A precise quantitative
study of the maximum overproduction of 4He, accounting for all oscillations
effects is provided. It was shown that the considerable spectrum distortion
10
of the electron neutrino caused by oscillations, which effects the kinetics of
the neutron-proton transitions during nucleons freezing, plays the dominant
role in the overproduction of 4He.
Enormous overproduction of 4He (up to 32% in the resonant case and
14% in the non-resonant case) was found possible in case the sterile neutrino
state was empty at the start of oscillations.
The results of this analysis can be useful for constraining nonstandard
physics, predicting active-sterile neutrino oscillations, like extra-dimensions,
producing oscillations, supernova bursts employing oscillations, etc.. It can
be of interest also for models of galactic chemical evolution.
Acknowledgments
I thank M. V. Chizhov for the help during the preparation of this paper.
This work has been completed during my visiting position at TH CERN. I
appreciate also the Regular Associateship at the Abdus Salam ICTP, Trieste.
References
[1] Kirilova, D. P., and Chizhov, M. V., 2001, CERN-TH/2001-020, astro-
ph/0108341
[2] Dolgov A., hep-ph/0202122.
[3] Izotov, Y. I., and Thuan, T. X., Ap. J. 500 (1998) 188.
[4] Lopez, R., and Turner, M. S., Phys. Rev. D 59 (1999) 103502.
[5] Meara J. et al., Ap.J. 552 (2001) 718.
[6] Pryke S. et al., ApJ 5681 (2002) 46.
[7] Netterfield C. et al., ApJ 571 (2002) 604.
[8] Bernardis P. et al., ApJ 564 (2002) 559.
[9] Padin S. et al., ApJ Letters 549 (2001) L1.
[10] Stompor B. et al., ApJ 561 (2001) L7.
11
[11] Wang X., Tegmark M., Zaldarriaga M., Phys.Rev. D65 (2002) 123001
[12] Douspis M. et al., Astr.Astrop. 379 (2001) 1.
[13] Dolgov, A. D., Sov. J. Nucl. Phys. 33 (1981) 700.
[14] Fargion, D., and Shepkin, M., Phys. Lett. B 146 (1984) 46.
[15] Kirilova, D. P., 1988, JINR E2-88-301, Dubna, Russia.
[16] Barbieri, R., and Dolgov, A., Phys. Lett. B 237 (1990) 440.
[17] Dolgov A.,Kirilova D., Int.J.Mod.Phys.A3 (1988) 267.
[18] Kirilova, D. P., and Chizhov, M. V., Phys. Lett. B 393 (1997) 375.
[19] Kirilova, D. P., and Chizhov, M. V., Nucl. Phys. B 591 (2000) 457.
[20] Kirilova, D. P., and Chizhov, M. V., Phys. Rev. D 58 (1998) 073004.
[21] Mikheyev, S., and Smirnov, A., 1986, in VI Moriond Meeting on Massive
Neutrinos in Particle Physics and Astrophysics, eds. O. Fackler and J.
Tran Thanh Van, Editions Frontiers, Tignes, p. 355.
[22] Kirilova, D., and Chizhov, M.,1996, in 17 International Conference on
Neutrino Physics and Astrophysics, NEUTRINO 96, eds. K. Enqvist,
K. Huitu, and J. Maalampi, World Scientific, Helsinki, p. 478.
[23] Kirilova, D. P., and Chizhov, M. V., 1999, hep-ph/9908525.
[24] R. Foot, M. J. Thomson and R. R. Volkas, Phys. Rev. D 53 (1996)
R5349
[25] Kirilova, D. P., and Chizhov, M. V., Nucl. Phys. B Proc. Suppl. 100
(2001) 360.
[26] Shi X., Schramm D. N., and Fields B. D., Phys. Rev. D 48, 2563 (1993).
[27] Shi X., and Fuller G. M., Phys. Rev. D 59, 063006 (1999).
[28] Kirilova D., CERN-TH-2002-209
12
Figure 1a: The figure illustrates the degree of distortion of the electron
neutrino energy spectrum x2ρLL(x), where x = E/T , caused by oscillations
with mass difference δm2 = 10−7 eV2 and sin2 2ϑ = 0.1 at a character-
istic temperature 1 MeV. The dashed curve gives the equilibrium neutrino
spectrum for comparison.
13
Figure 1b: The figure illustrates the degree of distortion of the electron
neutrino energy spectrum x2ρLL(x), where x = E/T , caused by oscillations
with mass difference δm2 = 10−7 eV2 and sin2 2ϑ = 0.1 at a characteristic
temperature 0.7 MeV.
14
Figure 1c: The figure illustrates the degree of spectrum distortion of
the electron neutrino caused by oscillations with mass difference δm2 = 10−7
eV2 and sin2 2ϑ = 0.1 at a characteristic temperature 0.5 MeV.
15
Figure 2: Maximum primordial 4He abundance for the resonant (upper
curve) and the non-resonant oscillation case (lower curve), as a function of the
neutrino mass differences. The non-resonant case is calculated at maximum
mixing, while in the resonant case the helium abundance is calculated at the
resonant mixing angle for the corresponding mass difference.
16
Figure 3: The dependence of the relative increase of primordial helium
on the mixing angle for the resonant (r.h.s.) and non-resonant (l.h.s.) oscil-
lation case. The upper curve corresponds to δm2 = 10−7 eV2, the lower one
to δm2 = 10−8 eV2.
17
|
astro-ph/0204159 | 1 | 0204 | 2002-04-09T15:52:48 | Is RX J185635-375 a Quark Star? | [
"astro-ph"
] | Deep Chandra LETG+HRC-S observations of the isolated neutron star candidate RX J1856.5-3754 have been analysed to search for metallic and resonance cyclotron spectral features and for pulsation behaviour. As found from earlier observations, the X-ray spectrum is well-represented by a ~ 60 eV (7e5 K) blackbody. No unequivocal evidence of spectral line or edge features has been found, arguing against metal-dominated models. The data contain no evidence for pulsation and we place a 99% confidence upper limit of 2.7% on the unaccelerated pulse fraction over a wide frequency range from 1e-4 to 100 Hz. We argue that the derived interstellar medium neutral hydrogen column density of 8e19 <= N_H <= 1.1e20 per sq. cm favours the larger distance from two recent HST parallax analyses, placing RX J1856.5-3754 at ~ 140 pc instead of ~ 60 pc, and in the outskirts of the R CrA dark molecular cloud. That such a comparatively rare region of high ISM density is precisely where an isolated neutron star re-heated by accretion of interstellar matter would be expected is either entirely coincidental, or current theoretical arguments excluding this scenario for RX J1856.5-3754 are premature. Taken at face value, the combined observational evidence -- a lack of spectral and temporal features and an implied radius at infinity of 3.8-8.2 km that is too small for current neutron star models -- points to a more compact object, such as allowed for quark matter equations of state. | astro-ph | astro-ph |
To appear in the Astrophysical Journal
Is RXJ1856.5-3754 a Quark Star?
Jeremy J. Drake1 , Herman L. Marshall2 , Stefan Dreizler3 , Peter E. Freeman1 ,
Antonella Fruscione1 , Michael Juda1 , Vinay Kashyap1 , Fabrizio Nicastro1 , Deron
O. Pease1 , Bradford J. Wargelin1 , Klaus Werner3
1Smithsonian Astrophysical Observatory, MS-3, 60 Garden Street, Cambridge, MA 02138
2MIT Center for Space Research, Cambridge, MA 02139
3 Institut fur Astronomie und Astrophysik, Universitat Tubingen, Sand 1, 72076 Tubingen,
Germany
ABSTRACT
Deep Chandra LETG+HRC-S observations of the isolated neutron star can-
didate RX J1856.5−3754 have been analysed to search for metallic and resonance
cyclotron spectral features and for pulsation behaviour. As found from earlier
observations, the X-ray spectrum is well-represented by a ∼ 60 eV (7 × 105 K)
blackbody. No unequivocal evidence of spectral line or edge features has been
found, arguing against metal-dominated models. The data contain no evidence
for pulsation and we place a 99% confidence upper limit of 2.7% on the unac-
celerated pulse fraction over a wide frequency range from 10−4 to 100 Hz. We
argue that the derived interstellar medium neutral hydrogen column density of
8 × 1019 ≤ NH ≤ 1.1 × 1020 cm−2 favours the larger distance from two recent
HST parallax analyses, placing RX J1856.5−3754 at ∼ 140 pc instead of ∼ 60 pc,
and in the outskirts of the R CrA dark molecular cloud. That such a compar-
atively rare region of high ISM density is precisely where an isolated neutron
star re-heated by accretion of interstellar matter would be expected is either en-
tirely coincidental, or current theoretical arguments excluding this scenario for
RX J1856.5−3754 are premature. Taken at face value, the combined observa-
tional evidence—a lack of spectral and temporal features, and an implied radius
R∞ = 3.8-8.2 km that is too small for current neutron star models—points to a
more compact ob ject, such as allowed for quark matter equations of state.
Subject headings: stars:
X-rays: stars
individual (RX J1856.5−3754) — stars: neutron —
– 2 –
1.
Introduction
The structure and evolution of neutron stars depends on the properties of matter at
nuclear and supranuclear densities. Such conditions are not achievable in terrestrial labora-
tories and the theoretical description of such superdense matter remains uncertain. It has
therefore been hoped that neutron stars, and in particular those that are isolated and not
complicated by strong accretion or magnetospheric signatures, might provide some empirical
insights: observations of their masses, radii and cooling characteristics could, in principle,
provide useful constraints for the equation of state (EOS) of dense matter (e.g. Lattimer &
Prakash 2001 and references therein).
In a relatively brief period of 106 -107 yr, a hot, isolated neutron star (INS) born in a
supernova explosion can cool, cease pulsar activity and become essentially inactive (see, e.g.,
the review of Treves et al. 2000). Of the estimated 108 -109 isolated neutron stars thought
to inhabit the Galaxy, only a tiny fraction are therefore expected to be sufficiently young
to remain hot and visible in X-rays. One possible mechanism capable of sustaining thermal
X-ray emission in an older INS is accretion of material from the ISM. To date, only a handful
of these older INS candidates have been found.
The soft X-ray source RX J1856.5−3754 discovered by Walter, Wolk & Neuhauser (1996)
is the brightest, and probably the closest (Kaplan, van Kerkwijk & Anderson 2002) of the
INS candidates. It was identified with a very faint (V ≃ 25.6) optical counterpart by Walter
& Matthews (1997), was found to have an optical flux about a factor of 2-3 higher than that
predicted by the Rayleigh-Jeans tail of the ∼ 55 eV blackbody spectrum that represents the
low resolution ROSAT PSPC spectrum (e.g. Walter et al. 1996; Campana et al. 1997, Pons et
al. 2001), and lies in the line-of-sight toward the dark molecular cloud R CrA. However, the
exact nature of RX J1856.5−3754 remains unknown—whether it is a fairly young, cooling
INS, perhaps undetected as a pulsar because of unfortunate beam alignment, or an older
ob ject reheated by ISM accretion.
Very Large Telescope (VLT) observations have recently revealed an Hα nebula round
RX J1856.5−3754 and a blackbody spectrum through the UV-optical range (van Kerkwijk
& Kulkarni 2001a,b). Hubble Space Telescope (HST) astrometry was used by Walter (2001)
and Kaplan et al. (2002) to estimate a parallax and proper motion, but with conflicting
results (16.5 ± 2.3 mas vs. 7 ± 2 mas). Walter (2001) argued that the proper motion points
to the Sco-Cen OB association and an age of ∼ 106 yr. However, Pons et al. (2001) failed
to detect the expected pulsation signature in ROSAT and ASCA data. Modelling X-ray,
EUV, UV and optical spectra using the Walter (2001) parallax and assuming a metal-
dominated atmosphere resulted in stellar radii too small for current EOS and smaller than
the Schwarzschild radius for a canonical 1.4M⊙ star, leading them to conclude that the
– 3 –
surface temperature distribution could be inhomogeneous. In contrast, the same atmospheric
analysis using the larger distance of Kaplan et al. (2002) would yield a radius consistent with
current theory. If the atmosphere is indeed metal-dominated, then line and edge features
should be visible in high resolution X-ray spectra.
RX J1856.5−3754 was observed in 2000 March for 55ks by the Chandra X-ray Observa-
tory Low Energy Transmission Grating (LETG) and the High Resolution Camera spectro-
scopic microchannel plate detector array (HRC-S) to look for spectral features and pulsar
activity. Within relatively large statistical uncertainties, Burwitz et al. (2001) failed to de-
tect either significant departures from a blackbody spectrum or pulsations in these data.
The prospect of important scientific gains from a longer observation prompted, at the re-
quest of different researchers, the undertaking of a very recent set of Chandra observations of
RX J1856.5−3754 under director’s discretionary time using LETG+HRC-S. A period search
using these data by Ransom, Gaensler & Slane (2002) has already placed an upper limit on
a pulse fraction of 4.5%. This Letter presents spectroscopic and independent timing analyses
of these data.
2. Observations and Data Reduction
The observations analysed in this paper Pipeline-processed (CXC software version 6.3.1)
photon event lists were reduced and analysed using the CIAO software package version 2.2,
and independently using custom IDL1 software. Processing included extra pulse-height fil-
tering to reduce background (Wargelin et al., in preparation) and barycentric correction of
event times. Dispersed photon events were extracted using the now standard “bow tie” win-
dow, and a small circular region (31-pixel radius2 ) was used to extract the 0th order events.
We note that an extraneous bright feature (likely a ghost image of a bright off-axis source)
appeared on the +1 order outer HRC-S plate which introduces spurious features into the
extracted spectrum and background near 110 A. We applied corrections to extracted event
times to ameliorate an HRC electronics problem that assigns time tags for each event to the
event that immediately follows. If every event were telemetered to the ground, correct times
could be easily reassigned, but because of the high HRC background rate and a telemetry
limit of 184 events/sec, only valid events, which make up approximately 45% of the total
1 Interactive Data Language, Research Systems Inc.
2Our 31 pixel radius extraction region for the 0th order source corresponds to an encircled energy fraction
of about (92 ± 3)%, which combines with an average deadtime of 0.59% to yield a correction factor for a 2π
steradian aperture of 1.09.
– 4 –
events in these observations, are recorded in data received on the ground. We have reas-
signed the time of every valid event to the preceding valid event during ground processing;
the event times will then be correct about Nvalid /Ntotal of the time. One can place an upper
limit of δ t on the time errors associated with this process, however, simply by excluding
events with a time-shift of more than δ t; the average timing error of such events will, of
course, be less than δ t. However, if δ t is too small, too many events get excluded to perform
sensitive timing studies. We have adopted a δ t of 2 ms, retaining 20% of the original counts
in the corrected data set.
3. Count Rates and Timing Analysis
Count rates were derived for each observation segment from event lists filtered to exclude
times of high background and telemetry saturation, when the telemetered valid event rate
exceeded 184 count s−1 , and to exclude high pulse-height events that arise entirely from
background. These rates corresponding to 0th order only are listed in Table 1.
The count rates for the different observation segments are statistically consistent, though
the 2000 March observation (113) rate of 0.2195 ± 0.0020 count s−1 lies 1.9% and 1.9σ above
the rate for the combined 2001 October series rate of 0.2155 ± 0.0008. Such deviations will be
obtained by chance about 3% of the time when the count rates do not vary; it is thus more
likely attributable to quantum efficiency (QE) variations in the detector on small scales.
Indeed, Pons et al. (2001) note two ROSAT HRI observations obtained 3 yr apart that are
consistent to 1%. The fluxes corresponding to HRI and PSPC count rates are higher than
that obtained from our Chandra data by 20% and 30%, respectively. As remarked by Burwitz
et al. (2001), these differences are likely attributable to absolute calibration uncertainties.
We have also examined the ASCA SIS observation described Pons et al. (2001) and find
fluxes 30-40 % lower than obtained by Chandra for the 20-30 A range, but in agreement
with Chandra shortward of 20 A. The EUVE DS count rate of Pons et al. (2001) is also
consistent with the Chandra observation within allowed uncertainties.
Our search for pulsations used three different techniques, none of which found any
evidence for significant variability. In contrast to Ransom et al. (2002), we did not include
a deceleration term; this will be discussed below.
We applied the Bayesian method of Gregory & Loredo (1992) to both the time-corrected
and uncorrected 0th order event lists. The method tests for variability by comparing the fits
of periodic stepwise models to the data with the fit of a constant model. The odds favoring
variability (based on eqn. 5.28 of Gregory & Loredo 1992, with a maximum number of steps
– 5 –
mmax = 12 and limiting angular frequencies ωlo and ωhi equal to 10−4 and 103 , respectively)
were found to be 1.45 × 10−4 for the whole data set and 3.75 × 10−3 for the data filtered on
δ t = 2 ms, both to be compared with an odds value of 102 needed for a confident pulsation
or variability detection.
Our second method employed an FFT analysis applied to the combined 0th and 1st
order events, followed by a likelihood ratio test (LRT) to determine limits on the pulse
fraction. The FFT power distribution was consistent with shot noise. For the LRT of a
given period, P , the data were binned into N phase bins, giving ni counts in each. The
source model was y = A + f cos(φi + φ0 ) where φi is the phase of bin i and φ0 is the phase
of the pulse and f /A is the pulse fraction. The likelihood equations were then solved for A
and the process applied to > 500 frequencies where the FFT power exceeded a critical level
in the frequency range 0.001-50 Hz. By including also the dispersed events we improved the
signal-to-noise of the result by a factor of 1.47 compared to using 0th order alone and could
obtain a pulse fraction limit lower than the value of 4.5% obtained by Ransom et al. (2002)
in their unaccelerated search. A pulse fraction upper limit (99% confidence) of 2.7% was
derived applying our likelihood ratio method using all data, including the dispersed events
with 1 < λ < 70 A. Taking only the events limited by δ t < 2 ms, the pulse fraction limit is
10%.
Thirdly, we computed the Lomb-Scargle periodogram for both the time-corrected and
uncorrected photon arrival time differences in the frequency range 0.01-1 Hz for the events
in ObsID’s 3380 3381, 3382 and 3399. Again, no significant peaks were present.
The assumption of a negligible deceleration term in our period search restricts the range
of periods and dipole magnetic field strengths for which our search is valid. The coherence
limit for phase slippage by 10 % over the duration of the 2001 October observations implies
that our result is valid for a magnetic field upper limit B < 2.3 × 1013P 3/2 G for period
P s (e.g. Shapiro & Teukolsky 1983). As noted by Ransom et al. (2002), this range would
exclude very young and energetic neutron stars, such as the Crab and Vela pulsars, though
most of these younger ob jects are also conspicuously strong radio pulsars. All anomalous
X-ray pulsars would lie within our sensitivity limit range. RX J1856.5−3754 is also most
unlikely to be an extremely young ob ject based on its modest temperature and luminosity,
which are consistent with an ob ject of age ∼ 105 yr on canonical cooling curves (e.g. Tsuruta
1997).
– 6 –
4. Spectral Analysis and Model Parameter Estimation
Spectral analysis in the form of model parameter estimation was undertaken using the
CIAO/Sherpa fitting engine and independently using specially-written IDL software. Cur-
sory inspection of the spectrum leads immediately to the conclusion that there are no obvious
features indicative of absorption lines or edges. We found that blackbody models represent
the high resolution spectra well, in agreement with earlier studies. We modelled +1 and −1
orders both separately and simultaneously, and added together; results from these differ-
ent approaches were statistically indistinguishable. Representative results of model fits and
residuals are illustrated in Figure 1. Two sets of best-fit parameters were obtained from in-
dependent analyses that invoked (1) the existing first and higher order CXC calibration3 and
(2) the same first order effective area with higher orders modified slightly to improve model
fits to sources with power law spectra (Marshall et al., in preparation). Both sets are con-
sistent with the results of Burwitz et al. (2001) based on the 2000 March observation alone.
Parameters and 1σ statistical uncertainties for best-fit models were: (1) T = 61.2 ± 0.3 eV;
NH = (1.10 ± 0.02) × 1020 cm−2 ; X-ray luminosity (2.96 ± 0.03) × 1031D2
100 erg s−1 , where
D100 is the distance in units of 100 pc; (2) T = 61.1 ± 0.3 eV; NH = (0.81 ± 0.02) × 1020 cm−2 ;
X-ray luminosity (3.16 ± 0.03) × 1031D2
100 erg s−1 . Parameters producing minima in the χ2
test statistic were not sensitive to the exact binning adopted, though of course the reduced χ2
values were: values ranged from 0.94 for data binned to a signal-to-noise ratio of S/N = 10,
to 1.7 for S/N = 30. The latter value is dominated by residual effective area calibration
uncertainties. To investigate the effects of these uncertainties, which are estimated to be
about 15% absolutely, a first order polynomial term was included in the source model to
mimic an effective area lower by 15% at 20A and higher by 15% at 100 A, and vice-versa.
Such a slope skews the blackbody curve and leads systematically to higher or lower temper-
ature solutions by about 1 eV—clearly the true temperature uncertainty is driven by this
uncertainty in the effective area. Allowing for this uncertainty, we adopt a final tempera-
ture of 61.2 ± 1.0—in agreement with, but much more tightly constrained than, the ROSAT
temperatures derived by Burwitz et al. (2001; 63 ± 3 eV) and Pons et al. (2002; 55.3 ± 5.5).
We note that a temperature of 61 eV results in an optical flux only 10% higher than that
from a 55 eV blackbody, so that the discrepency between the observed optical flux and that
predicted by the hot blackbody noted by Walter & Matthews (1997) and Pons et al. (2002)
remains essentially unchanged.
Blackbody models were found by both Pons et al. (2001) and Burwitz et al. (2001) to
represent observed spectra better than sophisticated model atmospheres, though a uniform
3Version dated 2000 October 31; http://asc.harvard.edu/cal/Links/Letg/User
– 7 –
temperature blackbody model was formally excluded in X-ray-EUV-UV-optical modelling
in the former work. However, additional cooler components were found to contribute at
most only a few percent to the observed ROSAT PSPC X-ray flux. We have ruled out
the significant presence of additional thermal and non-thermal emission components by trial
of models combining two blackbodies, and a blackbody component with arbitrary power
laws: in both cases the additional components were completely unconstrained and resulted
in no improvement in the goodness of fit. The 3σ upper limit to power law flux is 5.2 ×
1028D2
100 erg s−1keV−1 at 1 keV. Our blackbody models from methods (1) and (2) correspond
formally to a radius over distance ratio (angular size) of R∞/D100 = 4.12 ± 0.68 km/100pc,
where R∞ = R/p1 − 2GM/Rc2 is the “radiation radius” corresponding to the true radius
R for a star of mass M , and the quoted uncertainty represents the combined temperature
determination uncertainty (±1 eV) and the (dominant) absolute effective area uncertainty
of the LETG+HRC-S combination (±15% ).
In all model comparisons for binning at S/N > 10 the residual differences between the
observed counts and the best-fit model show systematic departures from normal statistical
deviations (Figure 1). Broad deviations are characterised by a general overprediction of
observed counts for λ < 30 A and in the region 75-100 A which is dominated by higher order
flux, underprediction by an average of ∼ 10 % for the range 25-38 A, and deviations around
the instrumental C K edge region 40-44 A. The latter results from a residual calibration error
in the HRC-S UV/Ion shield. Other deviations could arise either as a result of impropriety of
a blackbody model for RX J1856.5−3754, or through calibration errors which are currently
estimated to be ≤ 15% over broad spectral ranges and less over narrower ranges (Drake
et al., in preparation). An apparent edge at 60 A and flux excess in the range 60-70 A
arises because of one of the HRC-S plate gap boundaries coincides with small residual QE
differences between positive and relative negative order outer plates.
In the case of narrow line or edge features, in different combinations of spectral order,
data reduction method and binning size, we identify possible structure in residuals at 26.5,
27.6, 34.4, 32.4, and 35.9 A (emission like features), and at 28.2, 39.1, and 86.5 A (absorption
like features), that might be tempting to attribute to the source. However, we cannot exclude
the possibility that any are chance fluctuations at the 1% level. The issue is complicated
by possible calibration uncertainties on smaller scales, believed to be at a level of about
5% or less, that should be largely smoothed out by dither. We have also used a 60ks
observation (ObsID 331) of PKS 2155−304, currently thought to be a featureless continuum
in the spectral range we are concerned with here (Marshall et al., in preparation), as a
flat field source to aid in feature identification. This spectrum comprises about eight times
the number of first order counts of the combined RX J1856.5−3754 data. Moreover, we
have imposed a constraint that features must appear in both positive and negative orders
– 8 –
in the RX J1856.5−3754 spectrum. We examined deviations from the model on scales up
to 3 A and find only the expected normal distribution of residuals after allowing for smooth
departures resulting from calibration. The Kolmogorov-Smirnov test applied to deviations
on different scales also revealed no evidence for significant features these scales. In summary,
all significant deviations in the residuals that have been found can reasonably be explained
by instrumental effects. Equivalent width upper limits were derived by applying counting
statistics to a convolution of the spectrum with a triangular kernel (Figure 2).
5.
Interstellar Medium Absorption
The distance of 62 pc derived by Walter (2001) seems at odds with the neutral H column
density, NH , of 1020 cm−2 derived from the Chandra spectra: measured NH values for ob jects
at this distance based on different techniques are typically in the range 1018 -1019 cm−2 (e.g.
Fruscione et al. 1994). Walter (2001) indeed remarked on this, citing reddening values
EB−V of up to 0.1 derived by Knude & Høg (1998) in support of the distance. However,
these reddening values show considerable scatter at low reddening and are based only on a
relatively coarse attribution of spectral type to the stars considered.
We have estimated NH and the mean local neutral hydrogen number density, nH , in the
line-of-sight toward RX J1856.5−3754 as a function of distance by spatial interpolation in
the measurements compiled by Fruscione et al. (1994) and Diplas & Savage (1994) using a
technique developed by P. Jelinsky (unpublished). We illustrate this in Figure 3, together
with the allowed distance ranges from Walter (2001) and Kaplan et al. (2002). The value
of NH estimated in this way at a distance of 60pc is an order of magnitude lower than the
X-ray measurement, but is in good agreement with the distance of 140pc derived by the
latter authors. This distance would place RX J1856.5−3754 on the outskirts of the R CrA
cloud using the cloud distance of Knude & Høg (1998) of ∼ 170pc and within the cloud using
the canonical cloud distance of ∼ 130 pc. In either case, RX J1856.5−3754 would likely lie
in a region of relatively high ISM density (∼ 1-10 cm−3 ). While these estimates of NH are
crude and will smooth out any small-scale ISM inhomogeneities, the larger distance is easier
to reconcile with the measured column. The cloud distance of Knude & Høg (1998) then
represents an upper limit to the distance of RX J1856.5−3754.
– 9 –
6. Discussion
Our results, combined with the recent analysis of Ransom et al. (2002), demonstrate a
lack of pulsed features above a level of 2.7 % (unaccelerated search; 4.5% from the accelerated
search of Ransom et al.) and no unequivocal detection of spectral features. This dearth of
indices with which to restrict parameter space precludes an obvious answer to the problem
of the nature and origin of RX J1856.5−3754.
The apparent lack of electron or proton resonance cyclotron absorption suggests that
magnetic field strengths in the ranges (1-7) × 1010 and (0.2-1.3) × 1014 G are less likely,
as discussed by Paerels et al. (2001) for RXJ 0720.4−3125 and by Burwitz et al. (2001),
but, as emphasised by the latter, should not be excluded owing to possible difficulties in
detecting the absorption features.
Indeed, neutron stars with different levels of magnetic
field up to 1015 G have now been observed with high resolution X-ray spectrometers and
none have so far shown absorption features that are intrinsic to the stellar photosphere (e.g.
RXJ 0720.4−3125—Paerels et al. 2001; PSR 0656+14—Marshall & Schulz 2002; Vela—
Pavlov et al. 2001; and 4U 0142+61—Juett et al. 2002).
Our derived angular size based on modelling of the LETGS spectra, R∞/D100 = 4.12 ±
0.68 km/100pc, is consistent with that of Burwitz et al. (2001), though the revised allowed
distance range of 111-200 pc (Kaplan et al. 2002), together with the distance upper limit
constraint based on the R CrA cloud, D100 ≤ 1.70, now implies a radiation radius in the
range R∞ = 3.8-8.2 km. The high end of this range, corresponding to the largest allowed
distance, is still inconsistent with current “normal” NS equations of state (R∞ & 12 km;
e.g. Lattimer & Prakash 2000), as well as with those with extreme softening, such as kaon
condensate models.
Pons et al. (2001) find that heavy element-dominated atmosphere models provide a
plausible match to the low resolution ROSAT PSPC spectra and UV and optical fluxes of
RX J1856.5−3754, while Kaplan et al. (2002) alleviate conflicts with standard EOS through
their revised distance. The heavy element models yield larger radii by virtue of having cooler
effective temperatures than blackbody spectra with similar energy distributions. However,
Burwitz et al. (2001) argued against such uniform temperature heavy element-dominated
atmosphere solutions based on a lack of the expected spectral line features. The appar-
ently featureless but much higher quality LETGS spectra presented here strengthen these
conclusions.
An alternative favoured by Pons et al. (2001), Burwitz et al. (2001) and Ransom et al.
(2002) is a non-uniform temperature model—that we are only seeing a localised hot region
on the surface of a cooler star. The latter authors argue that the gravitational smearing
– 10 –
effects described by Psaltis, Ozel, & DeDeo (2000) account for a lack of observed pulsations.
However, the pulse fraction expectations of Psaltis et al. (2000) indicate X-ray pulse fraction
levels below our 2.7 % limit would be seen only ∼ 10-15 % of the time.
Pulsation would also be expected for a young (∼ 106yr), cooling INS with a strong
magnetic field. The alternative—that RX J1856.5−3754 is an older ob ject re-heated by ISM
accretion—has been dismissed by Kaplan et al. (2002) and van Kerkwijk & Kulkarni (2001a),
largely based on a Bondi-Hoyle accretion rate for the space velocity of Walter (2001) and
Kaplan et al. (2002) that would be much too low to explain the observed luminosity, and
on a possibly strong, accretion-inhibiting magnetic field (Pons et al. 2001). Nevertheless,
RX J1856.5−3754 appears to be in the outskirts of the R CrA cloud and, based on its HST-
derived velocity vector, has likely passed through more dense regions than it now resides
in. Finding one of only a handful of INS candidates in such a region by chance is very
unlikely, yet it is just such dense ISM regions that are expected to power accretion-heated
INSs. Based on the large velocity, however, accretion could only be significant if the Bondi-
Hoyle formalism were to be inapplicable for RX J1856.5−3754. While we tentatively ascribe
the ∼ 2% change in observed 0th order count rate in the 19 months separating the two
observation sets to detector effects, such variability could be accommodated by an accretion
model as the star passed through ISM density fluctuations on ∼AU scales.
The lack of spectral features is also consistent with a pure hydrogen atmosphere model
that is expected to result from modest accretion, whereby heavier elements undergo rapid
gravitational settling. Pons et al. (2001) argue that standard pure H models overpredict the
optical flux of RX J1856.5−3754 by a factor of 30 and that the magnetic accreting models of
Zane, Turolla & Treves (2000), while capable of reproducing the observed X-ray to optical
flux ratio, would need to be two orders of magnitude brighter and an order of magnitude
hotter than observed. However, different accretion scenarios and atmospheric models might
bear examination in light of the otherwise coincidental location of RX J1856.5−3754.
The slightly unfavourable odds of a non-uniform temperature model failing to show signs
of pulsation leads us to consider a third possibility. Taken at face value, the distance, NH ,
and lack of spectral and temporal features and pulsations favour a more compact ob ject than
current NS models permit, but one that is allowed for in strange quark matter solutions (e.g.
Lattimer & Prakash 2000). There now exists a body of evidence from heavy-ion collision
experiments supporting the viability of a quark-gluon plasma (Heinz 2001), and the possible
existence of “strange stars” (e.g. Alcock et al. 1986) is perhaps not as speculative as it once
was. As noted by Pons et al. (2001), such an ob ject would be expected to have a thermal
spectrum as we observe. Such a suggestion is not unprecedented and there now exists a small
handful of ob jects whose apparent compactness could be explained if they are composed of
– 11 –
quark matter (e.g. Cheng et al. 1998; Li et al. 1995; Xu et al. 1999). Of the existing quark
star candidates, RX J1856.5−3754 arguably presents the strongest and most direct case. If
this case survives future scrutiny, then the likelihood of such an ob ject being the brightest
and closest of the current few INS candidates would add some support to speculation that
such a state of matter is a common product of supernovae explosions, or a common phase or
endpoint in the evolution of a neutron star (e.g. Alcock et al. 1986; Kapoor & Shukre 2001;
Xu, Zhang & Qiao 2001).
We extend warm thanks to Pete Ratzlaff for invaluable assistance and “wonder scripts”
developed at short notice. We also thank members of the CfA HEAD for useful comments and
corrections that enabled us to improve the manuscript. The SAO authors were supported by
NASA contract NAS8-39073 to the Chandra X-ray Center during the course of this research.
HLM was supported by NASA contract SAO SV1-61010.
Alcock, C., Farhi, E., & Olinto, A. 1986, ApJ, 310, 261
REFERENCES
Burwitz, V., Zavlin, V. E., Neuhauser, R., Predehl, P., Trumper, J., & Brinkman, A. C.
2001, A&A, 379, L35.
Campana, S., Mereghetti, S., & Sidoli, L. 1997, A&A, 320, 783
Caraveo, P. A., Bignami, G. F., & Trumper, J. E. 1996, A&A Rev., 7, 209
Cheng, K. S., Dai, Z. G., Wei, D. M., & Lu, T. 1998, Science, 280, 407
Diplas, A. & Savage, B. D. 1994, ApJS, 93, 211
Fruscione, A., Hawkins, I., Jelinsky, P., & Wiercigroch, A. 1994, ApJS, 94, 127
Gregory, P. C. & Loredo, T. J. 1992, ApJ, 398, 146
Heinz, U. 2001, Nuclear Physics A, 685, 414
Li, X.-D., Bombaci, I., Dey, M., Dey, J., & van den Heuvel, E. P. J. 1999, Physical Review
Letters, 83, 3776
Juett, A. M., Marshall, H. L., Chakrabarty, D., & Schulz, N. S. 2002, ApJ, 568, L31.
Kaplan, D. L., van Kerkwijk, M. H., & Anderson, J. 2002, ApJ, in press
– 12 –
Kapoor, R. C. & Shukre, C. S. 2001, A&A, 375, 405
Knude, J. & Høg, E. 1998, A&A, 338, 897
Lattimer, J. M. & Prakash, M. 2001, ApJ, 550, 426
Lattimer, J. M. & Prakash, M. 2000, Phys. Rep., 333, 121
Marshall, H.L., & Schulz, N.S. 2002, ApJ, in press
Paerels, F. et al. 2001, A&A, 365, L298
Pavlov, G. G., Zavlin, V. E., Sanwal, D., Burwitz, V., & Garmire, G. P. 2001, ApJ, 552,
L129
Pons, J. A., Walter, F. M., Lattimer, J. M., Prakash, M., Neuhauser, R., & An, P. 2002,
ApJ, 564, 981.
Psaltis, D., Ozel, F., & DeDeo, S. 2000, ApJ, 544, 390
Ransom, S. M., Gaensler, B. M., & Slane, P. O. 2002, ApJ, in press
Shapiro, S. L. & Teukolsky, S. A. 1983, Black Holes, White Dwarfs, and Neutron Stars. New
York: John Wiley and Sons 1983
Treves, A., Turolla, R., Zane, S., & Colpi, M. 2000, PASP, 112, 297
Tsuruta, S. 1998, Phys. Rep., 292, 1
van Kerkwijk, M. H. & Kulkarni, S. R. 2001a, A&A, 380, 221
van Kerkwijk, M. H. & Kulkarni, S. R. 2001b, A&A, 378, 986
Walter, F. M. 2001, ApJ, 549, 433
Walter, F. M. & Matthews, L. D. 1997, Nature, 389, 358
Walter, F. M., Wolk, S. J., & Neuhauser, R. 1996, Nature, 379, 233
Xu, R. X., Qiao, G. J., & Zhang, B. 1999, ApJ, 522, L109
Xu, R. X., Zhang, B., & Qiao, G. J. 2001, Astroparticle Physics, 15, 101
Zane, S., Turolla, R., & Treves, A. 2000, ApJ, 537, 387
This preprint was prepared with the AAS LATEX macros v5.0.
– 13 –
Fig. 1.— The combined positive and negative order spectra of RX J1856.5−3754 binned at
0.5 A intervals shown with the best fit blackbody model with parameters corresponding to
method (2) in §4 and residuals (observations−model). The deviations from this model are
consistent with Poisson statistics after allowing for calibration uncertainties at the C K-edge
and over broader wavelength intervals. The apparent edge at 60 A results primarily from
one of the HRC-S plate gap boundaries and small residual QE differences between positive
and relative negative order outer plates.
– 14 –
Fig. 2.— The 3σ equivalent width upper limit to line features as a function of wavelength
based on a convolution of the spectrum with a triangular kernel.
– 15 –
Fig. 3.— The estimated NH (solid curves) and nH (dotted curves) as a function of distance in
the line-of-sight toward RX J1856.5−3754. The pairs of curves represent the likely ranges of
these quantities and correspond to deviations of factors of 3 from a smooth locus through the
results of the 3D interpolations. Vertical shaded regions indicate the distance ranges found
by Walter (2001; W) and Kaplan et al. (2002; KvKA). The horizontal stripe illustrates the
allowed range (±1σ) of NH derived in this study.
– 16 –
Table 1. Summary of Chandra LETG+HRC-S Observations
Obs ID
UT Start
UT End
Exposure
[s]
Net Eventsa
0th + 1st
0th
0th Rateb
[Hz]
113
3382
3380
3381
3399
2000-03-10 07:55:12
2001-10-08 08:18:49
2001-10-10 05:06:28
2001-10-12 19:19:26
2001-10-15 11:47:06
2000-03-10 23:37:24
2001-10-09 03:01:50
2001-10-12 04:00:48
2001-10-14 09:14:28
2001-10-15 14:42:59
55121
101172
166325
169956
9282
12202
20949
35097
36011
1962
— 0.2195 ± 0.0020
0.2166 ± 0.0016
86516
0.2154 ± 0.0013
135230
141349
0.2157 ± 0.0013
0.2126 ± 0.0051
7136
aThe net event numbers includes background events; while including the 1st order events increases the
net source events, the number of background events also increases.
bRates were derived from 0th order data when valid event rates did not exceed 184 count s−1 ; this
selection criterion yielded 417786 seconds of the 501856s total exposure time.
|
astro-ph/0111323 | 1 | 0111 | 2001-11-16T10:18:23 | CN and HNC Line Emission in IR Luminous Galaxies | [
"astro-ph"
] | We have observed HNC 1-0, CN 1-0 and 2-1 line emission in a sample of 13 IR luminous (LIRGs, L_IR > 10E11 Lo) starburst and Seyfert galaxies. HNC 1-0 is detected in 9, CN 1-0 is detected in 10 and CN 2-1 in 7 of the galaxies. We also report the first detection of HC3N (10-9) emission in Arp220. The excitation of HNC and CN emission requires densities n > 10E4 cm-3. We compare their intensities to that of the usual high density tracer HCN. The I(HCN)/I(HNC}) and I(HCN)/I(CN) 1-0 line intensity ratios vary significantly, from 0.5 to >6, among the galaxies. This implies that the actual properties of the dense gas is varying among galaxies who otherwise have similar I(CO)/I(HCN) line intensity ratios. We suggest that the HNC emission is not a reliable tracer of cold (10 K) gas at the center of LIRGs, as it often is in the disk of the Milky Way. Instead, the HNC abundance may remain substantial, despite high gas temperatures, because the emission is emerging from regions where the HCN and HNC formation and destruction processes are dominated by ion-neutral reactions which are not strongly dependent on kinetic temperature. We find five galaxies (four AGNs and one starburst) where the I(HCN)/I(HNC) intensity ratio is close to unity. In other AGNs, however, I(HCN)/I(HNC}) is >4. The CN emission is on average a factor of two fainter than HCN, but the variation is large and there seems to be a trend of reduced relative CN luminosity with increasing IR luminosity. One galaxy, NGC3690, has a CN luminosity twice that of HCN and its ISM is thus strongly affected by UV radiation. We discuss the I(HCN)/I(HNC) and I(HCN)/I(CN) line ratios as indicators of starburst evolution. | astro-ph | astro-ph |
Astronomy & Astrophysics manuscript no.
(will be inserted by hand later)
CN and HNC Line Emission in IR Luminous Galaxies
S. Aalto1, A. G. Polatidis1, S. Huttemeister1
,
2, S. J. Curran3
1 Onsala Rymdobservatorium, Chalmers Tekniska Hogskola, S - 439 92 Onsala, Sweden
2 Astronomisches Institut der Universitat Bochum, Universitatsstrasse 150, D - 44780 Bochum, Germany
3 School of Physics, University of New South Wales, Sydney NSW 2052, Australia
Received / Accepted
I(HNC) 1-0 and I(HCN)
Abstract. We have observed HNC 1-0, CN 1-0 & 2-1 line emission in a sample of 13 IR luminous (LIRGs,
LIR > 1011 L⊙) starburst and Seyfert galaxies. HNC 1-0 is detected in 9, CN 1-0 is detected in 10 and CN 2-1 in 7
of the galaxies and all are new detections. We also report the first detection of HC3N (10-9) emission in Arp 220.
The excitation of HNC and CN emission requires densities n > 104 cm−3. We compare their intensities to that of
the usual high density tracer HCN. The I(HCN)
I(CN) 1-0 line intensity ratios vary significantly, from
∼ 6, among the galaxies. This implies that the actual properties of the dense gas is varying among galaxies
0.5 to >
who otherwise have similar I(CO)
I(HCN) line intensity ratios. We suggest that the HNC emission is not a reliable tracer
of cold (10 K) gas at the center of LIR galaxies, as it often is in the disk of the Milky Way. Instead, the HNC
abundance may remain substantial, despite high gas temperatures, because the emission is emerging from regions
where the HCN and HNC formation and destruction processes are dominated by ion-neutral reactions which are
not strongly dependent on kinetic temperature. We find five galaxies (Mrk 231, NGC 7469, NGC 7130, IC 694
and NGC 2623) where the I(HCN)
I(HNC) intensity ratio is close to unity. Four are classified as active galaxies and one
as a starburst. In other active galaxies, however, the I(HCN)
I(HNC) is > 4. The CN emission is on average a factor of
two fainter than the HCN for the luminous IR galaxies, but the variation is large and there seems to be a trend
of reduced relative CN luminosity with increasing IR luminosity. This trend is discussed in terms of other PDR
tracers such as the [C II] 158 µm line emission. One object, NGC 3690, has a CN luminosity twice that of HCN
and its ISM is thus strongly affected by UV radiation. We discuss the I(HCN)
line ratios as indicators
of starburst evolution. However, faint HNC emission is expected both in a shock dominated ISM as well as for a
cloud ensemble dominated by dense warm gas in the very early stages of a starburst. Additional information will
help resolve the dichotomy.
I(HNC) and I(HCN)
I(CN)
Key words. galaxies: evolution -- galaxies: ISM -- galaxies: starburst -- radio lines: galaxies -- radio lines: ISM
1. Introduction
The polar molecule HCN (dipole moment 2.98 debye) is
commonly used as a tracer of dense molecular gas, i.e. gas
at n(H2) ≥ 104 cm−3. In particular in distant luminous
(LIR > 1011 L⊙, LIRGs) and ultraluminous (LIR > 1012
L⊙, ULIRGs) systems the HCN 1 -- 0 line is the prototypi-
cal tracer of dense gas content (e.g., Solomon et al. 1992;
Helfer & Blitz 1993; Curran et al. 2000 (CAB)). Solomon
et al. (1992) find a tighter correlation between FIR and
HCN luminosity than the one found between FIR and CO.
They suggest that, in general, the IR luminosities origi-
nate from star formation rather than AGN activity in FIR
luminous galaxies. The HCN to CO intensity ratio, how-
ever, varies substantially ( 1
40 ) among luminous galax-
ies, and it is unclear whether this difference can simply
be interpreted as variations in the dense gas content or
3 - 1
Send offprint requests to: S. Aalto
is also due to abundance and/or excitation effects. Apart
from being collisionally excited, HCN may become excited
via electron collisions (at X(e) ≈ 10−5) or be pumped
by 14 µm continuum radiation through vibrational tran-
sitions in its degenerate bending mode. It is also difficult
to know if the gas is really engaged in star formation, or
if it is simply dense in response to being near the central
potential of the galaxy (e.g., Helfer & Blitz 1993; Aalto et
al. 1995) where other mechanisms (AGN, turbulence etc)
may heat the gas and dust.
In order to understand the activities in the centers of
luminous galaxies it is essential to also understand the
prevailing conditions of the dense gas. Apart from ob-
serving higher transitions of HCN it is important also
to study the emission from other high density tracers.
One such tracer is the HNC molecule, the isomer of (and
chemically linked to) HCN. For example, at high temper-
atures HNC can be transferred into HCN via the reaction
2
S. Aalto, A. G. Polatidis, S. Huttemeister, S. J. Curran: CN,HNC in Galaxies
HNC + H → HCN + H. It is predicted, e.g. in (maybe
oversimplified) chemical steady state models, but also by
shock models, that the HCN
HNC ratio increases with increas-
ing temperature and gas density (e.g., Schilke et al. 1992
(S92)). This is supported by the fact that the measured
HCN
HNC abundance ratio is especially high in the vicinity of
the hot core of Orion KL. Most of the temperature de-
pendence is between 10 and 50 K, after which there is a
considerable flattening (S92).
Compared to these results, the HCN
HNC intensity ratios
found (so far) in nearby starburst galaxies are rather low
(ranging from 1-5) closer to dark clouds than to hot cores
(Huttemeister et al. 1995 (H95)). This result is in apparent
∼ 50 K)
contradiction with the idea that the gas is warm (T >
in the centers of starburst galaxies (e.g., Wild et al. 1992;
Wall et al.1993). However, Aalto et al. (1995) suggest that
the dense cores of the molecular clouds of the starburst
NGC 1808 are cold (10 K) and thus these cores could
be responsible for the HNC emission in NGC 1808, but
possibly also in other galaxies.
The radical CN is another tracer of dense gas, with
a somewhat lower (by a factor of 5) critical density
than HCN. Observations of the CN emission towards the
Orion A molecular complex (Rodriguez-Franco et al. 1998)
show that the morphology of the CN emission is dom-
inated by the ionization fronts of the HII regions. The
authors conclude that this molecule is an excellent tracer
of regions affected by UV radiation. Thus, the emission
from the CN molecule should serve as a measure of the
relative importance of gas in Photon Dominated Regions
(PDRs).
>
We have searched for HNC and CN emission in
a sample of LIRG and ULIRG galaxies with warm
( f (60 µm)
∼ 0.75) FIR colours. We were interested to see
f (100 µm)
whether the HNC emission would be relatively fainter
compared to the cooler, nearby objects studied by H95. Is
HNC a reliable cold gas tracer, or would we find evidence
for the contrary? We furthermore wanted to assess the rel-
ative importance of dense PDRs in these objects through
comparing the CN line brightness with that of HCN. If in-
deed the HNC emission is a tracer of the amount of cold,
dense gas, then perhaps an anti-correlation between the
CN and HNC emission is to be expected. Many of the
galaxies in the survey are powered by prodigious rates of
star formation and thus a bright CN line relative to that
HCN is to be expected. Some of the galaxies are domi-
nated by an AGN where the expected CN brightness may
also be high (e.g., Krolik & Kallman 1983).
In Sect. 2, we present the observations and in Sect. 3
the results in terms of line intensities and line ratios. In
Sect. 4.1 we discuss the interpretation of the HNC results
and in Sect. 4.2 we discuss CN. In 4.3 possible connec-
tions to starburst evolution and scenarios of the possibly
dominating gas components are discussed.
Table 1. Beamsizes and efficiencies
Transition
ν [GHz]
HPBW [′′]
ηmb
OSO SEST OSO SEST
HCN 1-0
HNC 1-0
CN 1-0
CO 1-0
88.6
90.6
113.5
115.2
44
42
34
33
57
55
46
45
0.59
0.59
0.50
0.50
0.75
0.75
0.70
0.70
2. Observations
I(CO)
I(HCN) 1-0 intensity ratios <
We have used the SEST and OSO 20m to measure the
HNC 1-0 (90.663 GHz) and the CN 1-0 113.491 GHz (1-
0, J=3/2-1/2, F =5/2-3/2) line intensity in a selection
of 13 LIRGs and ULIRGs. We also include observations
of NGC 1808 which is of lower luminosity. The selected
∼ 15
galaxies all have global
(apart from NGC 3256 and NGC 1808). For the south-
ern galaxies observed with SEST, we were also able to
measure the CN 2-1 line (226.874 GHz (2-1, J=5/2-3/2,
F =7/2-5/2), 226.659 GHz (2-1 J=3/2-1/2 F=5/2-3/2)).
The CN 1-0 113.191 GHz line (1-0 J=1/2-1/2 F=3/2-3/2)
is shifted +806 km s−1 from the main line and we have ob-
tained limits to its intensity in several cases. For Arp 220
the two CN 1-0 spingroups are blended because the line
is so broad. Thus, even in the 1 GHz correlator backend
(see below) it was necessary to observe CN 1-0 at two dif-
ferent LO settings and then join the spectra together to
get enough baseline. For four galaxies the bandwidth was
wide enough to also include the 90.983 GHz 10-9 line of
HC3N.
Observations were made in 1999 October (HNC,
OSO), December (HNC, SEST) and 2000 June (CN,
OSO), August (CN, SEST). For OSO, the system tem-
peratures were typically 300 K for HNC and 500-600
K for CN. For SEST, typical system temperatures were
230 K for the HNC measurements and 400 K for both the
113 GHz and the 226 GHz CN observations. Pointing was
checked regularly on SiO masers and the rms was found
to be 2′′ for OSO, and 3′′ for SEST. Arp 220 was observed
both with OSO (CN) and SEST (HNC). We have also
measured the 115 GHz CO 1-0 and the HCN 1-0 lines for
some galaxies where we did not have values from the lit-
erature. Beamsizes and efficiencies are shown in Table 1.
For the OSO observations a 500 MHz filterbank was used
for backends for all observations, and for some a 1 GHz
autocorrelator was also used. For the SEST observations
we alternated between a 500 MHz and 1 GHz backend de-
pending on whether simultaneous observations with the 1
and 3 mm receiver were taking place. We used the software
package xs (written by P. Bergman) to subtract baselines
and add spectra.
3. Results
S. Aalto, A. G. Polatidis, S. Huttemeister, S. J. Curran: CN,HNC in Galaxies
3
Table 2. Integrated Line Intensitiesa
Galaxy
Telescope
vc
I(HNC) 1-0b
I(CN) 1-0c
I(CN)2-1d
5/2 - 3/2
3/2 - 1/2
I(CO) 1-0
km s−1
K km s−1
K km s−1
K km s−1
K km s−1
K km s−1
K km s−1
Arp 220
"
IC 694f
NGC 3690f
Mrk 231
Mrk 273
NGC 34
NGC 1614
NGC 2146
NGC 2623
NGC 6240
NGC 3256
NGC 7130
NGC 7469
OSO
SEST
OSO
OSO
OSO
OSO
SEST
SEST
OSO
OSO
SEST
SEST
SEST
OSO
5400
5400
3100
3050
12650
11320
5930
4500
900
5500
7335
2800
4840
4960
. . .
0.95 ± 0.2
0.75 ± 0.2
. . .
0.5 ± 0.1
2.0 ± 0.3
. . .
<
∼ 0.85
1.5 ± 0.3
0.5 ± 0.2:
<
<
∼ 0.25
∼ 0.3
. . .
. . .
0.6 ± 0.15
0.4 ± 0.2:
0.6 ± 0.05
0.4 ± 0.05
0.7 ± 0.1
<
∼ 0.25
∼ 0.3
<
0.6 ± 0.3:
1.2 ± 0.3
<
∼ 0.6
0.7 ± 0.1
1.2 ± 0.05
0.5 ± 0.1
0.8 ± 0.1
. . .
0.65 ± 0.2h
. . .
. . .
. . .
. . .
0.25 ± 0.05
0.5 ± 0.1
. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
1.3 ± 0.1
0.55 ± 0.05
0.45 ± 0.05
. . .
0.9 ± 0.1e
0.4 ± 0.05
0.3 ± 0.05
. . .
0.4 ± 0.1e
0.15 ± 0.05
0.15 ± 0.05
. . .
15.0 ± 2.0g
13.0 ± 2.0
14.0 ± 1.0
11.0 ± 1.0
g
g
5.5 ± 0.5g
9.0 ± 0.5
43.0 ± 1.0
13.0 ± 0.5
15.5 ± 0.5
45 ± 1.5
9.5 ± 0.5g
g
NGC 1808
SEST
960
1.2 ± 0.1
3.8 ± 0.1
2.0 ± 0.1
1.4 ± 0.1
0.7 ± 0.1
66 ± 0.5
A and in K km s−1, and upper limits are all 3σ. SEST, Tmb: 27 Jy K−1 (115 GHz), 41 Jy
a): Integrated line intensities are in T ∗
K−1 (230 GHz) for point sources; OSO, Tmb: 27 Jy K−1 (115 GHz) for point sources.
b): In Arp 220 we also detect the 90.98 GHz (10-9) line of HC3N with the intensity 0.4 ± 0.15. For NGC 3256 the (3σ) limit to
the line is 0.18, for NGC 7130 it is 0.15, and for NGC 1808 we have a very tentative detection with the intensity 0.2 ± 0.1. For
Arp 220 we also find an integrated HCN intensity in the SEST beam of 2.5 ± 0.4.
c): This is only the integrated intensity of the 1-0 J=3/2-1/2 line. To get the total 1-0 intensity multiply with 1.3 (in the
optically thin case). For Arp 220 both lines are blended and included in the integrated line intensity listed here.
d): The total integrated intensity from the two 5/2 - 3/2 and 3/2 - 1/2 spingroups together. The following two columns contain
the integrated intensity in each spin group whenever it was possible to measure.
e): Line blending is severe therefore the intensity ratio was estimated from fitting of two gaussians with fixed linewidths of
∆V = 450 km s−1 at v1 = 7200 km s−1and v2 = 7500 km s−1.
f): IC 694 and NGC 3690 are also known as the merger Arp 299.
g): See CAB.
h): very broad lines, may be affected by unknown baseline errors.
3.1. HNC line intensities and ratios
The line intensities are presented in Table 2 and the ra-
tios in Table 3. CO spectra are presented in Figure 1 and
HNC spectra (plus a SEST HCN spectrum for Arp 220)
in Figure 2. For some of the galaxies the center veloc-
ity vc is put to 0. This is because the observer in that
case chose to set the velcocity to zero and work with the
redshifted frequency instead. The velocity used to calcu-
late the redshifted frequency (thus the velocity the band
was centered on) can be found in Table 2. All galaxies,
∼ 1011 L⊙
except NGC 1808, have FIR luminosities LFIR
(see Table 3). Five of the investigated sources (Mrk 231,
NGC 7469, NGC 2623, IC 694 and NGC 7130) have global
HCN
HNC luminosity ratios close to unity. The rest have ra-
tios ranging from 2 to >
∼ 6. The HNC luminous objects
are all AGNs (three Seyferts, one LINER (NGC 2623))
except for IC 694 which we suspect is a starburst (e.g.,
Polatidis & Aalto 2000), but there are also Seyfert galax-
ies with faint HNC emission (such as Mrk 273, NGC 6240
and NGC 34). We do not find that an increase in FIR lu-
minosity is followed by an increase in HCN
HNC line intensity
>
ratio. Instead, there might be a weak trend to the opposite
and two (Mrk 231 and Arp 220) of the three ULIRGs in
our sample have relatively bright HNC emission (see Sect.
3.1.1).
3.1.1. Comparisons with other HNC surveys
In the H95 HNC survey of nearby starburst galaxies, the
majority of the sources show HCN
HNC line ratios greater than
or equal to two. Three objects have I(HCN)≈ I(HNC):
the nearby starburst NGC 253, the nearby post-starburst
NGC 7331 and the Seyfert NGC 3079. For most of the
galaxies in that sample, the line ratios are not global, but
reflect the conditions in the inner 0.2 - 1 kpc of the galaxy.
Since both the HCN and CO emission for most of our
sample galaxies comes from the inner kpc (e.g., Bryant
1996; Scoville et al. 1997; Downes & Solomon 1998; Bryant
& Scoville 1999), it is meaningful to compare the HCN
HNC
ratios of the two samples. In H95 the average HCN
HNC line
intensity ratio is 2 for 14 galaxies (excluding limits and
their value for Arp 220 (see below)). In our sample, the
ratio is somewhat lower, 1.6, when only detections are
4
S. Aalto, A. G. Polatidis, S. Huttemeister, S. J. Curran: CN,HNC in Galaxies
Table 3. Line Ratios and FIR luminosities
Galaxy
typea
CO
CN 1-0
b
CO
HCN
HCN
CN
HCN
HNC
HNC
CN
CN 5/2−1/2
3/2−1/2 CN 2−1
1−0
c
log L(FIR)d
Arp 220
IC 694
NGC 3690
Mrk 231
Mrk 273
NGC 34
NGC 1614
NGC 2146
NGC 2623
NGC 6240
NGC 3256
NGC 7130
NGC 7469
HII
HII
HII
Sey 1
Sey 2
Sey 2/HII
HII
HII
LINER
Seyfert 2
HII
Sey 2/HII
Sey 1/HII
>
>
>
7.5
∼ 17
6
8
∼ 10
∼ 17
12
28
∼ 22
18
29
14
11
>
NGC 1808f HII
13
8
11
13
5
2g
4
14
16
12
9
16
13
8h
18
1
>
>
>
∼ 1.2
0.5
1.6
∼ 5
∼ 4
1
1.8
∼ 2
2
2
1.1
1.4
>
>
>
1.6
1.0e
. . .
1
∼ 4
∼ 4
. . .
. . .
1.4
2.7
3
1.2
1.2
>
0.5
∼ 1
. . .
1.6
. . .
. . .
. . .
. . .
∼ 1.3
0.7
0.7
1
1
>
0.7
2
0.4
. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
2.1
2.6
2.4
. . .
2.2
0.15
. . .
. . .
. . .
. . .
>
∼ 0.35
0.35
. . .
. . .
0.6
0.14
0.25
. . .
0.25
12.13
11.77
. . .
12.37
12.04
11.16
11.43
11.00
11.49
11.69
11.52
11.26
11.41
10.48
a): Type of central activity that is suggested to dominate the IR luminosity.
b): The HCN data come from CAB apart from for: IC 694 & NGC 3690 (Arp 299) (Solomon et al. 1992); NGC 1614 (Aalto et
al. 1995); NGC 2623, NGC 6240 (Bryant 1996); NGC 3256 (Aalto et al. 1995). The Bryant (1996) ratios are measured with the
OVRO interferometer, but since both the CO and HCN emission is compact for both NGC 2623 and NGC 6240 the line ratios
are global and can be compared with single dish ratios. For Arp 299 (IC 694 & NGC 3690) the situation is more complicated
since the high density tracer emission is emerging from compact structures, while a fraction of the CO emission comes from
more extended gas (e.g., Aalto et al. 1997). In (Solomon et al. 1992) HCN was measured in a 28′′ beam (IRAM).
c): We have not mapped the CN line emission which is of course necessary for an accurate determination of the CN 2−1
1−0 line
ratio. For most galaxies we expect the CN distribution to be effectively a point source in the beam -- except for NGC 1808
where we adopt a source size of 18′′ (Aalto et al. 1995).
d): The FIR luminosities come from Bryant (1996) (apart from NGC 1808 (Aalto et al. 1994); NGC 34 (CAB); NGC 3256,
NGC 7130, NGC 2146 (Aalto et al. 1991))
e): We have assumed that the HNC emission is a point source (see footnote b) and recalculated the line intensity to that of the
28′′ IRAM HCN beam.
f): Line ratios are not global, but apply only for the inner 2-3 KPC.
g): We use the temperature line ratio of 2 instead of the integrated line ratio of 1 given in CAB (because the HCN line appear
too broad).
h): There is an error in table 4 of CAB. The correct ratio (8) can be found in table 3 in CAB.
included (but when limits are included the ratio increases
to 2). Joining the H95 galaxies with ours in one sample
we still can find no strong trend in the line ratio with
increasing FIR luminosity. We note, however, that we have
a larger number of galaxies with ratios close to unity (6)
compared to H95 (3) despite our smaller sample.
We observed HNC in Arp 220 to see if we could repro-
duce the remarkable result in H95 that HNC was brighter
than HCN. We observed HNC on two consecutive days,
with observations of HCN in between, but could not con-
firm the bright HNC emission found earlier. Instead we
obtain an HCN
HNC intensity ratio of 1.4, which is more con-
sistent with results found in other galaxies.
3.1.2. HC3N
For some galaxies the bandwidth was large enough to in-
clude the HC3N 10-9 line, shifted in velocity by -1000
km s−1 from the HNC line. The line is detected in Arp 220
at 40% of the HNC 1-0 intensity. In NGC 1808 the line is
tentatively detected at 16% of the HNC 1-0 intensity and
in NGC 7130 and NGC 3256 we have upper limits to the
line (see footnote to Table 2).
3.2. CN line ratios and intensities
CN 1-0: The line intensities are presented in Table 2 and
the ratios in Table 3. The CN 1-0 spectra are displayed in
Figure 3. For most of the galaxies we only detect one of
the spingroups in the 1-0 transition. Because of the broad
line in Arp 220 the two groups are however blended --
even though they are separated by 800 km s−1. Also for
NGC 6240 the second spingroup is somewhat blended with
the first one.
The HCN
CN 1-0 integrated line ratio varies substan-
tially: from 0.5 to greater than four. I(CN) >
∼ I(HCN) in
three objects (Arp 220, NGC 3690 and NGC 1808), while
CN remains undetected in IC 694, Mrk 273, NGC 34
and NGC 2623. We also find three galaxies (Mrk 231,
NGC 2623, and IC 694) where I(CN) <
∼ I(HNC). In four
galaxies (Arp 220, NGC 1808, NGC 3256, NGC 6240)
I(HNC) < I(CN) -- three are starburst galaxies and
S. Aalto, A. G. Polatidis, S. Huttemeister, S. J. Curran: CN,HNC in Galaxies
5
NGC 6240 is classified as a Seyfert. In H95, CN was
brighter than HNC in all four sample galaxies where CN
was measured.
The HCN
CN line intensity ratio seems to increase slightly
with FIR luminosity. Dividing our galaxies into two lumi-
nosity bins we find that the average ratio of 1.5 for the
lower luminosity galaxies and 2.3 for the higher luminos-
ity objects. This, somewhat surprising, result is discussed
in Sect. 4.2. We included NGC 3690 in the lower lumi-
nosity batch since most of the FIR emission is believed to
originate in IC 694.
CN 2-1: The 2-1 spectra are presented in Figure 4.
Both 2-1 spingroups are detected in four of the cases (in
NGC 6240 the blending is severe). In NGC 34 and in
NGC 1614 we detect only the brighter J=5/2-3/2 tran-
sition. The CN 2 -- 1 spectrum of Arp 220 shows a much
weaker line than the 1 -- 0 spectrum, and emission is only
detected in the lower velocity part (vc ≈ 5200 km s−1)
of the spectrum. Also the CN 1-0 line peaks around 5200
km s−1. The CO emission peaks at vc ≈ 5400 km s−1. This
implies that most of the CN emission emerges from the
western nucleus (see Scoville et al. 1997). The HC3N line
on the other hand appears to peak at vc = 5550 km s−1,
and thus most of the emission likely emerges from the
eastern nucleus. This tentative velocity difference should
be investigated at higher resolution.
3.2.1. The CN 2−1
1−0 ratio
The CN 2−1
1−0 line intensity ratio (see footnote to Table 3
on how it was calculated) suggests the CN emission be-
ing subthermally excited apart from in NGC 6240 (and
NGC 34 with an upper limit to the CN 1-0 emission). In
general the ratio is lower than 0.4 indicating gas densities
∼ 5 × 104 cm−3 (Fuente et al. 1995) which are about one
order of magnitude below the critical density.
<
3.2.2. The CN spingroup ratio
We can use the 1mm spingroup ratio to estimate the av-
erage optical depth of the two lines. The ratio between
the lines is quite large, >2 for all cases where it can be
measured. The spingroup ratios are very accurate since
they were obtained at the same time (apart from possi-
ble baseline errors). In NGC 6240 the line blending is so
severe that the ratio is difficult to determine. The LTE ra-
tio should be close to 1.8 (Bachiller et al. 1997) and ratios
around 2 thus indicate that the lines are of low optical
depth. For most galaxies, the other spingroup of the 1-0
line (around 113.191 GHz (J=1/2-1/2)) was outside the
observed bandwidth. In the optically thin case, this line
is about 1/3 of the J=3/2-1/2 at 113.490. For Arp 220,
the two lines are blended (see above) and we fitted two
gaussians each centered on one of the spingroups. The fits
shows that a line intensity ratio between the two lines of
1/3 is possible.
4. Discussion
We initially expected that the relative HNC luminosity
would be lower in our sample of warm, luminous galaxies
compared to the H95 sample. Instead (see Sect. 3.1.1) the
very luminous galaxies even seemed to be somewhat more
luminous, on average, in HNC even though this change is
not statistically significant. A possibility could be that the
telescope beam is picking up more extended, cooler gas in
the distant, more luminous, galaxies. However, most of
them are known to have very compact molecular cloud
distributions and it is unlikely that significant HNC emis-
sion emerges from the outskirts of the galaxies. Our results
seem to challenge the notion of HNC as a reliable tracer
of cold gas.
Furthermore, the variation in HCN
HNC line ratio is large
among the galaxies that otherwise have similar properties.
For example, despite having similar CO
HCN 1-0 intensity ra-
tios, Mrk 273 and Mrk 231 have quite different HCN
HNC and
CN line intensity ratios. The HCN
HCN
HNC ratio of Mrk 231 is
close to unity, while HNC is not detected in Mrk 273 re-
sulting in a HCN
∼ 6. It is unlikely that the dense gas
is cold in Mrk 231 but warm in Mrk 273 (both are hot,
ultraluminous AGN/Starburst mergers) and therefore the
interpretation of HNC needs to be reevaluated.
HNC ratio >
4.1. Why is HNC often bright in centers of galaxies?
There are a number of possible explanations. Let us ex-
amine them one by one and see how they can be tested:
1. The dense gas is cold.
Is it possible that a significant fraction of the dense gas
is cold in centers of starburst galaxies? Possibly. Aalto et
al. (1994) discuss the presence of cold dense gas in the
"mild" starburst galaxy NGC 1808. The high dust tem-
peratures observed towards NGC 1808 could be explained
by clouds with hot surfaces and cold interiors. In this sce-
nario, the HNC 1-0 emission would emerge from the cold
cloud cores, while HCN 1-0 emission also would come from
the outer, warmer parts. Gas at densities > 105 cm−3
should become thermalized with the dust. Thus, if the
HNC emission is coming from dense, cold (10-20 K) gas
there should be submm dust continuum emission associ-
ated with it. Therefore, a study of the submm and mm
excess in conjunction with the strength of the HNC emis-
sion should be quite interesting. However, the most HNC
luminous object in our sample, Mrk 213, shows very weak
mm thermal dust emission and Braine & Dumke (1998)
find a dust temperature of 70 K -- the kinetic tempera-
ture of the associated dense gas should be at least as high.
This is not consistent with the idea that the HNC emis-
sion arises from a cold component. Based on ISOPHOT
data Klaas et al. (1997) find that the bulk of the dust
mass in Arp 220 is at a temperature of about 50 K. Thus,
the bright HNC 1-0 emission is unlikely originate in cold
cloud cores.
We conclude that for objects like Mrk 231 and Arp 220,
where there is no mm excess emission, and where the over-
6
S. Aalto, A. G. Polatidis, S. Huttemeister, S. J. Curran: CN,HNC in Galaxies
Arp 220
NGC 34
IC 694
NGC 1614
NGC 1808
NGC 2146
NGC 2623
NGC 3256
NGC 3690
NGC 6240
NGC 7130
Fig. 1. CO spectra of all galaxies, apart from MRK 231, MRK 273 and NGC 7469 which can be found in CAB. The
A. The velocity resolution ranges from 10 TO 50 km s−1. It was selected to be 10% of the FWHM of the
scale is in T ∗
line itself.
all dust temperature is high, HNC 1-0 does not trace cold
dense gas. Instead one or several of the scenarios below
must apply. However, for less extreme objects such as
NGC 1808 there could be enough mass in cold cores that
at least a fraction of the HNC emission could emerge from
them.
2. Chemistry.
Steady state chemistry models by S92 show both a
temperature and a density dependence in the [HCN]
[HNC] abun-
dance ratio. For example, there is a very significant differ-
ence between n(H2)=104 cm−3 and 107 cm−3. For Tkin=
50 K, [HCN]
[HNC] =67 for 107. If the bulk
[HNC] =0.8 for 104 and [HCN]
of the HCN and HNC emission is emerging from gas of
densities 104 −105 then the relative HNC abundance there
may be substantial, despite the high temperature. The
reason for this is that at lower densities reactions with
HCNH+ (HCN and HNC reacts with H+
3 to form HCNH+)
become more important. The ion abundance is higher and
once HCN and HNC become protonised, HCNH+ will re-
combine to produce either HCN or HNC with 50% prob-
ability. At higher densities, the ion abundance is likely
lower and reactions like HNC + O → CO + NH become
more important at high temperatures. This scenario is in-
teresting since the electron and ion abundance is likely
HC N3
cV
HC N3
S. Aalto, A. G. Polatidis, S. Huttemeister, S. J. Curran: CN,HNC in Galaxies
7
Arp 220
HNC
IC 694
Mrk 231
NGC 1808
HNC
NGC 2623
NGC 3256
NGC 6240
NGC 7130
NGC 7469
Fig. 2. HNC 1-0 spectra for the galaxies for which we claim a detection. The scale is in T ∗
A. The 10-9 HC3N line
is indicated in the spectra for Arp 220 and NGC 1808. The velocity resolution ranges from 10 to 50 km s−1. It was
selected to be 10% of the FWHM of the line itself, but for a few cases the S/N of the spectrum required further
smoothing.
higher in PDRs. Therefore, in a PDR chemistry, the con-
nection between HNC abundance and kinetic temperature
may also be weak since we there expect the HCNH+ reac-
tions to be important. Since the CN 2−1
1−0 ratios we measure
indicate subthermal excitation it is reasonable to assume
that most of the HCN,HNC and CN emission is indeed
emerging from gas where the density is below 105 cm−3.
3. Optical depth effects
In Orion, S92 find peak-to-peak intensity ratios be-
tween HCN
HNC of 3 to 4 towards the hot core and ridge.
However, the abundance ratio is much higher, ≈80. Thus,
it is possible that the fairly bright HNC emission in some
galaxy centers is caused by optical depth effects. In objects
where the HCN emission is subthermally excited (like in
Mrk 231) the optical depth of the 1-0 line could be quite
high and perhaps explain part of the apparently too-bright
HNC.
4. IR pumping
Both HCN and HNC may be pumped by an intense
mid-IR radiation field boosting the emission from low den-
sity regions. There has been no direct evidence IR pump-
ing is dominating the HCN excitation in external galax-
ies. However, Barvainis et al. (1997) suggest IR pump-
ing as a possible mechanism behind the HCN emission
of the Cloverleaf quasar. Ultraluminous galaxies, such as
Mrk 231 and Arp 220, have central mid-IR sources with
optically thick radiation temperatures well in excess of
those necessary to pump the HCN molecule (Soifer et al.
1999). For HNC the coupling to the field is even stronger
than for HCN, thus increasing the probability for IR
pumping in extreme galaxies, such as Mrk 231. Even if
the HNC abundance is lower compared to HCN the HNC
emission may have a higher filling factor due to the IR
pumping (if it allows emission from gas clouds otherwise
at too low density to excite the HNC molecule). A com-
parative excitation study of HCN and HNC would help
8
S. Aalto, A. G. Polatidis, S. Huttemeister, S. J. Curran: CN,HNC in Galaxies
Arp 220
Mrk 231
NGC 1614
NGC 1808
NGC 2146
NGC 3256
NGC 3690
NGC 6240
NGC 7130
NGC 7469
Fig. 3. CN 1-0 spectra for the galaxies with detections. The scale is in T ∗
A. The spectrum for Arp 220 shows the
Gaussian fit for both spingroups, since the line is broad enough for them to be blended. The velocity resolution ranges
from 10 to 50 km s−1. It was selected to be 10% of the FWHM of the line itself, but for a few cases the S/N of the
spectrum required further smoothing.
cast light on this issue. Of course, the HNC abundance
must be high enough so that the pumping can be effective
and it must thus happen in regions where the chemistry
is dominated by ion-neutral reactions.
4.2. Interpreting the CN emission
4.2.1. CN chemistry
Studies of Galactic molecular clouds have shown that the
[CN]
[HCN] abundance ratio is increasing in the outer regions
of UV irradiated clouds (e.g., Greaves & Church 1996;
S. Aalto, A. G. Polatidis, S. Huttemeister, S. J. Curran: CN,HNC in Galaxies
9
Arp 220
J=3/2-1/2
NGC 34
J=3/2-1/2
NGC 1614
J=3/2-1/2
J=5/2-3/2
J=5/2-3/2
J=5/2-3/2
NGC 1808
J=3/2-1/2
NGC 3256
J=3/2-1/2
NGC 6240
J=3/2-1/2
J=5/2-3/2
J=5/2-3/2
J=5/2-3/2
NGC 7130
J=3/2-1/2
J=5/2-3/2
Fig. 4. CN 2-1 spectra for the galaxies with detections. The scale is in T ∗
A. The band is centered in between the two
spingroups, apart from Arp 220, where the spectrometer was centered on the J=5/2-3/2 line. The two spingroups are
marked with arrows and are blended for all galaxies, apart from NGC 7130. FOR Arp 220 emission is not detected at
the line center (marked by the arrow) instead, there is a tentative detection at V =5170 km s−1 (blueshifted from Vc
by approximately 250 km s−1). The velocity resolution ranges from 10 to 50 km s−1. It was selected to be 10% of the
FWHM of the line itself, but for a few cases the S/N of the spectrum required further smoothing.
Rodriguez-Franco et al. 1998). The abundance of the CN
radical becomes enhanced at the inner edge of a PDR (at
an AV of about 2 magnitudes) via the reaction N + C2 →
CN+C or via N+CH → CN+H. At larger depths into the
cloud the CN abundance radically declines and the [HCN]
[CN]
abundance ratio increases (Jansen et al. 1995). Most of the
CN is present in a part of the cloud where the abundance
of free electrons is rather large, X(e) ≈ 3×10−5. CN is also
a photodissociation product of HCN. Thus the CN abun-
dance should be favoured in a molecular cloud ensemble
dominated by PDRs. Furthermore, chemical models (e.g.,
Krolik & Kallman 1983; Lepp & Dalgarno 1996) show that
the CN abundance should also be enhanced when the X-
ray ionization rates are high -- as might be the case near
an AGN.
It is of course difficult to translate a measured HCN
CN 1-0
line intensity ratio to an abundance ratio between the two
species. The spingroup line ratios (see Sect. 3.2.1.) show
that the CN 2-1 emission is optically thin for most galax-
ies we have measured. The emission of the CN molecule is
distributed in a greater number of transitions than HCN,
thus the optical depth per transition is often lower for CN,
reducing the intensity per line. The CN luminosity then
becomes a measure of the total number of CN molecules
(at least in a comparative sense, given a constant excita-
tion situation from galaxy to galaxy). For the CN 1-0 line
we have only information for Arp 220 where the relative
faintness of the second spingroup suggests that the optical
depth of the 1-0 line is also low. So, the measured (total)
HCN
CN line intensity ratio will give a reasonable idea of the
abundance ratio if also the HCN line is close to being op-
10
S. Aalto, A. G. Polatidis, S. Huttemeister, S. J. Curran: CN,HNC in Galaxies
tically thin (and a lower limit to the abundance ratio if
it is not) and if the same excitation temperature can be
assumed. The critical density of the CN line is lower by a
factor of a few, so its Tex is likely somewhat higher.
4.2.2. AGNs or starbursts?
∼ I(HCN)
There are only a few galaxies where I(CN) >
(Arp 220, NGC 3690 and NGC 1808) -- all three are
starburst galaxies. For these galaxies N (CN) is greater
than N (HCN) and the dense gas is to a large degree af-
fected by photodissociation. The most extreme example
is NGC 3690 where the CN line is on average more than
a factor of two brighter than HCN. In Arp 220 the CN
appears to be bright only towards the western of the two
nuclei (Sect. 3.2). This is interesting, since the situation is
similar for the IC 694/NGC 3690 system: bright CN emis-
sion towards one of the galaxy nuclei only. It is interesting
to speculate on whether the burst towards NGC 3690 is
older than the one in IC 694 because of the development
of the PDRs. However, the IC 694 burst is more compact
and it is possible that the actual properties of the ISM are
intrinsically different.
We were surprised to find that CN was difficult to
detect in several of the brightest galaxies like Mrk 273,
NGC 2623, NGC 6240, IC 694 and NGC 34. Four of these
galaxies are AGNs and it is tempting to speculate that
the CN deficiency is related to the nuclear activity. This
seems contrary to models (see above) which predict an in-
crease in the CN abundance in an X-ray chemistry. High
resolution studies of nearby systems which contain both
starburst and AGN activity (like NGC 1068) will reveal
whether CN emission is associated with one or both of the
activities.
4.2.3. The CN and the [C II] 158 µm line
In ULIRGs such as Arp 220 and Mrk 231 the [C II] 158
µm fine structure line is found to be abnormally faint com-
pared to other, less FIR luminous, starburst galaxies like
NGC 3690 (e.g., Luhman et al. 1998). This is interesting,
since one would expect the emission from a standard PDR
tracer, like the [C II] line, to be bright in a galaxy that
is believed to be powered largely by mighty starbursts.
Malhotra reports a decreasing trend in F[CII]
with increas-
FFIR
ing f (60)
f (100) µm flux ratio.
Several possible explanations for the [C II] faintness are
brought forward (e.g., Malhotra et al. 1997; Luhman et al.
1998; van der Werf 2001). The PDRs may be quenched in
the high pressure, high density environment in the deep
potentials of the ULIRGs and the HII regions exist in
forms of small-volume, ultracompact HII regions that are
dust-bounded. The [C II] line may become saturated ei-
ther in low density (n ∝ 102 cm−3) regions of very high
UV fields (G0 ∝ 103) or in dense (n ∝ 105 cm−3) re-
gions of more moderate UV fields (G0 = 5 − 10). A soft
UV field from an aging starburst is another possibility. A
higher dust-to-gas ratio would also decrease the expected
F[CII]
FFIR
ratio.
The molecular ISM of ULIRGs seems to be character-
ized by subthermally excited CO and very bright emis-
sion from HCN (e.g., Downes & Solomon 1998; Aalto et
al. 1995; Solomon et al. 1992). Crudely, this can be mod-
elled as dense clouds (n = 104 − 105 cm−3) embedded in
a low density (n = 100 − 500 cm−3) continous medium.
This simple scenario may fit well with the two scenarios
resulting in saturated [C II] emission.
We have three ULIRGs in our sample: Arp 220,
Mrk 231 and Mrk 273 their HCN
CN line intensity ratio
changes from 1 to >
∼ 6. The deficiency of CN in Mrk 273
is consistent with the lack of [C II] emission and can be
an indication that the PDRs are not forming in the dense
gas. In Arp 220 the CN emission from the western nucleus
is strong and an indicator that a fair fraction of the dense
gas is in fact in a PDR state. The UV radiation is strongly
affecting the dense molecular clouds here. Clearly we need
more information on the properties of the dense gas there
to find an explanation for the lack of [C II] emission.
4.3. Line ratios and starburst evolution
We speculate whether line ratios of dense gas tracers can
be used to explore the evolutionary stage of a starburst.
The observed galaxies can be sorted into rough categories
based on their line ratios.
4.3.1. Warm dense gas -- a young starburst or shocks?
Objects that show bright HCN emission, but little or no
HNC or CN, may be dominated by warm dense gas early
in their starburst development. The Orion KL region is
an example of a Galactic warm (T >
∼ 50 K), dense core
where I(HCN) is significantly greater than both I(CN)
and I(HNC) (e.g., Ungerechts et al. 1997). Also the emis-
sion from the 10-9 transition of HC3N is brighter than
the CN and HNC emission which is a typical signature
of warm, dense gas. However, shocked gas may be an im-
portant part of the molecular ISM in the center of a star-
burst galaxy -- in particular if there is a bar in the center
where clouds on intersecting orbits collide. The interac-
tion between supernova remnants at the surrounding ISM
may also lead to the presence of shocked gas. The effect
of the shock is to compress and heat the gas, which in
some respects will make it look like an ISM dominated by
warm dense cores. The major difference is that the shock,
partially or fully, destroys the dust grains and thereby re-
leasing molecules, such as SiO, into the ISM. Therefore,
SiO emission is often used as a tracer of shocked gas (e.g.,
Martin-Pintado et al. 1992). If the shocked gas is allowed
to cool after the shock (which may happen quickly since
it has been compressed) the HNC abundance will increase
rapidly (S92). An example of a galaxy that could have
a shock dominated ISM is NGC 6240. Very strong IR
emission is emerging from shock-excited H2 (van der Werf
S. Aalto, A. G. Polatidis, S. Huttemeister, S. J. Curran: CN,HNC in Galaxies
11
2001) in between the two merger nuclei and both HNC
and CN line emission is faint relative to that of HCN. In
a high pressure environment the gas may be dense and
warm -- but perhaps heated by dissipation of turbulence
rather than very young embedded stars. This may not oc-
cur during the very early stages of star formation, but
rather be a form of aftermath.
>
(e.g., Churchwell et al. 1984). We know, however, from
other studies of IC 694 that the dense gas is warm
(Tkin
∼ 80K) (Aalto et al. 1998) and the compact CO nu-
cleus of NGC 2623 likely harbours an ISM similar to that
of IC 694. The starburst may have evolved beyond a strong
radiative impact from the stars while the chemisty is still
dominated by ion-neutral reactions at moderate gas den-
sity and relatively high electron abundance.
4.3.2. PDRs
For those galaxies where I(CN) is >
∼ I(HCN) a significant
part of the ISM should be in PDRs. In the Galaxy they
are often found in interface regions between HII regions
and molecular clouds (e.g., Jansen et al. 1995) and near
planetary nebulae. CN luminous galaxies are very clearly
in the phase where the ISM is being strongly affected by
an intense UV field. The filling factor of UV illuminated
gas must be high, and the clouds probably not too large
since the [CN]
[HCN] abundance ratio drops dramatically with
increasing AV. The mechanical impact of the starburst
(superwinds - supernovae) may help in fragmenting the
molecular clouds. Emission from complex molecules, such
as HC3N, is faint because of photodestruction. Since the
chemistry now, to a large degree, involves ion-neutral re-
actions (see Sect. 4.1) the HNC abundance becomes less
dependent on temperature and I(HNC) may be signifi-
cant even from a warm PDR. For example, the temper-
ature around a planetary nebula may become very high
(≈ 100 K) but the HNC abundance is substantial enough
to result in HCN
HNC line intensity ratios close to unity (e.g.,
Herpin & Cernicharo 2000).
For Arp 220, it appears that HC3N is mainly emerging
from the eastern nucleus (see Sect 3.2.1) which would sup-
port the notion of an evolutionary difference between the
two nuclei. Rodriguez-Franco et al. (1998) show that the
emission from HC3N is bright toward hot, dense cores,
while the HC3N/CN abundance ratio is only 10−3 in
PDRs. Thus the eastern nucleus seems to be in an earlier
evolutionary phase where star formation has just begun.
In the mid-IR the western nucleus is more prominent than
the eastern one (Soifer et al. 1999).
As discussed in Sect. 4.2.1. the absence of PDR tracers,
such as CN emission and the 158 µm [C II] fine structure
line, does not necessarily mean that the burst is young
(or shock dominated). In high-pressure, dusty ULIRGs it
is possible that the formation of PDRs is suppressed --
or that the PDRs are associated with the diffuse lower
density molecular material where the classic PDR lines
will not be excited.
4.3.3. An evolved burst
Two galaxies (NGC 2623 and IC 694) show fairly bright
HNC emission but with undetected CN emission. Very
cold (10 K) and dense gas would result in I(HCN)=
0.3 − 0.5 × I(HNC) and I(CN) < I(HCN). In the Galaxy
such conditions dominate clouds like TMC-1 and TMC-2
5. Summary
We have undertaken a SEST and OSO survey of CN and
HNC line emission in a sample of 13 luminous IR galax-
ies, plus one more "normal" starburst (NGC 1808). This
survey is the first in its kind for IR luminous galaxies. The
main conclusions we draw from this survey are as follows:
1. We have detected HNC 1-0 in 9 of the galaxies,
while CN 1-0 is detected in 10 galaxies. The CN 2-1
lines (J=5/2-3/2, F =7/2-5/2; J=3/2-1/2 F=5/2-3/2)
(which could only be measured at SEST) are detected
in 7 galaxies -- in one (NGC 34) the CN 1-0 line was
not detected, while the 2-1 line was.
2. The line intensitites vary significantly from galaxy to
galaxy and the line ratios relative to the "standard"
high density tracer, HCN, are far from constant. The
HCN
HNC ratios vary from 1 to >
CN ratios range
from 0.5 to >
∼ 6. From this we can learn that the ac-
tual properties of the dense gas vary significantly from
galaxy to galaxy, even if their HCN luminosities are
similar.
∼ 6 and the HCN
3. We report the detection of HC3N 10-9 emission in
Arp 220 at 40% of the HNC 1-0 intensity. In NGC 1808
the line is tentatively detected.
4. We find that the IR luminous galaxies are at least as
HNC luminous as more nearby (IR-fainter) galaxies.
We conclude that the HNC emission is not a reliable
tracer of cold (10 K) gas in the center of luminous
IR galaxies, the way it may be in clouds in the disk
of the Milky Way. Instead, we list several alternative
(testable) explanations. Provided that the situation in
the centers of starburst galaxies can be approximated
with a steady state chemistry, the observed HCN
HNC line
ratios can be explained by the emission originating in
∼ 105 cm−3 where the chemistry is
gas of densities n <
dominated by ion-neutral, instead of neutral-neutral,
reactions. The temperature dependence of the HNC
production in the ion-neutral reactions is strongly sup-
pressed.
5. We were surprised to find that the CN emission of the
more luminous galaxies is often faint. This can be con-
sistent with the relative faintness of other PDR tracers
such as the 158 µm [C II] line. Both these phenom-
ena can be understood in terms of a two-component
molecular ISM consisting of low density molecular gas
surrounding dense clouds. However, the relative bright-
ness of CN in Arp 220 may be difficult to reconcile with
its relative [C II] faintness.
12
S. Aalto, A. G. Polatidis, S. Huttemeister, S. J. Curran: CN,HNC in Galaxies
Lepp, S., &, Dalgarno, A., 1996, A&A, 306, L21
Luhman M. L., Satyapal S., Fischer, J., et al., 1998, ApJ, 504,
L11
Malhotra S., Helou G., Stacey G., et al., 1997, ApJL 491, 27
Martin-Pintado, J., Bachiller, R., & Fuente, A., 1992, A&A
254, 315
Polatidis, A., & Aalto, S., 2000, in Proceedings of the 5th EVN
Symposium, Eds. J. Conway, A. Polatidis, R.Booth. Y.
Pihlstrm, Onsala Space Observatory, Chalmers Technical
University, Gothenburg, Sweden (2000), 127
Rodriguez-Franco, A., Martin-Pintado, J., & Fuente, A., 1998,
A&A, 329, 1097
Schilke P., Walmsley C.M., Pineau de Forets G., et al., 1992,
A&A 256, 595 (S92)
Scoville, N.Z., Yun, M.S, & Bryant, P. M., 1997, ApJ, 484, 702
Soifer B. T., Neugebauer G., Matthews K., et al., 1999, ApJ,
513, 207
Solomon P.M., Downes D., & Radford S.J.E., 1992, ApJ 387,
L55
Ungerechts H., Bergin E.A., Goldsmith P.F., et al., 1997, ApJ
482, 245
van der Werf, P.P., 2001, in Starburst galaxies - near and far,
proceedings of the Ringberg workshop, eds. D. Lutz & L.J.
Tacconi, in press
Wall, W.F., Jaffe, D.T, Bash, F.N., et al., 1993, ApJ, 414, 98
Wild, W., Harris, A.I., Eckart, A., et al., 1992, A&A, 265, 447
6. Within Arp 299 (the merging pair IC 694 and
NGC3690) there is a line ratio difference between the
two nuclei. NGC 3690 has a CN luminosity twice that
of HCN and its ISM is thus strongly affected by UV
radiation while CN is not detected towards IC 694.
This reflects an evolutionary gradient in the burst --
or that the star formation activity takes place in very
different environments in the two galaxies. For Arp 220
there may also be an evolutionary gradient between the
two nuclei because of differences in emission velocities
for CN and HC3N.
7. The measured line ratios can in general be used to
identify stages in the starburst evolution, but we con-
clude that the dichotomy between scenarios is a com-
plication. For example, faint HNC emission is expected
both in a shock dominated ISM as well as for a cloud
ensemble dominated by dense warm gas in the very
early stages of a starburst.
Acknowledgements. Many thanks to the OSO and SEST staff
for their help. We are grateful to M. Walmsley and P. Schilke
for discussions on the HNC chemistry. We thank the referee,
F. Herpin, for many useful comments and suggestions.
References
Aalto S., Black J.H., Booth R.S., Johansson L.E.B., 1991,
A&A, 247, 291
Aalto S., Booth R.S., Black J.H., Johansson, L.E.B., 1994,
A&A, 286, 365
Aalto, S., Booth, R.S., Black, J.H., Johansson, L.E.B. 1995,
A&A, 300, 369
Aalto S., Radford, S.J.E., Scoville, N.Z., Sargent, A.I., 1997,
ApJ, 475, L107
Aalto S., Radford, S.J.E., Scoville, NZ., Sargent, A.I., 1998,
IAU Symposium 186 'Galaxy Interactions at Low and High
Redshift', Ed. D. Sanders
Bachiller, R., Fuente, A., Bujarrabal, V., Colomer, F., Loup,
C., Omont, A., de Jong, T., 1997, A&A, 319, 235
Barvainis, R., Maloney, P., Antonucci, R., Alloin, D., 1997,
ApJ, 484, 695
Braine, J., & Dumke, M., 1998, A&A, 333, 38
Bryant, P.M., 1996, PhD thesis, Caltech
Bryant, P.M., & Scoville, N.Z., 1999, AJ, 117, 2632
Churchwell, E., Nash A. G., & Walmsley C. M., 1984, ApJ 287,
681
Curran, S.J., Aalto, S., & Booth, R.S. 2000, A&AS, 141, 193
(CAB)
Downes, D., & Solomon, P.M., 1998, ApJ, 507, 615
Fuente, A., Martin-Pintado, J., & Gaume, R, 1995, ApJ, 442,
L33
Greaves J. S., & Church S. E., 1996, MNRAS 283, 1179
Helfer, T.T., & Blitz L., 1993, ApJ, 419, 86
Herpin, F., & Cernicharo, J., 2000, ApJ, 530, L129
Huttemeister S, Henkel C., Mauersberger R., et al., 1995, A&A
295, 571
Jansen D. J., van Dishoeck,E. F., Black J. H., Spaans M., Sosin
C., 1995, A&A, 302, 223
Klaas, U., Haas, M., Heinrichsen, I., Schulz, B., 1997 A&A,
325, L21
Krolik, J.H., &, Kallman, T.R., 1983, ApJ, 267, 610
|
astro-ph/0101271 | 1 | 0101 | 2001-01-16T16:54:48 | The X-ray Emission in Post-Merger Ellipticals | [
"astro-ph"
] | The evolution in X-ray properties of early-type galaxies is largely unconstrained. In particular, little is known about how, and if, remnants of mergers generate hot gas halos. Here we examine the relationship between X-ray luminosity and galaxy age for a sample of early-type galaxies. Comparing normalized X-ray luminosity to three different age indicators we find that L_X/L_B increases with age, suggesting an increase in X-ray halo mass with time after a galaxy's last major star-formation episode. The long-term nature of this trend, which appears to continue across the full age range of our sample, poses a challenge for many models of hot halo formation. We conclude that models involving a declining rate of type Ia supernovae, and a transition from outflow to inflow of the gas originally lost by galactic stars, offers the most promising explanation for the observed evolution in X-ray luminosity. | astro-ph | astro-ph |
Mon. Not. R. Astron. Soc. 000, 1 -- 8 (0000)
Printed 27 October 2018
(MN LaTEX style file v1.4)
The X -- ray Emission in Post -- Merger Ellipticals
Ewan O'Sullivan1, Duncan A. Forbes1,2 and Trevor J. Ponman1
1School of Physics and Astronomy, University of Birmingham, Edgbaston, Birmingham B15 2TT
(E-mail: [email protected])
2Astrophysics & Supercomputing, Swinburne University, Hawthorn VIC 3122, Australia
27 October 2018
ABSTRACT
The evolution in X -- ray properties of early -- type galaxies is largely unconstrained. In
particular, little is known about how, and if, remnants of mergers generate hot gas
halos. Here we examine the relationship between X -- ray luminosity and galaxy age
for a sample of early -- type galaxies. Comparing normalized X -- ray luminosity to three
different age indicators we find that LX /LB increases with age, suggesting an increase
in X -- ray halo mass with time after a galaxy's last major star-formation episode. The
long-term nature of this trend, which appears to continue across the full age range of
our sample, poses a challenge for many models of hot halo formation. We conclude
that models involving a declining rate of type Ia supernovae, and a transition from
outflow to inflow of the gas originally lost by galactic stars, offers the most promising
explanation for the observed evolution in X-ray luminosity.
Key words:
evolution -- X-rays: galaxies
galaxies: interactions -- galaxies: elliptical and lenticular -- galaxies:
1
INTRODUCTION
A potential problem with forming ellipticals from merging
spirals is how to account for the different gas properties in
the two types of galaxies. In particular, spirals contain rel-
atively high masses of cold (T ∼ 100 K) gas, whereas el-
lipticals have very little. The opposite situation is true for
the hot (T ∼ 106 K) gas masses -- normal spirals contain
rather little hot gas, while ellipticals may possess exten-
sive hot halos. Since the total gas masses per unit stellar
mass are broadly comparable in early and late-type galax-
ies, this raises the possibility that inefficient merger -- induced
star formation might heat the cool interstellar gas in spirals
to form the hot halo in post -- merger ellipticals. Recently,
Georgakakis et al. (2000) have shown that the cold gas mass
indeed decreases in an 'evolutionary sequence' from merging
spirals to post -- merger ellipticals.
However, it is not clear that gas heated in an intense
starburst can be retained within the galactic potential. Read
& Ponman (1998) examined the X -- ray properties of eight
on -- going mergers placed in a chronological sequence. Their
study revealed that material is ejected from merging galax-
ies soon after the first encounter. Massive extensions of hot
gas are seen (involving up to 1010 M⊙) at the ultralumi-
nous peak of the interaction, as the two nuclei coalesce.
There is evidence for this in in the Antennae, Arp 220, and
NGC 2623. However, after this phase of peak activity, the
c(cid:13) 0000 RAS
X -- ray luminosity actually declines. For example, NGC 7252
(a prime example for the remnant of a merger that occurred
0.7 Gyr ago) shows some evidence of a hot halo, but one that
is much smaller than the halos seen in typical ellipticals.
Two post -- merger ellipticals were studied in the X -- ray
by Fabbiano & Schweizer (1995). They found that neither
galaxy had an extensive hot halo. Given the correlation be-
tween X -- ray luminosity and isophotal boxiness (thought to
be a signature of a past merger) demonstrated by Bender
et al. (1989), this was a somewhat surprising result. If post --
merger ellipticals are to resemble 'normal' ellipticals they
need to acquire a hot gas halo. Possible mechanisms include:
1) The late infall (i.e. after a few Gyr) of HI gas associated
with the tidal tails (Hibbard et al. 1994). This cold gas may
be shocked to X -- ray emitting temperatures as it falls back
into the merger remnant.
2) A reservoir of hot gas, possibly expelled at the nuclear
merger stage, might infall from large radii, where its low
density makes it undetectable to current X -- ray satellites.
3) After the initial violent starburst, the continued mass
loss from stars and heating from stellar winds recreates a
hot ISM.
The first study to attempt an 'evolutionary sequence'
for post -- merger ellipticals was that of Mackie & Fabbiano
(1997). For 32 galaxies they showed a weak trend for LX /LB
to increase with a decreasing Σ parameter. The Σ parame-
2
Ewan O'Sullivan et al.
ter is defined by Schweizer & Seitzer (1992), and is a mea-
sure of a galaxy's fine structure, i.e. optical disturbance.
They showed that it correlates with blue colours and Balmer
line strength, and is thus a rough indicator of dynamical
youth. The Mackie & Fabbiano trend therefore suggested
that post -- merger ellipticals became more X -- ray luminous,
for a given optical luminosity, as the galaxy aged. They also
plotted LX/LB against Hβ absorption line EW, which re-
vealed a similar trend. A more recent study by Sansom et al.
(2000) confirms the trend in X -- ray overluminosity with Σ
for 38 galaxies, and suggests that it might be explained by
the build up of hot gas in post -- merger galaxies. However,
both Σ and Hβ EW have their drawbacks -- Σ is only semi --
quantitative at best and the strength of Hβ is affected by
both stellar age and metallicity.
Here we reexamine the trend seen by Mackie & Fab-
biano (1997) for a larger sample, and investigate two new
measures of galaxy age: residual from the Fundamental
Plane and spectroscopic age. Using these three measures we
explore the X -- ray luminosity evolution of post -- merger el-
lipticals.
Throughout the paper we assume H0 = 75 km s−1
Mpc−1 and normalise LB using the solar luminosity in the
B band, LB⊙ = 5.2×1032 erg s−1.
2 RESULTS
We first reexamine the trend of normalized X -- ray lumi-
nosity with Fine Structure parameter Σ, presented initially
by Mackie & Fabbiano (1997) for 32 galaxies. Here we
use a sample of 47 early -- type galaxies with Σ taken from
Schweizer & Seitzer (1992), and normalized LX values from
our recent catalogue (O'Sullivan et al. 2000) of X -- ray lu-
minosities (mostly based on ROSAT data). The LX val-
ues are approximately bolometric. The LB values are based
wherever possible on BT magnitudes taken from Prugniel
& Simien (1996). If these are unavailable, values from NED
are used. Fig. 1 shows the sample of 47 early -- type galaxies
plotted in order of decreasing Σ or increasing age.
As noted by Mackie & Fabbiano, the scatter at low val-
ues of Σ is large, covering around 2 orders of magnitude in
LX /LB, while at higher Σ the scatter appears to be smaller
and LX /LB limited to low values. A similar trend was seen
by Sansom et al. (2000) for 38 galaxies. This suggests that
dynamically young galaxies have low X -- ray luminosities and
that aging produces a range of luminosities, presumably de-
pendent on other factors.
In order to test the strength of this relation we apply
Kendall's K test to the X -- ray detections from our sample.
The test does not assume any distribution in the data, and
the K statistic is unit normal distributed when at least 10
data points are present. For our data we find K = -2.31054
(using 27 data points), indicating an anti correlation be-
tween Log LX /LB and Σ of ∼2-σ significance. For com-
parison, we also apply the test to the sample of Sansom et
al. , and find a correlation of very similar significance, K =
-2.28749 for 30 detections.
2.1 LX /LB versus Spectroscopic Age
Fine structure only persists in dynamically young galaxies,
so Fig. 1 provides information only on the early galaxy evo-
lution of these objects. A more general measure of age is
needed to show how X -- ray properties evolve over a longer
timescale. Galaxy spectroscopic ages are now available for
a large number of early -- type galaxies from the catalogue of
Terlevich & Forbes (2000). These ages are generally based
on Hβ absorption line measurements and the stellar popu-
lation models of Worthey et al. (1994). The line index mea-
surements come from the galaxies' central regions, and are
luminosity weighted. Thus they are dominated by the last
major burst of star formation, which is presumably triggered
by a major merger event. Although not a reliable absolute
measure of the age of each galaxy, these spectroscopic ages
do provide us with a much more useful estimate of their ages
relative to one another. In Fig. 2 we show LX/LB plotted
against spectroscopic age (from Terlevich & Forbes 2000).
The plot shows a large degree of scatter, most notably
in the age range between 4 and 10 Gyrs. This is likely to
be in part caused by the uncertainties in calculating ages,
which we estimate lead a typical error of ∼20%. We also
expect a mean 1-σ error in X -- ray luminosity of ∼15%. Typ-
ical error bounds for points at 1, 7 and 15 Gys are shown
in Fig. 2 by the diamonds at the top of the plot. The graph
also contains a relatively large number of upper limits, so
we have used the survival analysis packages available under
IRAF to assess the strength of any correlation and fit regres-
sion lines. The Cox proportional hazard, Spearman's rho and
generalized Kendall's tau tests are used to determine corre-
lation strength. Linear regression fitting is carried out using
the expectation and maximization (EM) algorithm and the
Buckley -- James (BJ) algorithm. In all cases we found that
the two fitting algorithms agreed closely.
Using the full sample of 77 early-type galaxies, we find
a correlation of at least 99.6% significance between LX /LB
and age. Line fits to the sample produce slopes of 0.066 ±
0.022 (EM) and 0.063 ± 0.025 (BJ). The former is shown as
the dashed line in Fig. 2.
Recent work on the X -- ray properties of galaxies in
groups (Helsdon et al. 2000) lead us to suspect that some
of the scatter seen in Fig. 2 might be caused by inclusion of
group or cluster dominant galaxies (marked as crossed cir-
cles) in our sample. It is also important to note that there is
evidence to suggest that the largest ellipticals may have non --
solar abundance ratios (e.g. Carollo et al. 1993). This could
lead to their ages being underestimated by a small amount,
again adding to the scatter. We therefore removed from our
sample a number of galaxies listed by Garcia (1993) as group
dominant or known cDs and retested the sample. This im-
proves the correlation slightly (to >99.75%) and makes the
line fit slightly shallower; 0.063 ± 0.019 (EM) and 0.061 ±
0.022 (BJ). The EM fit is plotted as a solid line in Fig. 2.
To confirm that the increase in LX /LB occurs across the
range of ages, rather than just the first few Gyrs, we also
binned the sample into subsets by age. The age range of each
subset was chosen so as to have roughly equal numbers of
detections in each bin. We then calculated a mean LX /LB
for each bin, using the Kaplan -- Meier estimator. The results
are shown in Table 1, and as large crosses on Fig. 2. These
show the same trend for increasing LX /LB with age, and
c(cid:13) 0000 RAS, MNRAS 000, 1 -- 8
X -- rays from Post -- Mergers
3
32
31
30
29
10
8
6
4
2
0
Figure 1. Normalized X -- ray luminosity versus fine structure parameter Σ for early -- type galaxies. Filled circles are detections, arrows
denote upper limits. Dynamically young galaxies (high Σ) have low LX /LB .
Age
(Gyrs)
A < 4
4 ≤ A < 6
6 ≤ A < 8
8 ≤ A < 10
A ≥ 10
No. of detections No. of upper limits
Log LX /LB
(erg s−1 L−1
B⊙)
9
6
9
9
8
12
11
4
6
4
29.56
29.83
29.93
30.15
30.32
error
±0.11
±0.15
±0.22
±0.14
±0.09
Table 1. Mean LX /LB values for 5 age bins, calculated using the Kaplan -- Meier estimator.
indicate that the trend is continuous across the age range
covered by the data.
The Kaplan -- Meier estimator (Feigelson & Nelson 1985)
includes upper limits by constructing a probability distribu-
tion function for the data in which the probability asso-
ciated with each upper limit is redistributed equally over
detected values which lie below the upper limit. So long as
there is no systematic difference between the detected and
undetected systems, this should be a reasonable procedure.
A problem arises when the lowest point in a given bin is an
upper limit (since there are no lower points over which to re-
distribute the corresponding probability). When this occurs,
the Kaplan -- Meier estimator treats the limit as a detection
and hence may be biased.
Of the five bins chosen, two (the youngest and second
oldest) have upper limits as their lowest values. For the sec-
ond oldest bin, the problem data value is only slightly lower
than a detected point, so the bias will be small. However,
the youngest bin has two upper limits as its lowest points,
one of which is considerably below the nearest detection. To
check for bias, we recalculated a mean log LX /LB of 29.66
for this bin excluding these points, compared to 29.56 pre-
c(cid:13) 0000 RAS, MNRAS 000, 1 -- 8
4
Ewan O'Sullivan et al.
Figure 2. Log (LX /LB ) plotted against spectroscopic age. Crossed circles represent group or cluster dominant galaxies, filled circles all
other detections. Arrows denote upper limits and the filled triangle represents NGC 7252, a young post -- merger. The three diamonds at
the top of the graph show typical age and LX /LB errors for galaxies of age 1, 7 and 15 Gys. The two lines are fits to the data, including
(dashed line) or excluding (solid line) the group and cluster dominant galaxies. The large crosses represent mean LX /LB for the five age
bins described in the text. The dotted line shows the expected increase in LX /LB caused by the decrease of LB with age. Normalisation
of this line is arbitrary, and the position shown can be considered to represent a "worst case" scenario.
viously. This indicates that although the youngest bin may
be slightly biased, it still supports the trend observed.
As a further check of the robustness of our conclusions,
we examined the LX /LB:Age relation using detected points
only, and find a correlation significant at ∼2.1-σ. The upper
limits also show a correlation with age (at ∼2.3-σ). These
trends cannot be selection effects in the data, since neither
source distance, nor X-ray exposure time are correlated with
age. They therefore point to a genuine trend in the data.
Galaxies with larger spectroscopic ages typically have higher
LX /LB than younger ones.
As the merger -- induced starburst fades, we would expect
the blue luminosity to decline with time. This will cause an
increase in LX /LB, even for constant LX , which will com-
bine with any trends in LX which are present. With this
is mind, we show in Fig. 2 the expected change due to a
fading starburst from stellar population models (Worthey
1994). Here we have crudely assumed that the progenitor
galaxies consisted of a solar metallicity, 15 Gyr old popula-
tion, and that the new stars created in the merger also have
solar metallicity. The relative mass ratios are 90% progeni-
tor stars and 10% new stars. The total B band luminosity
fading of this composite stellar population is shown in Fig.
2 (we assumed that the fading of the progenitor stars after
reaching 18 Gyrs old is insignificant). It can be seen that the
starburst fades rapidly in the first few Gyrs, roughly match-
ing the trend seen in LX /LB, but at later times there is
very little change, whereas the observed mean LX /LB value
continues to rise. So although the LX /LB evolution at early
times could be simply driven by the fading starburst, this
cannot account for the evolution at late times.
A further possibility to be checked is that our sample
might contain a trend in mean optical luminosity with age.
Since LX/LB has been widely reported (see however Hels-
don et al. (2000)) to rise with LB, this could lead to a con-
sequent correlation between age and LX /LB. To test this,
we show in Fig. 3 a plot of log LB with measured age. An
anti -- correlation between LB with age could produce the ob-
c(cid:13) 0000 RAS, MNRAS 000, 1 -- 8
served trend in LX /LB. Fig. 3 does not appear to show any
such trend, and in order to test for a correlation, we apply
Kendall's K test to the data. We find a K statistic (which is
unit normal distributed) of ∼0.44 (for 77 galaxies), and ex-
cluding group dominant galaxies reduces this to ∼0.33 (for
65 galaxies). This result indicates that there is no trend in
the data, even at the 1σ level. We therefore conclude that
our sample does not show an anti -- correlation of LB with
age, and that the LX /LB trend with age reflects an increas-
ing X -- ray luminosity.
2.2 LX /LB versus Fundamental Plane Residual
As further evidence of X -- ray evolution with age we have
compared LX/LB to residual from the Fundamental Plane
(FP). Work by Forbes et al. (1998) has shown that galaxy
age appears to be a "fourth parameter" affecting the position
of a galaxy relative to the FP plane. Galaxies below the
plane (negative residual) are generally young, while older
objects lie on or above the plane (positive residual). Fig. 4
shows 212 galaxies with Fundamental Plane residuals taken
from Prugniel & Simien (1996).
As with Figs. 1 and 2, this shows a trend of increasing
LX /LB with age. For the complete data set, the correla-
tion strength is >99.98%. However, only a small number of
points lie outside the range -- 0.5<FP residual<0.5. These are
unlikely to be representative of the general population, but
will have a strong influence on the statistical tests. Exclud-
ing them lowers the correlation strength to ∼99.95%, with
a slope of 1.39 ± 0.40 (EM). This line is shown in Fig. 4.
Again the trend shown in Fig. 4 is one of increasing X -- ray
luminosity, relative to the optical, as a galaxy evolves.
3 DISCUSSION
In the previous section we have used three age estimators to
show a strong correlation between normalized X -- ray lumi-
nosity and galaxy age. The X -- ray emission from early -- type
galaxies is known to be produced by sources of two main
types; discrete sources such as X -- ray binaries, and hot gas.
The contribution of discrete sources has been shown to be
important in low luminosity early-type galaxies (Fabbiano
et al. 1994; Irwin & Sarazin 1998; Sarazin et al. 2000) and
in late-type galaxies, where they are generally the dominant
source of X -- ray emission. On the other hand, most early --
type galaxies have a strong hot gas component, and massive
ellipticals are certainly dominated by hot gas emission (Mat-
sushita et al. 2000; Matsushita 2000). As the contribution
of hot gas is known to vary a great deal, while the contri-
bution from discrete sources is generally believed to scale
with LB (Matsushita et al. 2000; Fabbiano et al. 1992) , it
seems likely that the trend we observe in LX /LB is caused
by changes in the hot gas content of these objects with age.
We now go on to discuss possible hot halo formation mecha-
nisms which may be responsible for this trend of increasing
X -- ray gas content with age.
3.1 Gas infall
It has been suggested (Hibbard & van Gorkom 1996) that
the hot gas in elliptical galaxies may be produced during
c(cid:13) 0000 RAS, MNRAS 000, 1 -- 8
X -- rays from Post -- Mergers
5
a merger by the shock heating or photoionization of cool
gas from the progenitor galaxies. In some ongoing mergers
the tidal tails are thought to contain up to half the Hi gas
originally present in the progenitor galaxies (Hibbard et al.
1994). When this gas falls back into the body of the galaxy,
shock heating should be capable of heating it to X -- ray tem-
peratures. An alternative is that the temperature of the gas
is caused by heating in the starburst phase of the merger. In
this case the hot gas would be blown out to large radii, but
might be contained by the dark matter halo of the galaxy
(Mathews & Brighenti 1998). At these large radii it would
be too diffuse to be detected, but would eventually fall back
into the galaxy, forming the observed X -- ray halo.
In terms of available masses of gas, both these mod-
els appear to be viable formation processes. Bregman et al.
(1992) derived X -- ray and cold gas masses for a large sample
of early -- type galaxies, finding MX of between ∼ 106 and
1011 M⊙. The mean X -- ray gas mass was a few 109 M⊙.
On the other hand, very few detections of Hi were made
in early -- type galaxies, although Sa galaxies were found to
contain between 107 and ∼ 5×1010 M⊙ of Hi . Studies of Hi
masses in later -- type galaxies (Huchtmeier & Richter 1989;
Huchtmeier & Richter 1988) give an average of a few 109
M⊙, depending on the range of morphologies chosen. Given
that large elliptical galaxies are likely to have been formed by
the merger of many smaller spiral galaxies, these results sug-
gest that fairly modest conversion efficiencies could produce
the expected X -- ray halo gas masses either from infalling Hi
or via starburst heating and later infall.
However, the infalling Hi model fails to explain the
timescale over which we observe the LX /LB trend. In the
well known merging galaxy NGC 7252, 50% of the cold gas
currently in the tidal tails is expected to fall back into the
main body of the galaxy within the next 2 -- 3 Gyr (Hibbard
& van Gorkom 1996). This suggests that we should expect
to see a rapid build up of X -- ray emission in the first few
gigayears after merger. This increase should be made even
more noticeable by the fading of the stellar population, as
this is the period during which LB is dropping most quickly.
Rapid generation of the X -- ray halo is inconsistent with our
results.
A model involving infall of hot gas is likely to have sim-
ilar problems. It is difficult to see how gas falling back into
the body of the galaxy can do so at the steady rate needed to
produce a smooth increase in LX /LB. Infall may be delayed,
as the hot gas reaches larger radii than the cold, but it will
still occur on a galaxy dynamical timescale. Models involv-
ing cooling and infall of gas onto central dominant galaxies
in groups or clusters (Brighenti & Mathews 1999) may be
able to produce an inflow over a longer period, but these are
unlikely to be generally applicable. Group dominant galax-
ies have exceptionally large dark halos, making them more
able to retain (or accrete) hot gas. They also lie at the bot-
tom of a group potential well, and may be at the centre of
a group X -- ray halo, providing them with an extra reservoir
of hot gas. These conditions do not apply to the majority of
early-type galaxies. Therefore we suggest that infall of (cold
or hot) gas is unlikely to be the dominant process in the
generation of X -- ray halos in typical ellipticals.
6
Ewan O'Sullivan et al.
11
10.5
10
9.5
9
0
5
10
15
Figure 3. Optical luminosity versus spectroscopic age. Crossed circles represent cluster or group dominant galaxies and filled circles
all other detections. The triangle is the recent merger remnant NGC 7252. The cE galaxy, M32, has been excluded from the plot and
associated statistical tests.
3.2 Ongoing stellar mass -- loss and galaxy winds
Since infall of either hot or cold gas seems unable to provide
the sort of long term trend in LX/LB seen in Fig. 2, we now
look to models involving gas generated by mass loss from the
galactic stars. This source of gas is fairly well understood,
and can provide the sort of gas masses required to account
for the halos of many early-type galaxies (though proba-
bly not the very brightest -- Brighenti & Mathews 1998),
provided that the gas is retained within the galactic poten-
tial. Two main timescales are involved. One is the mass loss
rate from stars, primarily giant stars. The second is the rate
of type Ia supernovae (SNIa) which provide the main heat
source (after the brief SNII phase). These two factors, and
the interplay between them, can potentially lead to changes
in the hot gas content of elliptical galaxies on timescales
much longer than a dynamical time, and so have the poten-
tial to explain the slow trend we observe. The stellar mass
loss rate is fairly well understood (e.g. Ciotti et al. 1991), and
for a single-aged stellar population it declines approximately
as t−1.3. However, the SNIa heating rate, which dominates
the effective specific energy of the injected gas, is highly un-
certain. Recent estimates of the SNIa rate in old stellar pop-
ulations (e.g. Cappellaro et al. 1999) have revised the classic
Tammann (1982) value of the rate in old stellar populations
down by a factor 3-4, and the evolution of this rate with
population age is very model-dependent. Given our contin-
uing ignorance about the precise nature of SNIa, it must
therefore be regarded as largely unknown.
The specific energy of the injected gas determines
whether gas is retained in a hydrostatic hot halo, or escapes
the galaxy as a wind. The way in which the specific energy
evolves, depends upon the evolution of the SNIa rate rela-
tive to the stellar wind loss rate. There are therefore two
fundamentally different classes of models: those in which
the SNIa rate is constant, or declines more slowly than the
stellar mass loss rate (e.g. Loewenstein & Mathews 1987;
David et al. 1991) and those in which the supernova rate
drops more quickly than the gas injection rate (Ciotti et al.
1991; Pelegrini & Ciotti 1998). Broadly speaking, when the
specific energy of the injected gas is smaller than its grav-
itational binding energy it will be retained in a hot halo.
It may subsequently cool in a luminous cooling flow. How-
ever, if the specific energy of the gas substantially exceeds
its binding energy it will escape from the galaxy in a fast
c(cid:13) 0000 RAS, MNRAS 000, 1 -- 8
X -- rays from Post -- Mergers
7
32
31
30
29
-0.8
-0.6
-0.4
-0.2
0
0.2
0.4
Figure 4. X -- ray luminosity versus Fundamental Plane residual for early -- type galaxies. Filled circles represent detections and arrows
upper limits. The solid line is a fit to all points with -- 0.5<FP residual<0.5
wind, which results in a much lower X -- ray luminosity. The
first class of models, in which the specific energy rises with
time, therefore produces an evolution from an early inflow
phase to a later wind phase, predicting a declining X-ray
luminosity with time. Such models are clearly inconsistent
with our results.
Models in which the SNIa rate drops more quickly than
t−1.3 evolve, in broad terms, from a low luminosity wind
phase, towards a more luminous halo/cooling flow phase.
The timescale for this evolution depends upon a variety of
factors, such as the depth and shape of the potential, but it
is typically ∼ 10 Gyr (Ciotti et al. 1991). In fact, none of
the models studied to date shows the rather simple mono-
tonic rise in LX with age which we observe. The models of
Ciotti et al. (1991) predict an initial luminous phase, when
the gas loss rate from stars is very high. The X -- ray luminos-
ity then declines as the wind loss rate drops, and then rises
again during the transition to a bound hydrostatic halo, at
which point the X-ray luminosity is essentially equal to the
SNIa luminosity. This whole development describes a fairly
symmetrical dip and rise, lasting ∼ 5 − 15 Gyr, which is
again not what we see in the data. However, this model de-
scribes a galaxy in which all stars are formed at t = 0. In
c(cid:13) 0000 RAS, MNRAS 000, 1 -- 8
contrast, the spectroscopic age of our galaxies probably de-
notes the time since a merger-induced starburst involving
only a few percent of the stellar mass. In this case, the early
stellar mass loss rate will be much lower than in the Ciotti
et al models, and the X-ray luminosity correspondingly less
(scaling approximately as the square of the mass loss rate).
The main change in LX will be a rise associated with the
slowing wind and transformation into a hydrostatic halo. In
addition, as shown in Fig. 2, the decrease in optical lumi-
nosity as the starburst population fades will produce a rise
in LX /LB over the first 1-2 Gyr.
4 CONCLUDING REMARKS
We have examined the evolution of the X -- ray properties
of early -- type galaxies. For three galaxy age estimators (fine
structure parameter, Fundamental Plane residuals and spec-
troscopic ages) we find that the normalized X -- ray luminos-
ity evolves with time. In particular, the X -- ray luminosity,
which reflects the mass of hot halo gas, appears to increase
at a steady rate over ∼ 10 Gyrs.
Comparing the long term trend in LX/LB which we
8
Ewan O'Sullivan et al.
observe with expectations from possible mechanisms for hot
halo formation, we conclude that the only viable mechanism
appears to be the slow evolution from an outflowing wind
to hydrostatic halo phase driven by a declining SNIa rate.
Infalling gas seems unlikely to be the main cause of such
a long term trend. Our results suggest that some of the
scatter seen in the global LX versus LB relation is due to
the evolutionary state, and past merger history, of early --
type galaxies.
Hibbard J. E., Guhathakurta P., van Gorkom J. H.,
Schweizer F., 1994, AJ, 107, 67
Huchtmeier W. K., Richter O.-G., 1988, A&A, 203, 237
Huchtmeier W. K., Richter O.-G., 1989, A&A, 210, 1
Irwin J., Sarazin C., 1998, ApJ, 494, L33
Loewenstein M., Mathews W. G., 1987, ApJ, 319, 614
Acknowledgements
We would like to thank R. Brown for early work on this
project and S. Helsdon, W. Mathews and A. Renzini for
useful discussions. We also thank an anonymous referee for
several suggestions which have improved the paper. This
research has made use of the NASA/IPAC Extragalactic
Database (NED) which is operated by the Jet Propulsion
Laboratory, California Institute of Technology, under con-
tract with the National Aeronautics and Space Adminis-
tration. E. O'S. acknowledges the receipt of a PPARC stu-
dentship.
REFERENCES
Mackie G., Fabbiano G.,
in Arnaboldi M.,
Da Costa G. S., Saha P., eds, ASP Conf. Ser. 116:
The Nature of Elliptical Galaxies; 2nd Stromlo Sym-
posium. p. 401
1997,
Mathews W. G., Brighenti F., 1998, in ASP Conf. Ser. 136:
Galactic Halos. p. 277
Matsushita K., Ohashi T., Makishima K., 2000, PASJ, 52,
685
Matsushita K., 2000, preprint, astro-ph/0009276
O'Sullivan E., Forbes D. A., Ponman T. J., 2000, in prepa-
ration
Pelegrini S., Ciotti L., 1998, A&A, 333, 433
Bender R., Surma P., Dobereiner S., Mollenhoff C., Made-
Prugniel P., Simien F., 1996, A&A, 309, 749
jsky R., 1989, A&A, 217, 35
Bregman J. N., Hogg D. E., Roberts M. S., 1992, ApJ, 387,
484
Brighenti F., Mathews W. G., 1998, ApJ, 495, 239
Brighenti F., Mathews W. G., 1999, ApJ, 512, 65
Read A. M., Ponman T. J., 1998, MNRAS, 297, 143
Sansom A. E., Hibbard J., Schweizer F., 2000, AJ, 120, 1946
Sarazin C. L., Irwin J. A., Bregman J. N., 2000, preprint,
astro-ph/0009448
Schweizer F., Seitzer P., 1992, AJ, 104, 1039
Cappellaro E., Evans R., Turatto M., 1999, A&A, 351, 459
Carollo C. M., Danziger I. J., Buson L., 1993, MNRAS, 265,
553
Tammann G. A. in Rees M., Stoneham R., eds, Supernovae:
A Survey of Current Research, Dordrecht: Reidel,
1982
Ciotti L., D'Ercole A., Pelegrini S., Renzini A., 1991, ApJ,
Terlevich A. I., Forbes D. A., 2000, MNRAS, submitted
376, 380
Worthey G., Faber S. M., Gonzalez J. J., Burstein D., 1994,
David L. P., Forman W., Jones C., 1991, ApJ, 369, 121
ApJS, 94, 687
Fabbiano G., Schweizer F., 1995, ApJ, 447, 572
Fabbiano G., Kim D. W., Trinchieri G., 1992, ApJS, 80, 531
Fabbiano G., Kim D.-W., Trinchieri G., 1994, ApJ, 429, 94
Feigelson E. D., Nelson P. I., 1985, ApJ, 293, 192
Forbes D., Ponman T., Brown R., 1998, ApJ, 508, L43
Garcia A. M., 1993, A&AS, 100, 47
Georgakakis A., Forbes D. A., Norris R. P., 2000, MNRAS,
318, 124
Helsdon S. F., Ponman T. J., O'Sullivan E., Forbes D. A.,
2000, MNRAS, submitted
Hibbard J. E., van Gorkom J. H., 1996, AJ, 111, 655
c(cid:13) 0000 RAS, MNRAS 000, 1 -- 8
|
astro-ph/0209100 | 1 | 0209 | 2002-09-05T14:42:55 | Bose-Einstein condensation in dense nuclear matter and strong magnetic fields | [
"astro-ph",
"nucl-th"
] | Bose-Einstein condensation of antikaons in cold and dense beta-equilibrated matter under the influence of strong magnetic fields is studied within a relativistic mean field model. For magnetic fields $> 5 \times 10^{18}$G, the phase spaces of charged particles are modified resulting in compositional changes in the system. The threshold density of $K^-$ condensation is shifted to higher density compared with the field free case. In the presence of strong fields, the equation of state becomes stiffer than that of the zero field case. | astro-ph | astro-ph | Bose-Einstein condensation in dense nuclear matter and strong
magnetic fields
Prantick Dey(a), Abhijit Bhattacharyya(b) and Debades Bandyopadhyay(c)
(a)Physics Department, Raja Peary Mohan College, Uttarpara, Hooghly 712258, India
(b)Physics Department, Scottish Church College, 1 & 3 Urquhart Square, Calcutta 700006, India
(c)Saha Institute of Nuclear Physics, 1/AF Bidhannagar, Calcutta 700 064, India
Abstract
Bose-Einstein condensation of antikaons in cold and dense beta-equilibrated
matter under the influence of strong magnetic fields is studied within a rela-
tivistic mean field model. For magnetic fields > 5 × 1018G, the phase spaces
of charged particles are modified resulting in compositional changes in the
system. The threshold density of K − condensation is shifted to higher den-
sity compared with the field free case. In the presence of strong fields, the
equation of state becomes stiffer than that of the zero field case.
PACS: 03.75.Fi, 26.60.+c, 21.65.+f
2
0
0
2
p
e
S
5
1
v
0
0
1
9
0
2
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
1
Recently, it has been inferred that some soft gamma ray repeaters (SGRs) and perhaps
certain anomalous X-ray pulsars (AXPs) could be neutron stars having large magnetic fields
∼ 1014 − 1016 G [1]. Those objects are called "magnetars" [2]. Earlier large magnetic fields
∼ 1013G were estimated to be associated with the surfaces of some radio pulsars [3]. The
origin of such ultra strong magnetic fields is still an unsolved problem. An attractive idea
about the origin is that the small magnetic field of a progenitor star is amplified due to
the magnetic flux conservation during the gravitational collapse of the star [3]. Recently,
Thomson and Duncan argued that a convective dynamo mechanism might result in large
fields ∼ 1015G [4]. On the other hand, it is presumed from the scalar virial theorem [5]
based on Newtonian gravity that the limiting interior field in neutron stars could be as large
as ∼ 1018G [6]. From the general relativistic calculation of axis-symmetric neutron stars in
magnetic fields, it follows that neutron stars could sustain magnetic fields ∼ 1018G [7,8].
Because of highly conducting core, such large interior fields may be frozen and could not
be directly accessible to observation. Its effects may be manifested in various observables
such as the mass-radius relationship, neutrino emissivity etc. Motivated by the existence of
large fields in the core of neutron stars, its influence on the gross properties of neutron stars
was studied by various groups [6,9,10,11]. The calculations in the relativistic mean field
(RMF) approach showed that the equation of state (EoS) was modified due to the Landau
quantization and also by the interaction of magnetic moments of baryons with the field [11].
The intense magnetic field was found to change the composition of beta equilibrated matter
relevant to neutron stars drastically [9,10,11]. The neutrino emissivity in neutron stars was
reported to be enhanced in strong magnetic fields [12].
Besides strong interior fields, many exotic forms of matter may exist in the dense core
of neutron stars. One such possibility is the appearance of the Bose-Einstein condensate of
strange particles. Nelson and Kaplan first pointed out that antikaons may undergo the Bose-
Einstein condensation (BEC) in dense matter at zero temperature because of the attractive
s-wave antikaon-nucleon interaction [13]. Later, this idea was applied to neutron stars by
various authors [14,15,16]. Bose-Einstein condensation in a magnetic field is an old and
interesting problem in other branches of physics also namely condensed matter physics and
statistical physics. It was shown by Schafroth [17] that a non-relativistic Bose gas could not
condense in an external magnetic field. There are some calculations on the condensation
of relativistic charged Bose gas in magnetic fields in the literature [18,19]. Elmfors and
collaborators [18] noted that the relativistic Bose gas might condense for spatial dimension
d ≥ 5. They showed that the number density of bosons in the ground state diverges for
d < 5 in the presence of a magnetic field. On the other hand, it was argued [19] that the
condensation of bosons in a magnetic field could occur in three dimension if the chemical
potential of bosons was taken as a function of density, temperature and magnetic field [19].
In this case, the BEC would be a diffuse one because there is no definite critical temperature.
It was also shown in the latter calculation [19] that the number density of bosons in the
ground state was finite. Recently, Suh and Mathews [10] have studied pion condensation in
a beta equilibrated non-interacting n-p-e-µ system in magnetic fields.
In this paper, we investigate the influence of strong magnetic fields on the Bose-Einstein
condensation of antikaons in cold and dense matter relevant to neutron stars. This may have
profound implications on the gross properties of neutron stars. This method of studying the
BEC in strong magnetic fields and dense matter is rather general; therefore it should be of
2
correspondingly broad interest.
We consider strong magnetic field effects on antikaon condensation in the beta equili-
brated neutron star matter composed of neutrons, protons, electrons, muons and K − mesons
within the framework of a relativistic field theoretical model [20]. As the constituents in
neutron stars are highly degenerate, the chemical potentials of baryons are larger than the
temperature of the system. Therefore, the gross properties of neutron stars are calculated
at zero temperature. The total Lagrangian density may be written as the sum of baryonic,
kaonic and leptonic parts i.e. L = LB + LK + Ll. In a uniform magnetic field, the baryonic
Lagrangian density [21] is given by
¯ψB (iγµDµ − mB + gσBσ − gωBγµωµ − gρBγµtB · ρµ − κBσµνF µν) ψB
LB = XB=n,p
1
+
2 (cid:16)∂µσ∂µσ − m2
− Xk=ω,ρ(cid:20) 1
4 (cid:16)∂µV k
σσ2(cid:17) − U(σ)
µ(cid:17)2
ν − ∂νV k
−
1
2
m2
k(V k
µ )2(cid:21) +
1
4
g4 (ωµωµ)2 −
1
4
F µνFµν .
(1)
Here ψB denotes the Dirac spinor for baryon B with vacuum mass mB and isospin operator
tB. The scalar self-interaction term [22] is, U(σ) = g2σ3/3 + g3σ4/4. Following Ref. [11],
the interaction of anomalous magnetic moments of baryons with magnetic fields is given
by the last term under the summation in Eq.(1). Here, F µν is the electromagnetic field
tensor, σµν = [γµ, γν]/2 and κB is the experimentally measured value of magnetic moment
for baryon B. The covariant derivative for a charged particle is Dµ = ∂µ + iqAµ with the
choice of gauge corresponding to the constant magnetic field (Bm) along z-axis is A0 = 0,
A ≡ (0, xBm, 0). The form of 4-component spinor solutions for baryons is given by Ref. [11].
The (anti)kaon-nucleon interaction is treated in the same footing as that of the nucleon-
nucleon interaction [15]. Therefore, the kaonic Lagrangian density in a magnetic field is
given as,
K K ∗K ,
LK = D∗
µK ∗DµK − m∗2
(2)
where the covariant derivative Dµ = ∂µ + iqAµ + igωKωµ + igρK tK · ρµ. There is
no interaction term involving magnetic moments in the kaonic Lagrangian density be-
cause (anti)kaons having zero spin angular momentum do not possess magnetic mo-
ments. The effective mass of (anti)kaons in this minimal coupling scheme is given by
m∗
K = mK − gσKσ. The solution for negatively charged kaons in a magnetic field is
K ∝ (qBm/π)1/4 (1/√2nn!)e−iωK− t+ipyy+ipzze−qBmη2/2Hn(√qBmη), where η = x + py/qBm,
"H" denotes the Hermite polynomial with n the Landau principal quantum number. The
Lagrangian density for neutrons is obtained by putting q = 0 in the covariant derivatives of
Eq. (1). In the mean field approximation [20], the meson field equations in the presence of
antikaon condensate and magnetic field are
m2
ωω0 + g4ω3
m2
σσ = −
0 = XB
ρρ03 = XB
m2
m∗
K
K + qBm
∂U
∂σ
gσBnS
B + gσK
+XB
gωBnB − gωKnK − ,
gρBI3BnB + gρKI3K −nK − .
qm∗2
nK − ,
(3)
(4)
(5)
3
where nB and ns
B are baryon and scalar density for baryon B respectively; I3B = +1/2 for
protons, −1/2 for neutrons and I3K − = −1/2 for K − mesons. The expressions of the scalar
and baryon density corresponding to protons are given by [11]
p = qpBm
ns
2π2 Xν Xs
m∗
p
mp
mp − sκpBm
ln (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
and
Ep
f + kp
mp
f,ν,s
! ,
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
np = qpBm
2π2 Xν Xs
kp
f,ν,s ,
where the energy spectrum for protons is given by
Ep,ν,s = rk2
z +(cid:16)qm∗ 2
p + 2νqpBm + sκpBm(cid:17)2
+ gωpω0 +
1
2
gρpρ03 ,
and
mp = qm∗ 2
p + 2νqpBm + sκpBm ,
Similarly for neutrons, those expressions are given by [11]
kp
f,ν,s = rEp 2
ns
n =
m∗
n
4π2 Xs
nn =
1
2π2 Xs
1
3
k3
f,s +
1
2
.
f −(cid:16)qm∗ 2
kf,sEn
p + 2νqpBm + sκpBm(cid:17)2
f − m2 ln (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
m (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
f arcsin
sκnBm"mkf,s + En 2
f + kf,s
! ,
En
m
f −
En
π
2!# ,
(6)
(7)
(8)
(9)
(10)
(11)
(12)
(13)
(14)
(15)
and
where
and
En,s = rk2
z +(cid:16)qm∗ 2
n + k2
x + k2
y + sκnBm(cid:17)2
+ gωnω0 −
1
2
gρnρ03 ,
m = m∗
n + sκnBm ,
kf,s = qEn 2
f − m2 .
magnetic field is obtained as ωK − = qp2
Solving the equation of motion for antikaons, the in-medium energy of K − meson in a
K + qBm(2n + 1) − gωKω0 − gρKρ03/2. The
condition for the condensation of K − meson in a magnetic field is pz = 0 and n = 0.
The number density of K − meson in a magnetic field and in the ground state is obtained
z + m∗2
4
from the relation J K
2(ωK − + gωKω0 + gρKρ03/2)K ∗K. The total energy density is given by
µ = i(K ∗∂L/∂µK ∗ − ∂L/∂µK K) and it is given by, nK − = −J K −
0 =
ε =
1
2
m2
σσ2 +
+ XB=n,p
εB +Xl
1
3
g2σ3 +
1
4
g3σ4 +
1
2
m2
ωω2
0 +
3
4
g4ω4
0 +
1
2
m2
ρρ2
03
εl + ε ¯K ,
(16)
where εB and εl correspond to the kinetic energy densities of baryons and leptons respec-
tively. The kinetic energy densities of protons and neutrons in a magnetic field are given by
[11]
εp = qpBm
4π2 Xν Xs
and
Ep
f kp
f,ν,s + mp
εn =
1
1
2
4π2 Xs
+ (cid:18)1
sκnBm −
3
En 3
f kf,s +
sκnBmEn 3
2
3
1
4
m(cid:19)"mkf,sEn
Ep
2 ln (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
f arcsin
f + m3 ln (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
En
f + kp
mp
f,ν,s
! ,
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
π
m
f −
En
f + kf,s
2!
m (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
!# .
(17)
(18)
Similarly, the expression for the kinetic energy density of electrons has the same form as that
of protons but electrons are noninteracting and anomalous magnetic moment of electrons
is not considered here [11]. The energy density for antikaons in the condensate state is
ε ¯K = qm∗2
K + qBm nK −. The other terms in Eq. (16) represent interaction energy densities.
The pressure of the system follows from the relation P = µnnb − ε, where µn and nb are the
neutron chemical potential and total baryon density, respectively. In the core of neutron
stars, strangeness changing processes such as n ⇀↽ p + K − and e− ⇀↽ K − + νe occur. The
chemical equilibrium yields µn − µp = µK − = µe, where µp and µK − are respectively the
chemical potentials of protons and K − mesons. Employing Eq. (3) in conjunction with the
chemical equilibrium conditions and charge neutrality np − nK − − ne − nµ = 0, we obtain
the effective masses self-consistently.
In the effective field theoretical approach adopted here, two different sets of coupling
constants for nucleons and kaons with σ, ω and ρ meson are required. The nucleon-meson
coupling constants are obtained by fitting experimental data for binding energies and charge
radii for heavy nuclei [21]. This set of parameters is known as TM1 set. The values of
coupling constants are gσN = 10.0289, gωN = 12.6139, gρN = 4.6322, g2 = −7.2325f m−1,
g3 = 0.6183 and g4 = 71.3075. The incompressibility of matter at normal nuclear matter
density (n0 = 0.145f m−3) is 281 MeV for the TM1 model. According to the simple quark
model and isospin counting rule, the kaon-vector meson coupling constants are gωK = 1
3gωN
and gρK = gρN . On the other hand, the scalar coupling constant is obtained from the
real part of the antikaon optical potential at normal nuclear matter density i.e. U ¯K(n0) =
−gσKσ − gωKω0 [16]. In this calculation, we have taken U ¯K(n0) = −160 MeV and the scalar
coupling is gσK = 2.0098.
The TM1 model was adopted earlier for the description of heavy nuclei and the equation
of state for neutron stars [21]. Besides the non-linear σ meson terms, the model also includes
5
c are denoted by the dashed lines. The critical field for electrons (Be
non-linear ω meson term. It was shown [21] that the TM1 model reproduced scalar and
vector potentials close to those of the relativistic Brueckner Hartree Fock calculation using
the realistic nucleon-nucleon interaction [23]. Recently, the TM1 model was used for the
investigation of antikaon condensation in neutron star matter [16] for zero field. For TM1
parameter set, it was found that the phase transition was of second order [15,16]. In the TM1
model, the maximum masses and central densities of neutron stars without and with antikaon
condensation where UK − = −160 MeV are respectively, 2.179(1.857)M⊙ and 5.97(6.37)n0.
In Figure 1, number densities of various particles are plotted with baryon density. The
particle densities for Bm = 0 are shown by the solid lines, whereas those corresponding to
Bm = 1.5 × 105Be
c ) is
that value where cyclotron quantum is equal to or above the rest energy of an electron and
its value is Be
c = 4.414 × 1013G. Here we note that the formation of K − condensation is
delayed to higher density than the field free case. The threshold densities of K − condensation
corresponding to Bm = 0 and Bm = 1.5 × 105Be
c are 2.67n0 and 3.85n0 respectively. The
delayed appearance of K − condensation may be attributed to the stiffer EoS because of the
effects of magnetic moments. In the presence of the field, the enhancement of electron and
muon fraction are pronounced whereas the proton fraction is smaller than the zero field value
beyond 2.7n0. With the appearance of K − condensate, it would try to diminish electron and
muon density. On the other hand, the phase spaces of electrons and muons are so strongly
modified in a quantizing field that their fractions are significantly increased. The net result
is the reduction in the density of K − condensate than that of the field free case. The proton
density increases after the onset of K − condensation. The neutron fraction also increases
because of the interaction of anomalous magnetic moment of neutrons with the field. This
may have important effects on the equation of state.
In the presence of magnetic fields > 5 × 1018G, the nucleon effective mass is enhanced
in the high density regime than that of the field free case. This may be attributed to the
effects of magnetic moments as it was also noted in Ref. [11]. The (anti)kaon effective mass
in magnetic fields does not change appreciably from the zero field case.
The onset of K − condensation is given by the equality of K − chemical potential (µK −)
with electron chemical potential (µe). In the presence of magnetic field, we find the hadronic
phase smoothly connects to the antikaon condensate phase resulting in a second order phase
transition as it is evident from equation of state (pressure versus energy density curve) in
Figure 2. For TM1 parameter set, we note that the phase transition is of second order with
and without magnetic field.
c (curve II) and Bm = 1.5× 105Be
In Figure 2, matter pressure (P ) versus matter energy density (ǫ) is displayed for Bm = 0
c (curve III). For Bm = 4× 104Be
(curve I), Bm = 4× 104Be
c ,
we note that the curve becomes slightly stiffer with the onset of K − condensation. This
stiffening may be attributed to the large enhancement in electron and muon fraction in the
field. However, this effect is reduced in the high density regime where electron and muon
fraction become small. As the field is further increased to Bm = 1.5 × 105Be
c , not only
electrons and muons are strongly Landau quantized, but also protons are populated in the
zeroth Landau level. It was shown [9,11] that Landau quantization of charged particles was
responsible for the softening in the equation of state. On the other hand, the effects of baryon
magnetic moments for Bm = 1.5 × 105Be
c overwhelm the effects of Landau quantization.
Consequently, the curve corresponding to Bm = 1.5× 105Be
c stiffens further. It is found here
6
m/(8π) = 4.814 × 10−8(Bm/Be
that the effects of magnetic moments are important for Bm > 105Be
c . Besides the effects
of Landau quantization and magnetic moments, the contribution of electromagnetic field to
the matter energy density and pressure is to be taken into account. The magnetic energy
density and pressure, εf = Pf = B2
c )2MeV f m−3, become
significant in the core of the star for Bm ≥ 105Be
c .
In this calculation, we have considered interior magnetic field > 5 × 1018G. However, it
was found in a recent calculation [24] that the maximum value of the magnetic field within
a star may not exceed 3 × 1018G for a particular choice of a constant current function but
independent of an EoS. In this case, the ratio of the maximum field to the average field is not
large because of small spatial gradient. The authors [24] argued that the value of maximum
field at any point may well exceed the average value as mentioned above for a different field
geometry. In that event the effects of strong magnetic field > 5 × 1018G on the threshold of
antikaon condensation, particle composition and EoS might be important.
To summarise, in this paper, we have focused on the formation of the antikaon conden-
sation in dense nuclear matter in the presence of magnetic fields. We have considered the
interaction of magnetic moments of baryons with the field and the magnetic energy density
and pressure in this work. In the presence of strong magnetic fields > 5 × 1018G, we find a
considerable change in the phase spaces of charged particles. The threshold density of K −
condensation is delayed to higher density in the presence of such a strong field and the EoS
becomes stiffer. For Bm > 1018G, the effects of magnetic moments are important and it adds
to further stiffening of the equation of state. Also, the electromagnetic field contribution
to the energy density and pressure becomes important in the core for Bm > 1018G. The
stiffening of the EoS in the presence of magnetic fields might have significant impact on
the gross properties of neutron stars such as the mass-radius relationship, cooling etc. It
is worth mentioning here that (anti)kaons do not interact with magnetic fields in the same
way as fermions do because their spin angular momentum is zero.
In this calculation, we do not include the role of hyperons, pion condensation and nucleon-
nucleon correlation on the antikaon condensation. Negatively charged hyperons, in particular
Σ− hyperon, could delay the onset of K − condensation [14]. However, it was estimated that
Σ−-nucleon interaction is highly repulsive in normal nuclear matter [25]. Recently, it has
been also shown that threshold densities of most hyperons including Σ− are substantially
increased in strong magnetic field Bm > 5 × 1018G both due to Landau quantisation and
magnetic moment interactions [24]. In this situation, Σ− hyperons might have no impact on
K − condensation. Pion condensation could occur in neutron stars because of the attractive
p-wave pion-nucleon interaction [26]. The condensation of π− may modify the electron
chemical potential which, in turn, would delay K − condensation. In this paper, we have
employed the RMF model which does not include nucleon-nucleon correlations.
It was
shown in non relativistic models [27] that nucleon-nucleon correlations shifted the threshold
density of K − condensation to higher density. We believe that the qualitative features of
strong magnetic fields presented here would survive even in other models which include
hyperons, pion condensation and nucleon-nucleon correlations. It would be interesting to
look into the neutrino emissivity from an antikaon condensed matter and the structure of
compact stars having antikaon condensate in the presence of a strong magnetic field.
7
REFERENCES
[1] Kouveliotou C, Dieter S, Strohmayer T, Paradijs J van, Fishman G J, Meegan C A and
Hurley K 1998 Nature 393 235; Kouveliotou C, Strohmayer T, Hurley K, Paradijs J van,
Finger M H, Dieter S, Woods P, Thompson C and Duncan R C 1999 Astrophys. J. 510
L115; Heyl J S and Kulkarni S R, 1998 Astrophys. J. 506 L61; Chatterjee P, Hernquist
L and Narayan R, 2000 Astrophys. J. 534 373.
[2] Duncan R C and Thompson C 1992 Astrophys. J. 392 L9; Thompson C and Duncan R
C 1995 MNRAS 275 255; Thompson C and Duncan R C 1996 Astrophys. J. 473 322;
Usov V V 1992 Nature 357 472; Paczynski B 1992 Acta. Astron. 42 145; Vasisht G and
Gotthelf E V 1997 Astrophys. J. 486 L129.
[3] Chanmugam G 1992 Annu. Rev. Astron. Astrophys. 30 143.
[4] Thompson C and Duncan R C 1993 Astrophys. J. 498 194.
[5] Chandrasekhar S and Fermi E 1953 Astrophys. J. 118 116.
[6] Lai D and Shapiro S L 1991 Astrophys. J. 383 745.
[7] Bocquet M, Bonazzola S, Gourgoulhon E and Novak J 1995 Astron. Astrophys.301 301.
[8] Cardall C Y, Prakash M and Lattimer J M 2001 Astrophys. J. 554 322.
[9] Chakrabarty S, Bandyopadhyay D and Pal S 1997 Phys. Rev. Lett. 78 2898; Bandyopad-
hyay D, Chakrabarty S and Pal S 1997 Phys. Rev. Lett. 79 2176.
[10] Suh I -S and Mathews G J 2001 Astrophys. J. 546 1126.
[11] Broderick A, Prakash M and Lattimer J M 2000 Astrophys. J. 537 351.
[12] Leinson L B and Perez A 1998 JHEP 09 020; Bandyopadhyay D, Chakrabarty S, Dey
P and Pal S 1998 Phys. Rev. D58 121301; Baiko D A and Yakovlev D G 1999 Astron.
Astrophys. 342 192; Dalen E N E van, Dieperink A E L, Sedrakian A and Timmermans
R G E 2000 Astron. Astrophys. 360 549.
[13] Kaplan D B and Nelson A E 1986 Phys. Lett. B175 57; Nelson A E and Kaplan D B
1987 Phys. Lett. B192 193.
[14] Prakash M, Bombaci I, Prakash M, Ellis P J, Lattimer J M and Knorren R 1997 Phys.
Rep. 280 1.
[15] Glendenning N K and Schaffner-Bielich J 1998 Phys. Rev. Lett. 81 4564; Glendenning N
K and Schaffner-Bielich J 1999 Phys. Rev. C60 025803.
[16] Pal S, Bandyopadhyay D and Greiner W 2000 Nucl. Phys. A674 553; Banik S and Bandy-
opadhyay D 2001 Phys. Rev. C63 035802.
[17] Schafroth M R 1955 Phys. Rev. 100 463.
[18] Elmfors P et al. 1995 Phys. Lett. B348 462 and references therein.
[19] Perez R H 1996 Phys. Lett. B379 148 and references therein.
[20] Serot B D and Walecka J D 1986 Adv. Nucl. Phys. 16 1.
[21] Sugahara Y and Toki H 1994 Nucl. Phys. A579 557.
[22] Boguta J and Bodmer A R 1977 Nucl. Phys. A292 413.
[23] Brockmann R and Machleidt R 1990 Phys. Rev. C42 1965.
[24] Broderick A, Prakash M and Lattimer J M 2001 astro-ph/0111516
[25] Friedman E, Gal A, Mares J and Cieply A 1999 Phys. Rev. C60 024314.
[26] Akmal A, Pandharipande V R and Ravenhall D G 1998 Phys. Rev. C58 1804.
[27] Pandharipande V R, Pethick C J, Thorsson V 1995 Phys. Rev. Lett. 75 4567; Carlson J,
Heiselberg H and Pandharipande V R 2001 Phys. Rev. C63 017603.
8
1.000
0.100
0.010
)
3
-
m
f
(
i
n
Bm = 1.5 x 105 Bc
e
n
p
e-
-
K-
n
p
e-
-
K-
0.001
1
2
3
n/n0
4
5
6
Fig.1 : The particle abundances are plotted with normalised baryon density for Bm = 0 and
Bm = 1.5×105Be
c . Solid lines indicate particle abundances for field free case whereas dashed
c ) is 4.414 × 1013G.
lines denote those with the magnetic field. The critical electron field (Be
9
m
m
III
II
I
100
)
3
-
m
f
V
e
M
(
P
10
200
400
600
e (MeV fm -3)
800
1000
Fig.2 : The matter pressure (P ) is shown as a function of matter energy density (ε) for
different values of Bm. The field free case is shown by curve I (solid line) and curve II
(dash-dotted line) and curve III (dashed line) represent calculations for Bm = 4× 104Be
c and
Bm = 1.5 × 105Be
c , respectively. The critical electron field (Be
c ) is 4.414 × 1013G.
10
|
astro-ph/0107505 | 1 | 0107 | 2001-07-26T14:54:47 | High Speed photometry of faint Cataclysmic Variables: I. V359 Cen, XZ Eri, HY Lup, V351 Pup, V630 Sgr, YY Tel, CQ Vel, CE-315 | [
"astro-ph"
] | The first results of a photometric survey of faint Cataclysmic Variables are presented. V359 Cen is an SU UMa star with a period of 112 min. Even though observed at quiescence, the mass transfer rate in this old nova may be sufficiently high that in such a short period system (with its implied small mass ratio) the disc may be excited into an elliptical shape with the result that the observed brightness modulation gives a superhump period rather than an orbital period. XZ Eri is an eclipsing dwarf nova with an orbital period (P(orb)) of 88.1 min. HY Lup has only slight variability. V351 Pup, the remnant of Nova Puppis 1991, has P(orb) = 2.837 h and a light curve that strongly resembles that of the magnetic Nova Cyg 1975. V630 Sgr is the first nova remnant that has both positive superhumps (P(sh) = 2.980 h) and eclipses (P(orb) = 2.831 h). The YY Tel identification is somewhat uncertain. The correct identification for CQ Vel is provided from discovery of its flickering activity. The light curve of CE-315, a recently discovered AM CVn star, shows similarities to that of GP Com, with no apparent orbital modulation. | astro-ph | astro-ph |
Mon. Not. R. Astron. Soc. 000, 000 -- 000 (0000)
Printed 26 October 2018
(MN LATEX style file v1.4)
High Speed photometry of faint Cataclysmic Variables: I.
V359 Cen, XZ Eri, HY Lup, V351 Pup, V630 Sgr, YY Tel,
CQ Vel, CE-315
Patrick A. Woudt⋆ and Brian Warner†
Department of Astronomy, University of Cape Town, Private Bag, Rondebosch 7700, South Africa
ABSTRACT
The first results of a photometric survey of faint Cataclysmic Variables are presented.
V359 Cen is an SU UMa star with a period of 112 min. Even though observed at
quiescence, the mass transfer rate in this old nova may be sufficiently high that in
such a short period system (with its implied small mass ratio) the disc may be excited
into an elliptical shape with the result that the observed brightness modulation gives
a superhump period rather than an orbital period. XZ Eri is an eclipsing dwarf nova
with an orbital period (Porb) of 88.1 min. HY Lup has only slight variability. V351
Pup, the remnant of Nova Puppis 1991, has Porb = 2.837 h and a light curve that
strongly resembles that of the magnetic Nova Cyg 1975. V630 Sgr is the first nova
remnant that has both positive superhumps (Psh = 2.980 h) and eclipses (Porb =
2.831 h). The YY Tel identification is somewhat uncertain. The correct identification
for CQ Vel is provided from discovery of its flickering activity. The light curve of CE-
315, a recently discovered AM CVn star, shows similarities to that of GP Com, with
no apparent orbital modulation.
Key words: techniques: photometric -- binaries: eclipsing -- close -- novae, cataclysmic
variables
1
INTRODUCTION
Cataclysmic Variable stars (CVs) are a class of interact-
ing close binaries which attract great attention. They en-
able a quantitative study of the many processes involved in
mass transfer through accretions discs or through magneti-
cally channeled accretion. A list of these physical processes,
which occur also in many other astrophysical environments,
is given in Warner (2000); a review of CVs in general is given
in Warner (1995a).
Fewer than 40 CVs have been studied in any great de-
tail (e.g., with doppler tomography, eclipse deconvolution).
These, and the phenomena seen in other CVs, show that
no two CVs are exactly alike. The different combinations of
masses of primaries and secondaries, temperatures, magnetic
field strengths, orbital inclinations and mass transfer rates
( M ) suffice to produce great variety of behaviour. Much of
this variety is already understood in terms of the compo-
nent physics accompanying the gross differences. The effects
of more subtle differences -- e.g., original chemical composi-
tion of the secondary, effects on the secondary of past nova
⋆ E-mail: [email protected]
† E-mail: [email protected]
c(cid:13) 0000 RAS
explosions, compositional and environmental (for instance,
irradiation) dependence of disc viscosity -- have still to be
teased out of the rich observational phenomenology.
Much of the progress in the past two or three decades
has come from the discovery of ever more variety of be-
haviour. There are often conspicuous effects in the longer
term light curves; examples are the ER UMa stars (dwarf
novae with outbursts every few days and superoutbursts ev-
ery few weeks), 'echo' outbursts after a superoutburst, con-
tinuous small outbursts in the discs of nova-like variables.
New phenomena continue to be found also in the 'close-up'
details; examples are short period quasi-periodic oscillations
(from ∼2 s in the X-ray flux of SS Cyg in outburst to ∼1000
s modulations in many high M discs), spiral waves in discs,
tilted discs, desynchronisation of polars by nova explosions.
Although some of the new phenomena have accompa-
nied the discovery of new CVs (e.g., the ∼ 108 G magnetic
fields in AR UMa and V884 Her), most have been discovered
in long known CVs (e.g., TeV gamma rays from AE Aqr, ZZ
Cet pulsations in GW Lib). It is clear, therefore, that while
the great majority of CVs remain under-observed new phe-
nomena may remain undiscovered. At a less exotic level, the
determination of a significant number of new orbital periods,
and in particular the discovery of eclipsing systems, provide
objects for more detailed structural studies. With the ad-
2
Patrick A. Woudt and Brian Warner
Table 1. Observing log.
Object
Type
Run No.
Date of obs. HJD of first obs.
(+2451000.0)
(start of night)
Length
(h)
tin
(s)
Tel.
<V>
(mag)
V359 Cen
SU UMa
XZ Eri
DN
HY Lup
NR
V351 Pup NR
V630 Sgr
NR
YY Tel
SU UMa
CQ Vel
NR
CE-315
AM CVn
S6165
S6167
S6170
S6186
S6189
S6197
S6216
S6156
S6158
S6163
S6191
S6157
S6160
S6162
S6185
S6192
S6124
S6126
S6128
S6130
S6100
S6218
S6223
S6231
S6071
S6209
S6212
29 Dec 2000
30 Dec 2000
1 Jan 2001
20 Feb 2001
21 Feb 2001
25 Feb 2001
18 May 2001
26 Dec 2000
27 Dec 2000
29 Dec 2000
22 Feb 2001
26 Dec 2000
27 Dec 2000
28 Dec 2000
20 Feb 2000
23 Feb 2000
24 Aug 2000
25 Aug 2000
26 Aug 2000
27 Aug 2000
5 Jun 2000
18 May 2001
21 May 2001
26 May 2001
908.50289
909.51368
911.51125
961.37008
962.41615
966.35024
1048.22183
905.29165
906.29347
908.30081
963.47488
905.47151
906.45415
907.43106
961.27432
964.26594
781.22202
782.21315
783.21314
784.20912
701.36536
1048.61265
1051.61806
1056.64150
10 Mar 2000
614.26263
15 May 2001
16 May 2001
1045.35878
1046.25888
2.15
1.95
2.02
2.05
4.03
2.22
4.02
4.02
0.36
0.65
3.76
3.06
3.62
3.96
1.93
1.20
6.72
7.15
7.16
6.15
1.39
1.72
1.57
1.07
4.07
4.26
6.63
30
30
30
30, 60
60
60
60
20, 60
30
20, 30
45
15
20
20
60
60
10
5
5, 10
5, 10
10
30, 60
60
30, 60
74-in
74-in
74-in
40-in
40-in
40-in
40-in
74-in
74-in
74-in
18.8
18.7
18.7
18.6
18.7
18.6
18.6
19.1
19.2
19.1
40-in
19.6
74-in
74-in
74-in
40-in
40-in
74-in
74-in
74-in
74-in
74-in
40-in
40-in
74-in
18.9
19.1
18.9
18.8
19.0
17.6
17.5
17.8
17.4
18.6
18.6
18.5
18.6
60
10
20
74-in
21.1
40-in
40-in
17.6:
17.6:
Notes: NR = Nova Remnant, DN = Dwarf Nova, tin is the integration time.
vent of 8-m class telescopes such studies can be made even
to 20th magnitude. The ongoing survey, of which this paper
presents the first general results, was initiated with this in
mind.
The need for such a survey, particularly in the south-
ern hemisphere and capable of succeeding in the crowded
fields in the general direction of the Galactic centre where
the majority of nova remnants reside, was recognised in the
early 1990s and as a result the CV group at the University
of Cape Town (UCT) used several years' equipment funds
(in advance) for the purchase of a state-of-the-art CCD de-
tector. The photometer and its software were designed and
constructed by Dr. D. O'Donoghue, who was a member
of the Astronomy Department at UCT at that time. The
UCT CCD photometer, completed in 1993, has been exten-
sively used on telescopes at the Sutherland site of the South
African Astronomical Observatory. An outline of its func-
tional properties can be found in O'Donoghue (1995) and
Koen & O'Donoghue (1995). In the CV group at UCT it
was first largely used for observations of well known CVs at
higher time resolution or improved signal-to-noise. Its use
for a general survey of faint CVs started in 1997, with the
almost immediate discovery of the first ZZ Cet (i.e., non-
radial) pulsations in the white dwarf primary of the SU
UMa type dwarf nova GW Lib (Warner & van Zyl 1998)
and independent discovery of the intermediate polar nature
of the nova remnant HZ Pup (see Abbott & Shafter 1997 for
detailed description). Follow-up observations of these led to
neglect of the survey for some time, but we are now actively
observing faint CVs again. We present here initial results.
The observations are given in Section 2 and some brief
conclusions in Section 3.
2 OBSERVATIONS
All observations have been made with the UCT CCD pho-
tometer attached either to the 40-in or 74-in reflectors at
Sutherland. In order to capture as many photons as possible
'white light' was used, with the result that extinction cor-
rections and magnitude calibrations are only approximate.
The observing log for the stars included in this paper is
given in Table 1.
2.1 V359 Centauri
V359 Cen was discovered bright in April 1929‡ and entered
the lists as a probable nova, unconfirmed because no spec-
tra were obtained. There are still no spectra available at
‡ Note that Duerbeck (1987) states 1939, and this has been prop-
agated in later literature.
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
High Speed photometry of faint Cataclysmic Variables: I.
3
Figure 1. Light curves of V359 Cen at quiescence phased ac-
cording to the 112 min period. The light curves are plotted top
to bottom in the same order as in Table 1. All light curves, except
for the top one, are displaced along the magnitude axis for display
purposes only. The mean magnitude of V359 Cen for each run is
given in Table 1.
minimum light, but in recent years further outbursts to a
maximum of V ∼ 13.8 have been observed at intervals ∼ 1
y, showing the star to be of SU UMa type. The quiescent
magnitude given in the Downes, Webbink & Shara (1997)
catalogue is 21.0 from J plates, and Munari & Zwitter (1998)
gave V ≥ 20.5 in 1996, but we find V ∼ 18.7, which is a great
deal easier to observe at moderate time resolution. Our May
2001 observation was obtained three weeks after a superout-
burst, but the star had returned to its quiescent brightness.
Our light curves are displayed in Fig. 1 and show a re-
current hump with an amplitude ∼ 0.4−0.5 mag. Our obser-
vations are not distributed adequately to remove aliases; the
best determined period from the February 2001 run is 112
min (0.0779 days). The variations of hump profile and their
occasional almost triangular shape resemble more what is
seen in superhumps than orbital modulations, but it is rare
to observe superhumps at quiescence -- the only well docu-
mented examples are EG Cnc which sustained 'late' super-
humps for at least 90 d after the 1996 superoutburst (at a
time when 'echo' normal outbursts were occuring) (Patter-
son et al. 1998) and AL Com with a persistent superhump as
well as an orbital hump at quiescence (Patterson et al. 1996).
If what we are seeing is an orbital hump then the inclination
of V359 Cen must be ∼ 65◦.
2.2 XZ Eridanus
XZ Eri has been known as a probable dwarf nova for nearly
70 years (Shapley & Hughes 1934). In recent years only three
outbursts have been detected: Mar/Apr 1995, Jan 1998,
Feb 1999. The first of these reached V ∼ 14.1, the others
reached V ∼ 14.6. CCD photometry at quiescence gives V
= 18.7 (Howell et al. 1991). The rarity and range of the
outbursts suggests that they are superoutbursts. Any nor-
mal outbursts may have gone unnoticed. There has been no
high speed photometry during an outburst.
A quiescence spectrum, obtained by Szkody & How-
ell (1992) shows the Balmer emission and continuum char-
acteristic of a dwarf nova. There is a hint of doubling in
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
Figure 2. The mean light curve of XZ Eri.
the strongest Balmer lines, indicative of a high inclination
system. Nearly 4 h of CCD photometry through a V filter
at quiescence, with 220 s integration on a 1.8-m telescope,
showed a scatter of ∼ 1.0 mag and, to quote the authors, no
indications of periodic modulation (Howell et al. 1991).
Our observations (Fig. 2) show XZ Eri to be, in brief,
a Z Cha lookalike, with the succession of immersions and
emersions of the white dwarf and bright spot all quite clear.
At V ∼ 19 we cannot achieve sufficient time resolution to
separate these components as clearly as in Z Cha (e.g., Wood
et al. 1986). Using our best estimate for the time of mid-
eclipse of the white dwarf we derive the following ephemeris:
HJDmin = 2451905.4419 + 0.d0612 E.
(1)
The short Porb (1.47 h = 88.1 min) implies that XZ Eri
will have superhumps when observed in outburst.
2.3 HY Lupi
HY Lup was Nova Lupi 1993, which reached V ∼ 8 and was
moderately fast (t2 ∼ 15 d). It is currently at V ∼ 18.9 and
has a resolved ejecta shell (Downes & Duerbeck 2000). No
pre-outburst candidate brighter than V ∼ 17 is detectable
(McNaught & Garradd 1993) and it is probable that the
remnant has not yet reached its quiescent brightness. No
high speed photometry has been published for this nova rem-
nant.
Our photometry (Fig. 3) shows little evidence for rapid
flickering and only slow variations of small amplitude. There
is no suggestion of any orbital modulation which would en-
courage further photometry. The Porb for HY Lup will need
to be obtained spectroscopically.
2.4 V351 Puppis
Nova Puppis 1991 (V351 Puppis) was a moderately fast nova
(t2 = 16 d), probably a neon nova (Pachoulakis & Saizar
1995). The pre-eruption magnitude was variable over the
range 21− > 22 (McNaught 1992). The eruption light curve
through 1996 is illustrated in Saizar et al. (1996); in 1998
the remnant was at V ∼ 19.6 (Downes & Duerbeck 2000).
There are one or two additional aspects of this nova
that are of interest. Sixteen months after optical maximum,
4
Patrick A. Woudt and Brian Warner
Figure 3. The light curve of HY Lup obtained on 22 February
2001. The data are binned along the horizontal axis by a factor
of five.
hard X-rays were detected by ROSAT which, from the ab-
sence of a soft X-ray component, were deduced to be caused
either by shocks in the ejected shell or by magnetically con-
trolled accretion onto the white dwarf (Orio et al. 1996).
Variability (time scale < days) observed in the X-ray flux
seemed to favour the latter interpretation. However, a later
analysis explains the absence of soft X-rays as due to in-
ternal and external absorption, allowing the hard X-rays to
be the product of surface nuclear burning, as in most novae
(Vanlandingham et al. 2001). Multiwavelength observations
of the nebular phase of the eruption made in 1992-1993 re-
vealed a strong red excess attributed to the secondary star,
implying that this component is either a giant or an irradi-
ated dwarf (Saizar et al. 1996). In the former case an orbital
period Porb ∼ days would be expected, in the latter Porb <∼ 4
h would be required in order for the secondary to be close
enough to the hot primary to be strongly irradiated. The
relatively small mass of the ejected shell of N Puppis 1991
implies a high mass for the white dwarf -- near the Chan-
drasekhar limit (Vanlandingham et al. 2001).
We observed V351 Pup in December 2000 and found it
to be modulated in brightness with a range of up to ∼1.2
mag and a period of 2.837 h. The light curves are given in
Fig. 4 and show variable amplitude, low amplitude flickering,
rounded maxima and a V-shaped minimum which may be
filled in to form a flattened region. The ephemeris from these
observations is
HJDmax = 2451905.56746 + 0.d1182 E.
(2)
The additional observations we made in February 2001
are too far removed in time to enable us to improve on the
precision of the period.
Our light curves for V351 Pup have a striking similar-
ity to those of V1500 Cyg in the late 1980s (see especially
Fig. 1 of Kaluzny & Chlebowski 1988). V1500 Cyg was Nova
Cygni 1975 and in the late 1980s it was about 15.3 mag be-
low its maximum brightness. Nova Pup 1991 was not discov-
ered until after maximum, but there are reasons to believe
it did not exceed V ∼ 5.5 (Saizar et al. 1996) and it there-
fore at present is ∼14 mag below maximum. V1500 Cyg is
a desynchronised polar (Stockman, Schmidt & Lamb 1988;
Kaluzny & Chlebowski 1988) with an orbital period of 3.35
h, which suggests the possibility that V351 Pup may also
be magnetic. With this in mind we asked Gary Schmidt to
check for circular polarization and he reports "A sequence
of spectropolarimetric observations was obtained by G. and
P. Schmidt on 2 January 2001 using the 74-in reflector at
Mt. Stromlo and covering the range 4220 -- 7300 A at ∼10
Figure 4. The light curves of V351 Pup obtained in December
2000, phased according to Eqn. 2. A vertical shift of 0.5 mag has
been introduced between light curves for display purposes.
A resolution. The summed spectrum, which spans a total
of 3.7 h, reveals a weak continuum (V ∼ 19.4) that is ap-
proximately flat in fλ, plus broad (±1700 km s−1) nebular
emission lines of H, He II and O III. The polarimetric results
fail to show significant circular polarization anywhere in the
observational sequence, and the coadded sum is unpolarized
to a 3σ upper limit of 1%".
Before dismissing the possibility that V351 Pup is a
magnetic nova we note that (a) the level of circular polariza-
tion in V1500 Cyg is ∼ 1.5% (Stockman et al. 1988), which,
compared with the 10 -- 30% in polars, shows the presence of
much unpolarized light, probably that from irradiation of
the secondary by the still hot primary; (b) V351 Pup is 10 y
from its eruption and less far in its recovery than was V1500
Cyg in the late 1980s -- and should therefore have even more
unpolarised light; (c) GQ Mus (Nova Mus 1983) also has a
polar-like orbital light curve and shows strong spectral sig-
natures of magnetic accretion, but no circular polarization
(Diaz & Steiner 1994); (d) similar remarks apply to V2214
Oph (Nova Oph 1988: Baptista et al. 1993).
2.5 V630 Saggitarii
V630 Sgr was the last of four novae detected in Saggitarius
in 1936. It was discovered in October of that year, at a time
when Harvard objective prism plates were being exposed,
whereby it attained the status and alias HD 321353. The
observed maximum was V ∼ 4.5, but it probably reached V
∼ 1.6 (Downes et al. 1997), and at minimum it is V ∼ 19,
giving it the unusually large eruption amplitude of ∼ 17.5
mag, even for a t2 = 4 d nova, suggesting a high inclination
system (Warner 1986: this conclusion arises from the appar-
ent intrinsic faintness of the nearly edge-on accretion disc
which such an object has in quiescence).
There have been no modern photometric or spectro-
scopic observations of V630 Sgr, which is a faint object in a
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
High Speed photometry of faint Cataclysmic Variables: I.
5
Figure 5. The light curves of V630 Sgr obtained in August 2000, phased according to Eqn. 3. These have been binned to 30 s to give
greater clarity of reproduction. The light curves are plotted top to bottom in the same order as in Table 1. All light curves, except for
the top one, are displaced along the magnitude axis for display purposes only. The mean magnitude of V630 Sgr for each run is given in
Table 1.
V630 Sgr
Figure 6. The Fourier transform of the light curves shown in
Fig. 5.
very crowded field. It is listed by Diaz & Steiner (1991), on
the basis of its similarity to V1500 Cyg in being a very fast
nova of large amplitude, as a candidate magnetic system.
Our light curves are shown in Fig. 5 and illustrate that
V630 Sgr is indeed a moderately high inclination system,
with eclipses 0.4 -- 0.6 mag deep. The ephemeris derived from
our light curves, measured from the fundamental Fourier
component. The time of the first minimum of our first light
curve of this set is given in the ephemeris, which is
HJDmin = 2451781.2480 + 0.d1180 E.
(3)
The orbital period of 2.831 h is just at the top of the
orbital period gap.
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
It is evident from Fig. 5 that there is a hump of com-
plex and variable profile that has a period comparable to
Porb. The Fourier amplitude spectrum of the complete set
of light curves is shown in Fig. 6. The main groups of power
are associated with the fundamental and strong harmon-
ics of the narrow eclipses. The window pattern of the data
set can be judged from the first and second harmonics. Su-
perimposed, but slightly offset from the Porb fundamental
window pattern, is a second pattern. Prewhitening at the
orbital frequency gives a clear pattern with an amplitude of
0.07 mag centred on Psh = 2.980 h, which is the period of
the drifting humps. Therefore, V630 Sgr is a nova remnant
that is also an eclipsing permanent superhumper. This is the
first CV found to combine all of these features. The super-
hump excess of 5.3% and beat period of 2.4 d are typical
for superhumps in superoutbursting dwarf novae and nova-
like variables (Warner 1995a). The beat period is evident in
the night-to-night variation of mean brightness, as seen in
Fig. 7. The Fourier transform of this delivers an approximate
period of 2.1 d and an amplitude of 0.22 mag.
The eclipse depth of ∼ 0.6 mag is relatively shallow,
but at the Porb of V630 Sgr it implies a quite high orbital
inclination. The width at half depth of the narrowest eclipses
is ∆ψ ∼ 0.085Porb. The secondary mass M (2) at Porb = 2.8
h is ∼ 0.24 M⊙ (Warner 1995a) and as V630 Sgr was a
very fast nova we expect M (1) ∼ 1.2 M⊙, giving a mass
ratio q ∼ 0.20. These figures are not quite compatible with
the theoretical ∆ψ, q, i relationship (Horne 1985), but they
serve to show that i ≥ 85◦. The presence of superhumps
shows V630 Sgr to be a high M system which will have a
disc radius at the 3:1 resonant radius, which is rd = 0.45a.
The Roche lobe radius for the secondary is R(2) = 0.25a for
q = 0.2. Therefore, even at large inclination, the eclipses are
partial, as is seen in the eclipse profiles in Fig. 5.
The observed eruption range of 17.5 mag is 3.5 mag
greater than expected for a t2 = 4 d nova (see Fig. 5.4 of
6
Patrick A. Woudt and Brian Warner
Figure 7. The long term optical behaviour of V630 Sgr.
Figure 8. The light curve of YY Tel obtained on 6 June 2000. The
lower light curve shows a comparison star of the same brightness
(displaced by 0.25 mag for display purposes).
Warner 1995a). Using the correction formula (Eqn. 2.63 of
Warner 1995a) for disc inclinations, removes the 3.5 mag
excess if i ∼ 88◦. We conclude that V630 Sgr is a very high
inclination system.
Figure 9. CCD image of YY Tel (the brighter star of the two
indicated by markers) taken on 21 May 2001. The field of view is
109′′ by 74′′, north is up and east is to the left.
2.6 YY Telescopium
YY Tel is a dwarf nova with a range V ∼ 14.4−19.3 (Downes
et al. 1997) suspected to be an SU UMa star of long out-
burst interval (O'Donoghue et al. 1991). The last reported
outburst was in October 1998, and no high speed photom-
etry has been reported, either in outburst or at quiescence.
The candidate identified at quiescence (Downes et al. 1997)
was observed spectroscopically by Zwitter & Munari (1996)
and showed only a continuum with no emission lines char-
acteristic of a CV. Their photometry gave V = 19.26, (B-V)
= 0.47, (U-B) = -0.28, placing the candidate near the black
body line at T ∼ 6700 K in the two colour diagram. This is
also a region frequented by late F type subdwarfs.
Our observation on 5 June 2000 showed the YY Tel
candidate to be constant in brightness (see Fig. 8). Re-
examination nearly one year later again showed absence of
any rapid variation but the star changed in brightness by
0.15 mag over three to five days. A nearby, fainter star,
marked in Fig. 9 (YY Tel is the brighter of the two stars
that are marked), shows marginal evidence for variation and
should be examined spectroscopically.
2.7 CQ Velorum
CQ Vel was Nova Velorum 1940, reaching V ∼ 9.0. It has
been identified with a 21st magnitude star (Duerbeck 1987)
but no confirmatory spectra or photometry have been ob-
tained. In our observations on 10 March 2000 we found
the candidate nova remnant to be of constant brightness
as were all the stars in the CCD frame, except for a star
slightly fainter than the candidate, 9′′ to the northwest
(Fig. 10). This star has the flickering characteristic of a CV,
see Fig. 11. These observations provide a correct identifica-
tion for CQ Vel.
2.8 CE-315
CE-315 is a recently discovered member of the AM CVn
class of CVs: systems transferring helium rather than hy-
drogen (Ruiz et al. 2001). It is only the second of this class
to be found in a long-term state of low M , the other being
GP Com (alias G61-29). Ruiz et al. found a period of 65.1
± 0.7 min from an S-wave in the spectra, which by analogy
with GP Com is thought to be the orbital period.
We have obtained two light curves for CE-315, which are
displayed in Fig. 12. There is no sign in the light curve, or
its Fourier Transform, of the orbital periodicity. CE-315 re-
sembles GP Com in the general appearance of its light curve
(see Warner 1972; Harrop-Allin 1996): much short time scale
flaring or flickering set on a slowly varying background. In
GP Com there are occasions when a few cycles at the 46.5
min orbital period appear in the light curve (Harrop-Allin
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
High Speed photometry of faint Cataclysmic Variables: I.
7
result of radiation of gravitational waves (Warner 1995a,b).
According to Eqn. 9.61 of Warner (1995a) the mass transfer
M ∼ 1.0 × 10−11 M⊙ y−1 for GP Com
rates should be
and M ∼ 3 × 10−12 M⊙ y−1 for CE-315. GP Com has X-
rays modulated at the orbital period (Beuermann & Thomas
1993). At B ∼ 17.2, CE-315 is ∼ 1.4 mag fainter than GP
M , so they are
Com, but this is accounted for by the lower
at similar distances and it would be worth seeking an X-ray
detection.
3 CONCLUSIONS
The results presented here cover a miscellany of CV types
and demonstrate the rewards waiting among the fainter
members of this class. For those objects where new phe-
nomena have been found (e.g., the possible quiescent super-
humps in V359 Cen), or where rarity is an incentive (e.g.,
the eclipsing superhumping nova remnant V630 Sgr), further
observations with larger telescopes are required. V351 Pup
may show circular polarization or other magnetic signatures
when it has descended the nova light curve further.
ACKNOWLEDGMENTS
We thank Dr. D. O'Donoghue for the use of his EAGLE
program for Fourier analysis of the light curves. PAW is
funded partly through strategic funds made available to BW
by the University of Cape Town. BW's research is funded
entirely by that university.
REFERENCES
Abbott T.M.C., Shafter A.W., 1997, IAU Colloq. No. 163, 679
Baptista R., Jablonski F.J., Cieslinski D., Steiner J.E., 1993, ApJ,
406, L67
Beuerman K., Thomas H.-C., 1993, Adv.Sp.Res., 13, 115
Diaz M.P., Steiner J.E., 1991, PASP, 103, 964
Diaz M.P., Steiner J.E., 1994, ApJ, 425, 252
Downes R.A., Duerbeck H.W., 2000, AJ, 120, 2007
Downes R.A., Webbink R.F., Shara M.M., 1997, PASP, 109, 345
Duerbeck H.W., 1987, Sp.Sci.Rev., 45, 1
Harrop-Allin M.K., 1996, MSc thesis, University of Cape Town
Horne K., 1985, MNRAS, 213, 129
Howell S.B., Szkody P., Kreidl T.J., Dobrzycka D., 1991, PASP,
103, 300
Kaluzny J., Chlebowski T., 1988, ApJ, 332, 287
Koen C., O'Donoghue D., 1995, ApJS, 101, 347
McNaught R., 1992, IAU Circ. 5422
McNaught R., Garradd G.J., 1993, IAU Circ. 5868
Munari U., Zwitter T., 1998, A&AS, 128, 277
O'Donoghue D., 1995, Baltic Astr, 4, 517
O'Donoghue D., Chen A., Marang F., Mittaz J.P.D., Winkler H.,
Warner B., 1991, MNRAS, 250, 363
Orio M., Balman S., Della Valle M., Gallagher J., Ogelman H.,
1996, ApJ, 466, 410
Pachoulakis I., Saizar P., 1995, ApSpSciLibr, 205, 303
Patterson J., Augustein T., Harvey D.A., Skillman D.R., Abbott
T.M.C., Thorstensen J., 1996, PASP, 108, 748
Patterson J., Kemp J., Skillman D.R., Harvey D.A., Shafter
A.W., Vanmunster T., Jensen L., Fried R., Kiyota S.,
Thorstensen J.R., Taylor C.J., 1998, PASP, 110, 1290
Ruiz M.T., Rojo P.M., Garay G., Maza J., 2001, ApJ, 552, 679
Figure 10. CCD image of CQ Vel (indicated by the markers)
taken on 10 March 2000. The field of view is 50′′ by 34′′, north
is up and east is to the left.
Figure 11. The light curve of CQ Vel (V = 21.1) obtained on 10
March 2000.
1996); if the same occurs in CE-315 it would need extensive
photometry to detect.
The fact that the two longest period AM CVn stars, GP
Com and CE-315, are low M systems (equivalent to dwarf
novae in hydrogen transferring CVs) is in accord with the
evolutionary model where orbital angular momentum is the
Figure 12. The light curves of CE-315 obtained on 15 and 16
May 2001. The data in the top panel have been binned along the
horizontal axis by a factor of four, the data in the bottom panel
by a factor of two.
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
8
Patrick A. Woudt and Brian Warner
Saizar P., Pachoulakis I., Shore S.N., Starrfield S., Williams R.E.,
Rothschild E., Sonneborn G., 1996, MNRAS, 279, 280
Shapley H., Hughes E.M., 1934, HA, 90, No. 4
Stockman H.S., Schmidt G.D., Lamb D.Q., 1988, ApJ, 332, 282
Szkody P., Howell S.B., 1992, ApJS, 78, 537
Vanlandingham K.M., Schwarz G.J., Shore S.N., Starrfield S.,
2001, AJ, 121, 1126
Warner B., 1972, MNRAS, 159, 315
Warner B., 1986, MNRAS, 222, 11
Warner B., 1995a, Cataclysmic Variable Stars, Cambridge Univ.
Press, Cambridge
Warner B., 1995b, ApSpSci, 225, 249
Warner B., 2000, PASP, 112, 1523
Warner B., van Zyl L., 1998, IAU Symp. No. 185, 321
Wood J., Horne K., Berriman G., Wade R.A., O'Donoghue D.,
Warner B., 1986, MNRAS, 219, 629
Zwitter T., Munari U., 1996, A&AS, 117, 449
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
|
astro-ph/9810060 | 1 | 9810 | 1998-10-06T00:00:37 | The Ultraviolet Luminosity Density of the Universe from Photometric Redshifts of Galaxies in the Hubble Deep Field | [
"astro-ph"
] | Studies of the Hubble Deep Field (HDF) and other deep surveys have revealed an apparent peak in the ultraviolet (UV) luminosity density, and therefore the star-formation rate density, of the Universe at redshifts 1<z<2. We use photometric redshifts of galaxies in the HDF to determine the comoving UV luminosity density and find that, when errors (in particular, sampling error) are properly accounted for, a flat distribution is statistically indistinguishable from a distribution peaked at z~1.5. Furthermore, we examine the effects of cosmological surface brightness (SB) dimming on these measurements by applying a uniform SB cut to all galaxy fluxes after correcting them to redshift z=5. We find that, comparing all galaxies at the same intrinsic surface brightness sensitivity, the UV luminosity density contributed by high intrinsic SB regions increases by almost two orders of magnitude from z~0 to z~5. This suggests that there exists a population of objects with very high star formation rates at high redshifts that apparently do not exist at low redshifts. The peak of star formation, then, likely occurs somewhere beyond z>2. | astro-ph | astro-ph | The Ultraviolet Luminosity Density of the Universe from
Photometric Redshifts of Galaxies in the Hubble Deep Field1,2
Sebastian M. Pascarelle and Kenneth M. Lanzetta
Department of Physics and Astronomy, State University of New York at Stony Brook,
Stony Brook, NY 11794-3800
and
Alberto Fern´andez-Soto
School of Physics, University of New South Wales, Sydney, NSW2052, Australia
8
9
9
1
t
c
O
6
1
v
0
6
0
0
1
8
9
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
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 NAS 5-26555.
2Based in part on observations with the KPNO 4-meter Mayall Telescope of NOAO, which is operated
by AURA, Inc., under cooperative agreement with the NSF.
-- 2 --
ABSTRACT
Studies of the Hubble Deep Field (HDF) and other deep surveys have revealed
an apparent peak in the ultraviolet (UV) luminosity density, and therefore the
star-formation rate density, of the Universe at redshifts 1 < z < 2. We use photo-
metric redshifts of galaxies in the HDF to determine the comoving UV luminosity
density and find that, when errors (in particular, sampling error) are properly
accounted for, a flat distribution is statistically indistinguishable from a distri-
bution peaked at z ≃ 1.5. Furthermore, we examine the effects of cosmological
surface brightness (SB) dimming on these measurements by applying a uniform
SB cut to all galaxy fluxes after correcting them to redshift z = 5. We find that,
comparing all galaxies at the same intrinsic surface brightness sensitivity, the UV
luminosity density contributed by high intrinsic SB regions increases by almost
two orders of magnitude from z ≃ 0 to z ≃ 5. This suggests that there exists a
population of objects with very high star formation rates at high redshifts that
apparently do not exist at low redshifts. The peak of star formation, then, likely
occurs somewhere beyond z > 2.
Subject headings: galaxies: evolution -- galaxies:
universe
formation -- cosmology: early
-- 3 --
1.
Introduction
Important properties of the Universe as a whole can be determined by analyzing com-
plete samples of galaxy redshifts to very faint magnitude limits. The comoving ultraviolet
(UV) luminosity density is one such property. Because it is dominated by massive, short-
lived stars, the UV emission of an actively star forming galaxy is nearly independent of star
formation history. For this reason, the comoving UV luminosity density is directly related to
the comoving star formation and metal production rate densities of the Universe (e.g., Cowie
1988; Fall, Charlot, & Pei 1996; Madau et al. 1996, hereafter M96). Lilly et al. (1996, here-
∼ 1 using the Canada -- France
after L96) measured the UV luminosity density at redshifts z <
Redshift Survey (CFRS) and found that it rises rapidly by a factor of ∼ 15 from z = 0 to
z = 1. Similar results were found in this redshift range by Cowie, Hu, & Songaila (1997).
M96 (later updated by Madau, Pozzetti, & Dickinson 1998, hereafter M98) measured
the UV luminosity density at redshifts z ≃ 2.75 and z ≃ 4.00 using the U- and B-band
"dropout" technique applied to the Hubble Deep Field (HDF) and found that it might be
lower than the L96 measurements at z <
∼ 1 by a factor of ∼ 2. This suggested that there
might be a peak in the UV luminosity density -- and therefore the star formation and metal
production rate densities -- at some redshift between z = 1 and 2, prompting a spate of
theoretical and observational interpretation (e.g., Shaver et al. 1996; Baugh et al. 1998;
Ferguson & Babul 1998; M98; Madau, Della Valle, & Panagia 1998; Silk & Rees 1998).
More recently, Connolly et al. (1997, hereafter C97) measured the UV luminosity density
from the HDF in the previously unsurveyed redshift range 1 <
∼ 2 and found that it appears
to peak at z ≃ 1.5, which is consistent with the results of L96 and M98. However the
photometric redshifts of galaxies in the HDF from Sawicki, Lin, and Yee (1997) result in
∼ 2.5, indicating that the peak at
a UV luminosity density that continues to increase to z >
z ≃ 1.5 may be questionable.
∼ z <
In this paper, we apply our photometric redshifts of galaxies identified in the HDF
(Lanzetta, Yahil, & Fern´andez-Soto 1996, hereafter LYF96; Lanzetta, Fern´andez-Soto, &
Yahil 1998; Fern´andez-Soto, Lanzetta, & Yahil 1998, hereafter FLY98) to measure the UV
luminosity density of the Universe at redshifts 0 < z < 6. Our analysis differs from previous
analyses in three important ways. First, our photometric redshifts are determined from
spectral template fits to optimal photometry of optical (Williams et al. 1996) and infrared
(Dickinson et al. 1998) images of the HDF. In contrast to the U- and B-band "dropout"
technique of M98, our analysis determines the most likely redshift of every galaxy in the
HDF, and in contrast to the photometric redshifts of C97, our analysis makes use of the J-,
H-, and K-band infrared photometry (instead of only the J-band photometry). Second, we
determine realistic uncertainties of the luminosity density measurements, including the effects
-- 4 --
of systematic and photometric error on the photometric redshifts and of sampling error.
Previous analyses have neglected sampling error, which in fact dominates the uncertainty of
the luminosity density measurements in the HDF. Third, we explicitly consider the effects
of cosmological (1 + z)3 surface brightness dimming on the observed luminosity density.
Previous analyses have neglected cosmological (1 + z)3 surface brightness dimming, which
varies by more than two orders of magnitude over the redshift range of galaxies identified in
the HDF image.
In § 2 we briefly describe our photometric redshift technique. In § 3 we present the results
of our measurement of the UV luminosity density, in § 4 we show how this determination is
affected by cosmological surface brightness dimming, and in § 5 we discuss these results and
present our conclusions.
2. Photometric Redshifts
The starting point of our analysis is the photometric redshift estimates of LYF96 and
FLY98. Because details of the photometric redshift estimation technique have been and
will be presented elsewhere (Lanzetta, Fern´andez-Soto, & Yahil 1998; FLY98), we simply
summarize the method here.
Galaxy photometry is determined from the optical F300W, F450W, F606W, and F814W
(Williams et al. 1996) and infrared J, H, and K (Dickinson et al. 1998) images of the HDF.
To measure fluxes and flux uncertainties in the optical images, we directly integrate within
the aperture mask of every object detected in the F814W image. To measure fluxes and flux
uncertainties in the infrared images, we (1) model the spatial profile of every object detected
in the F814W image as a convolution of the portion of the F814W image containing the object
with the appropriate point spread function of the infrared image and (2) determine a least-
squares fit of a linear sum of the model spatial profiles to the infrared image. The advantages
of this method over simple aperture photometry are that the flux measurements correctly
weight signal-to-noise ratio variations within the spatial profiles, and the flux uncertainty
measurements correctly include the contributions of nearby, overlapping neighbors.
Galaxy redshifts are determined from fits to the spectral templates of E/S0, Sbc, Scd,
and Irr galaxies, including the effects of intrinsic and intervening neutral hydrogen absorp-
tion. These effects are included as a function of redshift, with mean values taken from
Madau (1995) and Webb (1997). First, we integrate the redshifted spectral templates with
the throughputs of the F300W, F450W, F606W, F814W, J, H, and K filters, at redshifts
spanning z = 0 − 7. Next, we construct the "redshift likelihood function" of every object
-- 5 --
detected in the F814W image by calculating the relative likelihood of obtaining the mea-
sured fluxes and uncertainties given the modeled fluxes at an assumed redshift, maximizing
with respect to galaxy spectral type and arbitrary flux normalization. Finally, we determine
the maximum-likelihood redshift estimate of every object detected in the F814W image by
maximizing the redshift likelihood function with respect to redshift. The result of the most
recent application of this method is presented in the catalog of FLY98, which lists photomet-
ric redshift estimates of 1067 galaxies to a limiting magnitude threshold of AB(814) = 28.0
over an angular area of ∼4 arcmin2.
Spectroscopic redshifts of more than 100 galaxies in the HDF have been obtained using
the Keck telescope (Cohen et al. 1996; Cowie 1997; Steidel et al. 1996; Zepf et al. 1997;
Lowenthal et al. 1997). Comparison between the spectroscopic and photometric redshifts
indicates the following results: (1) At redshifts z < 2, the residuals between the spectro-
scopic and photometric redshifts are characterized by an RMS dispersion of σ = 0.09 and
a discordant fraction (> 3σ discrepant after sigma clipping) of 0%; (2) at redshifts z > 2,
the residuals between the spectroscopic and photometric redshifts are characterized by an
RMS dispersion of σ = 0.45 and a discordant fraction of 7%; and (3) these residuals between
the spectroscopic and photometric redshifts arise from cosmic variance with respect to the
spectral templates (rather than from photometric error), so a proper assessment of the er-
rors of the photometric redshifts of faint galaxies must include the effects of systematic and
photometric error.
3. Ultraviolet Luminosity Density
We determined the luminosity (per unit wavelength interval) of each galaxy at a rest-
frame wavelength ∼1500A by applying an empirical K-correction derived from the best-fit
spectral template to the measured galaxy photometry. The K-corrections are interpolated
from the measured photometry at redshifts z > 0.6, although they are extrapolated from the
measured photometry (by up to a factor of two in wavelength) at redshifts z < 0.6. Next,
we determined the comoving luminosity density versus redshift by arranging the galaxies
into redshift bins, summing the luminosities within the bins, and dividing by the appropri-
ate comoving volumes. Finally, we determined the uncertainty of the comoving luminosity
density versus redshift by applying a bootstrap resampling technique. For each iteration of
the bootstrap technique, we resampled the photometric catalog and redetermined the pho-
tometric redshift of each resampled galaxy, perturbing the photometry by random deviates
according to the measured photometric error and perturbing the redshift by a random de-
viate according to the RMS residuals described in §2. We then determined the comoving
-- 6 --
luminosity density versus redshift using the resampled, perturbed redshift catalog. We re-
peated the procedure 1000 times in order to determine the range of values compatible with
the observations. This procedure explicitly allows for sampling error, photometric error, and
cosmic variance with respect to the spectral templates.
The results are shown in the top panel of Figure 1, which plots the comoving UV
luminosity density versus redshift, and are given in the second column of Table 1. For
comparison, the data points from L96, C97, and M98 are also plotted in Figure 1a. Note
that our data are entirely consistent with the previous measurements to well within ≃ 1.5σ
(and all but our first data point are within <
∼ 1σ of previous data). The fact that we do not
appear to reproduce the steep rise in the UV luminosity density from z = 0 − 1 of L96 is
attributed only to the lowest redshift data point of L96, which is <
∼ 1.5σ discrepant. While
this is still consistent with our measurement, the small difference is likely due to the much
larger sample of galaxies and the much larger surface area covered by the L96 sample as
compared to that of the relatively small HDF field.
Table 1: Comoving UV Luminosity Density
Redshift
0.00 -- 0.50
0.50 -- 1.00
1.00 -- 1.50
1.50 -- 2.00
2.00 -- 3.00
3.00 -- 4.00
4.00 -- 5.00
5.00 -- 6.00
Luminosity Densitya Luminosity Density HDF Galaxies HDF Galaxies
above SB Cut
above SB Cut
0.12
0.22
0.14
26.29± 0.31
26.43± 0.23
26.62± 0.24
26.65± 0.24
26.44± 0.28
26.41± 0.34
26.52± 0.44
26.16± 0.59
0.37
0.12
0.27
0.21
0.38
0.13
0.25
0.26
24.07± 0.24
24.98± 0.25
25.69± 0.11
25.61± 0.10
26.24± 0.17
26.15± 0.15
26.44± 0.19
25.77± 0.28
0.10
0.17
0.14
0.20
0.28
152
241
228
205
141
62
50
8
1
7
23
49
85
60
50
8
aValues are log luminosity density at restframe 1500A in units of h−2
100 erg s−1 Hz−1 Mpc−3 (qo=0.5). Errors
include the effects of systematic and photometric error on the photometric redshifts and of sampling error.
With sampling errors properly accounted for, the errorbars generated from our bootstrap
code are more than a factor of two larger than those of C97 or M98 for the z > 2 UV
luminosity density. It can be seen that, after an increase at redshifts z ≃ 0 − 2, the UV
luminosity density remains constant to within errors to redshifts z ≃ 6. In other words, we
find no convincing evidence that the UV luminosity density decreases with redshift for z > 2.
-- 7 --
Fig. 1. -- (a), The UV luminosity density as a function of redshift as measured from galaxies
in the HDF using our photometric redshifts (filled circles). The errorbars were calculated
from a bootstrapping code which takes into account the effects of systematic and photometric
error on the photometric redshifts and the effects of sampling error. For comparison, we have
included data points from Lilly et al. (1996, open triangles), Connolly et al. (1997, open
circles), and Madau et al. (1998, open squares). Note that while our data appear to show
a possible increase in the UV luminosity density from z = 0 − 2, there is little evidence for
a decrease at higher redshifts to within the errors. (b), The UV luminosity density arising
from intrinsically high surface brightness (SB) regions, after applying a uniform SB cut so
that z < 5 galaxies are considered down to the same SB level as z >
∼ 5 galaxies. Comparing
the HDF galaxies at the same intrinsic SB sensitivity shows that the UV luminosity density
of high SB regions increases by almost two orders of magnitude from z ≃ 0 to z ≃ 5.
-- 8 --
4. Effects of Cosmological (1 + z)3 Surface Brightness Dimming
To meaningfully compare the comoving UV luminosity density of the local, low-redshift
Universe with that of the distant, high-redshift Universe, it is necessary to account for the
very large effect of cosmological (1 + z)3 surface brightness (SB) dimming. Because low-
redshift galaxies are viewed to much lower intrinsic SB thresholds than are high-redshift
galaxies, certain corrections must be applied in order to view all galaxies at a common
intrinsic SB threshold.
We applied these corrections to our catalog of 1067 galaxies. First, we corrected the flux
of each galaxy to the value that would be observed if the galaxy were placed at redshift z = 5.
Specifically, we applied monochromatic SB corrections and K-corrections on a pixel-by-pixel
basis, as in Bouwens, Broadhurst, & Silk (1998), and we also applied small corrections
to account for the different WFPC2 passbands that sample rest-frame 1500 A at different
redshifts.
Next, we applied a uniform SB cut on a pixel-by-pixel basis, excluding pixels that
failed to meet a minimum SB threshold. The threshold was determined by assuming that
objects are detected in the F814W image to within ∼ 1σ of sky, which corresponds to a SB of
1.27×10−33 erg s−1 cm−2 Hz−1 pixel−1. Next, we reversed the monochromatic SB corrections
and K-corrections on a pixel-by-pixel basis to bring each galaxy back to its actual redshift
and again calculated the comoving luminosity density versus redshift. In this way, only those
parts of the galaxies that are of high enough intrinsic SB to be detected at all redshifts up
to z = 5 -- given the actual sensitivity of the HDF F814W image -- are included into the
measurement.
Results are shown in the bottom panel of Figure 1, which plots the comoving UV
luminosity density versus redshift of high intrinsic surface brightness regions, and in the
third column of Table 1. Also listed in Table 1 are the number of galaxies within each
redshift bin and the number of galaxies in each bin that have at least one pixel above the
SB cut. It is evident that (1) the comoving UV luminosity density of high intrinsic surface
brightness regions increases by two orders of magnitude from z ≃ 0 to z ≃ 5, and (2)
star-forming objects seen at z > 2.5 are relatively rare at z < 2.5.
In other words, the
comoving UV luminosity density contributed by high intrinsic SB regions appears to increase
monotonically with redshift, at least out to z ≃ 5.
-- 9 --
5. Discussion and Conclusions
Previous determinations of the comoving UV luminosity density at z > 2 rely primarily
on the M98 Lyman break galaxy sample. The stated uncertainties of these measurements
include contributions from incompleteness at the faint end of the luminosity function (LF)
as well as from the volume normalization and color selection region. While these effects do
indeed contribute to the total uncertainty, they are by no means the dominant factors. For
any LF with power-law index α < 2, the luminosity density is dominated by the bright end
of the LF. Because the bright end of the LF is inherently poorly sampled, this sampling
error at the bright end of the LF in fact dominates the uncertainty of the luminosity density.
Indeed, at redshifts z > 2, our uncertainties -- which include the effects of sampling error --
are more than two times larger than those quoted by C97 and M98. This calls into question
the statistical significant of the "peak" in the comoving UV luminosity density at z ≃ 1.5.
Ignoring errors, one can see from Figure 1a that our z = 4 measurement of the UV
luminosity density is higher than that of M98. It is possible that this arises because the
M98 method of finding galaxies at that redshift by their definition does not find all galaxies
at z ≃ 4. Rather, their color-color polygon was designed to find objects that are almost
certainly at z ≃ 4 with little contamination from low-redshift objects, which is why the data
points of M96 were plotted as lower limits. In Figure 2, we plot the z ≃ 4 objects from our
photometric redshift catalog with the M98 polygon. One can see that even with the large
(B − V ) errors involved, the M98 technique may be missing half of the high-redshift galaxies
in the HDF.
Perhaps the single most important factor that must be included in such a study is the
enormous effect of the cosmological surface brightness dimming at very high redshifts. At
z < 0.1 this effect is less than a factor of 1.3, but at z = 5 it becomes a factor of 216.
Therefore, if we are to compare the UV luminosity density at z = 5 to that of the local
Universe, we must consider local galaxies only down to the same SB level as that sampled
at high-redshift.
At the highest redshifts, the severe cosmological SB dimming affecting the HDF observa-
tions allows only the very highest levels of UV luminosity "column density" (or equivalently,
star formation rate "column density") to be sampled. At progressively lower redshifts, we
are able to observe galaxies down to lower and lower SB thresholds. The uniform SB cut
that we applied to our galaxy sample to allow for this discrepancy corresponds to a star
formation rate column density of ≃ 0.1h2
100 M⊙ yr−1 kpc−2 at z = 5 (assuming qo = 0.5 and
using the relation of UV luminosity to star formation rate with a Salpeter IMF from M98).
As seen in Table 1 and the bottom panel of Figure 1, almost no z < 1 objects have star
formation rates above this level, while at high redshifts, there are many objects exceeding
-- 10 --
Fig. 2. -- Color-color diagram similar to that of Madau et al. (1996, 1998) containing z ≃ 4
galaxies selected from our photometric redshift catalog of HDF galaxies and plotted with
their measured photometric errors.
It can be seen that, while this technique does locate
z ≃ 4 galaxies, almost 50% of them may lie outside the selection polygon despite the large
(B − V ) errors involved.
-- 11 --
this cut. For example, only 7 out of 241 galaxies from the z = 0.5 − 1.0 redshift bin would
be visible at z = 5, and then only the brightest few image pixels of those objects would peak
above the SB cutoff.
Figure 1 provides us with another piece of evidence in favor of a UV luminosity density
that increases with redshift. In the lowest redshift bin (z = 0.0 − 0.5) of Figure 1a and b,
where the SB effect is smallest, the intrinsically high SB regions make up only 0.5% of the
total UV luminosity density. If this ratio is the same at high redshifts as it is at low redshifts
then our measurements of the UV luminosity density at high redshifts in Figure 1a may need
to be increased by up to a factor of 150. Although this ratio may be quite different at high
redshifts than it is at low redshifts, this suggests that the UV luminosity density could be a
strongly increasing function of redshift.
We have shown that when the low- and high-redshift Universes are observed on equal
footing, one gets a completely different impression of the global UV luminosity density than
previously thought. We find that objects with star formation rates comparable to those at
∼ 3 are very rare in the nearby Universe. This implies that a majority of the star formation
z >
may have occurred at very high redshifts, and therefore that a peak in the star formation
rate density of the Universe has not yet been observed and likely lies somewhere at z > 2.
SMP and KML acknowledge support from NASA grant NAGW-4422 and NSF grant
AST-9624216. AF acknowledges support from a grant from the Australian Research Council.
REFERENCES
Baugh, C. M., Cole, S., Frenk, C. S., & Lacey, C. G. 1998, ApJ, 498, 504
Bouwens, R., Broadhurst, T., & Silk, J. 1998, ApJ (submitted, preprint astro-ph/9710291)
Cohen, J. G., Cowie, L. L., Hogg, D. W., Songaila, A., Blandford, R., Hu, E. M., & Shopbell,
P. 1996, ApJ, 471, L5
Connolly, A. J., Szalay, A. S., Dickinson, M., SubbaRao, M. U., & Brunner, R. J. 1997, ApJ,
486, L11
Cowie, L. L. 1988, in The Post-Recombination Universe, eds Kaiser, N. & Lasenby, A.
(Dordrecht, Kluwer) 1
Cowie, L. L., Hu, E. M., & Songaila, A. 1997, ApJ, 481, L9
Cowie, L. L. 1997, http://www.ifa.hawaii.edu/∼cowie/tts/tts.html
Dickinson, M. et al. 1998 (in preparation)
-- 12 --
Fall, S. M., Charlot, S., & Pei, Y. 1996, ApJ, 464, L43
Ferguson, H. C., & Babul, A. 1998, MNRAS, 296, 585
Fern´andez-Soto, A., Lanzetta, K. M., & Yahil, A. 1998, ApJ (in press)
Lanzetta, K. M., Yahil, A., & Fern´andez-Soto, A. 1996, Nature, 381, 759
Lanzetta, K. M., Fern´andez-Soto, A., & Yahil, A. 1998, in "The Hubble Deep Field," pro-
ceedings of the Space Telescope Science Institute 1997 May Symposium, ed. M. Livio,
S. M. Fall, and P. Madau (in press, preprint astro-ph/9709166)
Lilly, S. J., & Le F`evre, O., Hammer, F., & Crampton, D. 1996, ApJ, 460, L1
Lowenthal, J. D. et al. 1997, ApJ, 481, 673
Madau, P. 1995, ApJ, 441, 18
Madau, P., Ferguson, H. C., Dickinson, M., Giavalisco, M., Steidel, C. C., & Fruchter, A.
1996, MNRAS, 283, 1388
Madau, P., Pozzetti, L., & Dickinson, M. 1998, ApJ, 498, 106
Madau, P., Della Valle, M., & Panagia, N. 1998, MNRAS, 297, 17
Sawicki, M. J., Lin, H., & Yee, H. K. C. 1997, AJ, 113, 1
Shaver, P. A., Wall, J. V., Kellermann, K. I., Jackson, C. A., & Hawkins, M. R. S. 1996,
Nature, 384, 439
Silk, J., & Rees, M. J. 1998, A&A, 331, L1
Steidel, C. C., Giavalisco, M., Dickinson, M., & Adelberger, K. L. 1996, AJ, 112, 352
Webb, J. K. 1997, (private communication)
Williams, R. E., et al. 1996, AJ, 112, 1335
Zepf, S. E., Moustakas, L. A., & Davis, M. 1997, ApJ, 474, L1
This preprint was prepared with the AAS LATEX macros v4.0.
|
astro-ph/0308027 | 1 | 0308 | 2003-08-02T22:39:26 | IRAS PSCz v.s. 1.2-Jy Velocity and Density Fields: A Spherical Harmonics Comparison | [
"astro-ph"
] | We have used the two IRAS redshift surveys, 1.2-Jy (Fisher et al 1995) and PSCz (Saunders et al 2000), to model the linear velocity fields within a redshift of 8,000 km/s and have compared them in redshift space. The two velocity fields only differ significantly in their monopole components. The monopole discrepancy cannot be solely ascribed to shot-noise errors and incomplete sky coverage. The mismatch seems to arise from incompleteness of the PSCz catalog at fluxes $\le$ 1.2 Jy. The 1.2-Jy and PSCz higher order velocity multipoles, particularly the dipole and quadrupole components, appear to be consistent, suggesting that the dipole residuals found by Davis, Nusser and Willick (1996) when comparing 1.2-Jy and MarkIII velocity fields probably originates from the MarkIII velocities calibration procedure rather than from uncertainties in the model velocity field. Our results illustrate the efficiency of the spherical harmonics decomposition techniques in detecting possible differences between real and model velocity fields. Similar analyses shall prove to be very useful in the future to test the reliability of next generation model velocity fields derived from new redshift catalogs like 2dFGRS (Colless et al 2001), SDSS (York et al 2000), 6dF and 2MRS. | astro-ph | astro-ph |
DRAFT VERSION NOVEMBER 12, 2018
Preprint typeset using LATEX style emulateapj v. 26/01/00
IRAS PSCZ V.S. IRAS 1.2-JY MODEL VELOCITY FIELDS: A SPHERICAL HARMONICS COMPARISON.
LUÍS TEODORO1, ENZO BRANCHINI2 AND CARLOS S. FRENK3
1T-8, Theoretical Division, Los Alamos National Laboratory, Los Alamos, New Mexico, 87545, USA
2Dipartimento di Fisica, Università degli Studi Roma TRE, Roma, Italia
3Department of Physics, University of Durham, Science Laboratories, Durham DH1 3LE, England.
Draft version November 12, 2018
ABSTRACT
We have used the two IRAS redshift surveys, 1.2-Jy (Fisher et al. 1995) and PSCz (Saunders et al. 2000), to
model the linear velocity fields within a redshift of 8 000 km s- 1 and have compared them in redshift space. The
two velocity fields only differ significantly in their monopole components. The monopole discrepancy cannot be
solely ascribed to shot-noise errors and incomplete sky coverage. The mismatch seems to arise from incomplete-
ness of the PSCz catalog at fluxes ≤ 1.2 Jy. The 1.2-Jy and PSCz higher order velocity multipoles, particularly
the dipole and quadrupole components, appear to be consistent, suggesting that the dipole residuals found by
Davis, Nusser and Willick (1996) when comparing 1.2-Jy and Mark III velocity fields probably originates from
the Mark III velocities calibration procedure rather than from uncertainties in the model velocity field. Our results
illustrate the efficiency of the spherical harmonics decomposition techniques in detecting possible differences
between real and model velocity fields. Similar analyses shall prove to be very useful in the future to test the
reliability of next generation model velocity fields derived from new redshift catalogs like 2dFGRS (Colless et
al. 2001), SDSS (York et al. 2000), 6dF and 2MRS.
Subject headings: cosmology -- theory -- galaxies large-scale structure of the Universe -- large-scale dynamics.
1. INTRODUCTION
In the framework of linear gravitational instability theory and
linear biasing a comparison between the predicted and observed
peculiar velocities allows us to measure the so called β param-
eter
β ≡
.
(1)
Ω0.6
m
b
i.e. a combination of the mass density parameter, Ωm, and the
linear bias parameter, b, which relates the mass overdensity
field δ to the density fluctuations in galaxy counts δg, through
δg = bδ. Several different techniques have been used to compare
measured peculiar velocities (taken from all available catalogs:
Mark III Willick 1997a, SFI Haynes 1999, ENEAR da Costa et
al. 2000, SEcat Zaroubi 2002) to predict velocities that, in most
cases, have been obtained from two redshift surveys of IRAS
galaxies: the 1.2-Jy and the PSCz catalogs.
Most analyses have investigated the consistency between
model and observed velocities, while only a few of them have
been devoted in comparing different datasets or in studying the
consistency of model velocity fields. All recent works show
a general consistency between models and data (Willick et al.
1997, hereafter WSDK; Willick and Strauss 1998; Sigad et
al. 1998; Dekel et al. 1999; Branchini et al. 1999, hereafter
B99; Nusser et al. 2000; Zaroubi et al. 2002) with one no-
ticeable exception represented by the work of Davis, Nusser
& Willick (1996, hereafter DNW). Their analysis showed that
the velocity residuals between the Mark III peculiar veloci-
ties and the 1.2-Jy predictions display a significant spatial cor-
relation. Subsequent analyses by WSDK and Willick and
1
Strauss (1998) were performed using the same dataset, a sim-
ilar model velocity field but a different comparison technique,
called VELMOD, which allows for independent calibrations of
Mark III velocities in each of the Mark III sub-catalogs. As
a result the observed and model velocity fields were brought
into better agreement, at least within a redshift distance s =
7 500 km s- 1 and returned a calibration inconsistent with the
original one performed by Willick et al. (1997a).
While these results suggest that the DNW mismatch orig-
inates from calibration problems in the Mark III catalog, the
possibility that systematic errors in the 1.2-Jy model velocities
also contribute to the discrepancy has been completely over-
looked. In particular, we ignored whether the velocity fields
predicted from the deeper IRAS PSCz catalog provides a better
match to the Mark III velocities than the shallower 1.2-Jy cata-
log, considered by DNW.
The main goal of this work is to check whether the discrep-
ancies between the 1.2-Jy and Mark III velocities can be as-
cribed, at least to some extent, to model uncertainties. For
this purpose we compare both the overdensity and the radial
velocity fields modeled from the redshift space distribution of
1.2-Jy and PSCz galaxies. This work is meant to stress the
usefulness of spherical harmonics decomposition techniques in
assessing the adequacy of a model velocity (or density) field
through velocity-velocity (or density-density) comparisons.
In § 2 we describe how to compare velocity and overden-
sity fields inferred from different redshift surveys. The galaxy
redshift catalogs used in our analyses are described in § 3. In
sections 4 and 5 we analyse the two model overdensity fields
and radial velocity fields. Our main results are discussed in § 6
2
PSCz v.s. 1.2-Jy Velocity and Density Fields
and our main conclusions are presented in § 7.
2. SPHERICAL HARMONIC COEFFICIENTS DECOMPOSITION
The relations between an arbitrary scalar field ψ(s) and the
real -- valued spherical harmonic coefficients ψlm(s) are
ψlm(s) =Z ψ(s)Ylm(s)dΩ
∞
l
ψ(s) =
ψlm(s)Ylm(s),
Xl=0
Xm=- l
(2)
(3)
where Ylm(s){l = 0, .., ∞; m ≤ l} denote the well known real-
valued spherical harmonics, defined as in Baker et al. (1999,
hereafter BDSLS) and Bunn (1995) (see also Jackson 1999).
For a given l these functions are normalized to the value of Yl0
at the North Galactic Pole.
in quadrature over the (2l + 1)-ψ lm's, where {m = - l, ...,l}:
The amplitude of the l-multipole ψl(s) is defined by the sum
ψl(s) =" m=l
Xm=- l
1
2
.
ψ2
lm(s)#
2.1. Peculiar velocity and overdensity fields from the
distribution of galaxies in redshift space
(4)
FIG. 1. -- The 1.2-Jy smoothing width as a function of redshift. The dots
represent the value of σn at centre of the 52 redshift bins.
Nusser & Davis (1994) show that in the linear regime the
peculiar velocity field is irrotational not only in real space but
also in redshift space and thus can be expressed as the gradient
of a scalar velocity potential: v(s) = - ∇Φ(s). At a given redshift
s, Φ(s) and δ(s) can be expanded in spherical harmonics and
related to each other through a modified Poisson equation:
1
lm(cid:1)′ -
s2(cid:0)s2Φ′
1
1 + β
l(l + 1)Φlm
s2
=
β
1 + β(cid:18)δlm - 1
s
d lnφ
d lns
Φlm
′(cid:19) ,
(5)
where φ(s) and δlm(s) denote the sample selection function
and the spherical components of the overdensity field, respec-
tively. Here prime expresses d/ds. Solving this differential
equation requires computing the redshift space density field
from the observed galaxy distribution on a spherical grid us-
ing Gaussian cells of approximately equal solid angle. These
cells are equally distributed in longitude (64 bins) and are cen-
tered at the Gaussian-Legendre 32 point quadrature formula in
the range - 1 ≤ cos(b) ≤ 1, where b is the Galactic latitude.
In this work we measure redshifts in the Local Group frame,
and use 52 Gaussian radial redshift bins out to a distance of
s = 18 000 km s- 1. When comparing two or more galaxy cata-
logs the size of the radial bin is set by requiring equal spatial
resolution and the minimal shot noise constant throughout the
volume. We do this setting the distance between the centers
of the Gaussian cells at distance s equal to the average galaxy-
galaxy separation, σ(s) = [¯nφ(s)]- 1/3, in the sparser catalog. The
mean number density is estimated as
1
1
(6)
¯n =
V Xi
,
φ(si)
where the sum is over the galaxies contained within the volume
V , which, in this work, is a sphere of radius 8 000 km s- 1.
1 + δ(sn) =
Hernquist and Katz (1989) pointed out that there are two pos-
sible ways to estimate the smoothed density field: the "scatter"
and "gather" approches. In the former approch each particle is
distributed in space, and the density estimate at a given point is
the superposition of individual smoothing spheres. In the lat-
ter case one defines a smoothing lenght at a given point and
weights the particle in its neighborhood by the resulting kernel.
The Gaussian-smoothed galaxy overdensity field at a grid cell
centred around sn is given by:
N
1
(2π)3/2σ3
n
1
φ(si)
exp(cid:20)-
Xi
(sn - si)2
2σ2
n
(cid:21) ,
(7)
where the sum is over the N galaxies in the catalog, which
clearly is a "gather" approach. The smoothing width at a red-
shift s, used in this paper is σn = max[320, σ(s)] km s- 1. Since
in this work we compare the PSCz catalog with the sparser
1.2-Jy catalog, we use the σn obtained from the 1.2-Jy catalog
shown in Fig. (1). This smoothing scheme is tailored to keep the
shot-noise uncertainty roughly constant throughout the sample
volume and is similar to the optimal Wiener filtering procedure
used by Lahav et al. (1994) and Fisher et al. (1995b). More-
over, the our kernel has a compact support, which simplifies
the computational task.
Eqn. (5) is valid in the linear regime, where a one-to-one
mapping between distance and redshift is guaranteed. This as-
sumption is not valid in high density regions, such as cluster of
galaxies, where shell crossing may occur triple valued-regions
may apear. To deal with the latter effect, we adopt the same
procedure as Yahil et al. (1991) and collapse the fingers of God
associated with the six richest clusters in the sample listed in
Table 2 of that work.
In this work we will consider galaxies within smax = 20 000
Teodoro, Branchini & Frenk
3
km s- 1 with the spherical harmonics decomposition limited to
lmax = 15. Outlying smax galaxies are assumed to be distributed
uniformly according to the mean number density of the parent
catalogs.
3. DATASETS
The IRAS PSCz catalog contains ≈ 15 500 IRAS PSC galax-
ies with a 60 µm flux larger than 0.6 Jy. The average depth of
this survey is ≈ 100 h- 1 Mpc . We restrict our analysis to a sub-
sample of 11 206 galaxies within 20 000 km s- 1 from the Local
Group (LG, hereafter). The areas not covered by the survey
amount to 16.0% of the sky and are preferentially located near
the galactic plane, in the so-called Zone of Avoidance (Saun-
ders et al. 2000). The empty areas have been filled-in with the
cloning procedure described in Branchini et al. (1999) .
The 1.2-Jy catalog (Fisher et al. 1995a) contains 5 331 IRAS
galaxies with a 60 µm flux larger than 1.2-Jy within 20 000
km s- 1. This catalog has a slightly larger sky coverage of (≈
- 1 Mpc). The
87.6%) and a smaller median redshift (≈ 84 h
same B99 filling technique has been used to restore all sky cov-
erage.
Fisher et al. (1995b) have found that uncertainties in filling-
in the empty areas of the 1.2-Jy survey do not appreciably affect
the spherical harmonics coefficients with l . 15. Similar con-
clusions are also valid for the IRAS PSCz survey (Teodoro et
al. 1999).
The selection functions of the two surveys, that quantify the
probability of a galaxy at distance s, have been determined as
maximum-likelihood fits to the smooth function
φ(s) =( (cid:0) rr
r(cid:1)2α(cid:16) r2
1
⋆
r2
⋆
+r2
s
+r2(cid:17)β
, s ≥ H0 rs,
, otherwise,
(8)
1 is expressed in h- 1 Mpc (Yahil et al. 1991).
where r = s/H0
The best-fit parameters, determined using the galaxies within
8 000 km s- 1 and assuming a flat universe with Ωm = 1, are
shown in Tab. (1). The quantity rs in Eqn. (8) accounts for
the incompleteness of faint galaxies in the innermost regions of
the catalogs (Rowan-Robinson et al. 1990, Yahil et al. 1991).
4. THE IRAS PSCz AND 1.2-Jy FIELDS OVERDENSITY
The first step of our analysis consists of applying Eqn. (7) to
compute the IRAS 1.2-Jy and PSCz overdensities onto a spheri-
cal grid of radius 18 000 km s- 1 and lmax = 15 with the smooth-
ing scheme given by Eqn. (7).
4.1. Overdensity Maps
In Figs. (2 -- 4) we show the Aitoff projections of the PSCz (top
panels) and 1.2-Jy (middle panels) galaxy overdensity fields in
three spherical shells centred at s = 1 000, 3 000 and 5 000 km s- 1,
respectively. The thickness of each shell is defined by the Gaus-
sian width at the distance of the shell. All maps are plotted in
Galactic coordinates. The bottom panels show the map of the
this article we write Hubble's constant as H0 = 100h
1Throughout
km s- 1 Mpc- 1
FIG. 2. -- PSCz (top), 1.2-Jy (middle) and residual (bottom) overdensity
fields in the radial shell centred at s = 1 000 km s- 1. The thick-line contour
indicates δ = 0 or δres = 0 (bottom). The light- and dark-gray contours show
over- and under-dense regions, respectively. In the top and middle panels the
contours are equally -- spaced by ∆δ = 1.0. The bottom panel the density resid-
uals are shown in steps of ∆δres = 0.125.
overdensity residuals δres(s) = δ0.6(s) - δ1.2(s), where the sub-
scripts 0.6 and 1.2 refer to PSCz and 1.2-Jy surveys, respec-
tively. In all plots the thick line indicates the zero overdensity
(or residual) contour. The overdensity (residual) contour spac-
ing ∆δ (∆δres) depends on the shell's depth and is indicated in
the Figure captions.
Detailed cosmographic studies of the
structures in the nearby universe have already been performed
using both the distribution of 1.2-Jy (Strauss and Willick 1995)
and PSCz galaxies2 and will not be repeated here. We only
stress that the main cosmic structures like the Virgo Cluster (l ≈
240◦, b ≈ 75◦, s = 1000 km s- 1), the local void spanning the re-
gion 330◦ . l . 120◦, - 90◦ . b . 30◦ at s = 1000 km s- 1, the
Hydra-Centaurus complex (l ≈ 270◦, b ≈ 20◦, s = 3000 km s- 1)
and (l ≈ 300◦, b ≈ 20◦, s = 3000 km s- 1) and the Perseus-
Pisces supercluster (l ≈ 140◦, b ≈ - 25◦, s = 3000 km s- 1) are
detected in both surveys.
The residuals, calculated at the same gridpoints as the den-
2See http://www-astro.physics.ox.ac.uk/∼wjs/pscz.html
4
PSCz v.s. 1.2-Jy Velocity and Density Fields
SELECTION FUNCTION PARAMETERS FOR IRAS CATALOGS.
TABLE 1
60µm Ngal a
Catalog f lim
[Jy]
α
β
γ
r b
s
r⋆
¯n
[ h- 1 Mpc] [ h- 1 Mpc] [h3Mpc- 3]
0.6 13 364 0.54 1.80 ...
PSCz
PSCz 0.7 0.7 11 170 0.57 1.80 ...
PSCz 1.2 1.2
5 519 0.43 1.86 ...
1.2
1.2-Jy
6 010 0.43 1.86 ...
c
0.6 13 364 0.99 3.45 1.93
PSCz evo.
6.0
6.0
5.0
5.0
6.0
87.00
76.35
50.40
50.60
76.02
6.12 ·10- 2
6.38 ·10- 2
4.62 ·10- 2
4.67 ·10- 2
5.76 ·10- 2
a Includes synthetic objects.
b Parameter kept constant in the likelihood maximization.
c The selection function parameters have been taken from Canavezes et al. (1999)
FIG. 3. -- Same as in Fig. (2). The shell is now centred at s = 3 000
km s- 1 and the contour spacings are ∆δ = 0.5 and ∆δres = 0.125.
FIG. 4. -- Same as in Fig. (2). The shell is now centred at s = 5 000
km s- 1 and the contour spacings are ∆δ = 0.5 and ∆δres = 0.125.
Teodoro, Branchini & Frenk
5
THE PARAMETERS OF THE δ1.2 - δ0.6 AND u1.2 - u0.6 LINEAR REGRESSIONS.
TABLE 2
Regression Ngrid
Ndof
A a
B
χ2
eff/Ndof
δ1.2 - δ0.6
u1.2 - u0.6
112 552 156.36
112 552 . 156.36
- 0.011 ± 0.014
60.0 ± 5.9
1.025 ± 0.036
1.040 ± 0.019
0.78
0.88
Notes:
a For the u - u comparison A is expressed in km s- 1.
Column 2: Ngrid, the number of gridpoints used for the regression;
column 3: Ndof, the number of independent volumes;
column 4: A, the zero -- point offset in the linear regression, and its 1 - σ error;
column 5: B, the slope of the linear regression, and its 1 - σ error;
column 6: χ2
eff/Ndof from the linear regression.
i. This expression is a generalization of Eqn. (12) in Dekel et
al. (1993) to the case of a grid with variable mesh-size since
Rs,i depends on the gridpoint position. The new statistics χ2
eff =
(Ngrid/Neff)χ2 is approximately distributed as a χ2 with Ndof =
Neff - 1 degrees of freedom and can be applied to infer the values
of A, B and their errors.
Fig. (5) shows the δ1.2 vs. δ0.6 scatter diagram obtained us-
ing all 112 552 gridpoints with s < 8 000 km s- 1 and b > 8o.
Within these limits the estimated overdensities are little affected
by uncertainties in the filling-in procedure, and we expect that
sity fields, seem to indicate that, as expected, the largest residu-
als are located near the galactic plane where uncertainties in the
filling-in procedure dominate the errors. In the inner shell, over-
densities in the 1.2-Jy galaxy distribution appear to be larger
than the PSCz ones while in the most external shell the situa-
tion is reversed.
4.2. Density Residuals
A way of quantifying possible systematic differences be-
tween the two fields is that of performing a point-to-point com-
parison.
In absence of systematic errors we expect the two
IRAS overdensity fields to be linearly related through
δ1.2 = Bδ0.6 + A
(9)
with A = 0 and B = 1. δ0.6 and δ1.2 are estimated at Cartesian
gridpoint positions. A zero -- point offset A 6= 0 indicates discrep-
ancies in the mean densities of the two catalogs while a slope
B 6= 1 reveals the presence of systematic errors. If the values of
δ0.6 and δ1.2 are independent and their errors are Gaussian then
we can compute the χ2 statistics
χ2 =
Ngrid
Xi=1
(δ1.2 - A - Bδ0.6)2
+ B2σ2
(cid:0)σ2
0.6(cid:1)
1.2
,
(10)
where σ0.6 and σ1.2 indicate the typical errors in δ and Ngrid is
the total number of points in the comparison. However, since
the gridspacing (250 km s- 1) is smaller than the smoothing length
(≥ 320 km s- 1), not all values of δ0.6 (or δ1.2) are independent.
Indeed, the effective number of independent points used in the
χ2 statistics, Neff is given by
N- 1
eff = Ngrid
- 2
exp(- r2
i j/2R2
s,i),
(11)
Ngrid
Ngrid
Xi=1
Xj=1
where ri j is the separation between gridpoints i and j and Rs,i
is the smoothing radius of the Gaussian kernel at the gridpoint
FIG. 5. -- 1.2-Jy vs. PSCz overdensity. Only 1883 randomly selected grid-
points with b > 8◦ and within s = 8 000 km s- 1 are shown in the scatter plot.
The continuous line represents the linear best fit characterized by the zeropoint
A and slope B indicated in the panel.
6
PSCz v.s. 1.2-Jy Velocity and Density Fields
consistent with recent determinations based on several velocity-
velocity comparisons (e.g. Nusser et al. 2000), a recent density-
density analysis (Zaroubi et al. 2002) and the study of mean
relative velocities of galaxy pairs (Juszkiewicz et al. 2000).
5.1. Radial Peculiar Velocity Maps
To obtain maps of the radial velocity field we have computed
the gradient of the velocity potential projected along the line of
sight: u(s) = s · ∇Φ at the same gridpoint positions used for the
analysis of the density field.
The top end mid panels of Figs. (7-9) are similar to those of
figs. 2 -- 4 and show the radial velocity maps in the same shells.
The bottom panels show the maps of the radial velocity residu-
als ures(s) ≡ u0.6(s) - u1.2(s). The thick line represents the zero
radial velocity (residual) contour. The velocity (residual) con-
tour spacing ∆u (∆ures) varies with the redshift and is indicated
in the Figure captions.
The main feature seen at all redshifts in both 1.2-Jy and
PSCz radial velocity maps is the dipole pattern resulting from
the LG reflex motion toward the CMB apex (l ≈ 276◦,b ≈ 30◦,
Kogut et al.1993).
Local deviations from this pattern arise from peculiar mo-
tions, like the infall onto Perseus-Pisces superclusters at (s ≈
3000 km s- 1, l ≈ 120◦, b ≈ - 40◦). The largest velocity resid-
uals are located near the galactic plane, but their contours are
broader due to the intrinsic non-locality of the peculiar velocity
field. The most striking feature is perhaps the fact that velocity
residuals, that are positive across a large fraction of the sky in
the first shell, becomes mostly negative in the remaining two
shells, suggesting possible systematic errors in either PSCz or
1.2-Jy velocity models.
shot-noise dominates the total error budget. To estimate the
shot-noise errors affecting both PSCz and 1.2-Jy overdensity
fields we have generated 100 bootstrap realizations of the ob-
served PSCz and 1.2-Jy surveys. These are obtained by replac-
ing each galaxy (including the 'synthetic' galaxies used to fill-
in the unobserved regions) with a number of objects drawn from
a Poisson deviate with mean unity. Shot noise errors in δ at the
generic gridpoint were set equal to the dispersion over the 100
realizations. The solid line in the plot is the best linear fit ob-
tained from the minimization of χ2
eff. As shown in Tab. (2), the
slope of the line, B = 1.025 ± 0.036, is consistent with unity,
indicating no systematic mismatch. On the other hand the zero-
point value A = - 0.011 ± 0.014 shows no systematic difference
in the mean density of the two IRAS catalogs.
The scatter around the best -- fitting line, σδ = 0.16 is similar to
average shot-noise error in the 1.2-Jy density field (σδ1.2
SN = 0.15).
The parameter S = χeff/Ndof = 0.78 indicates that bootstrap er-
rors slightly overestimate the true uncertainties.
4.3. Radial Density Contrast
An alternative way to characterize possible discrepancies be-
tween the two IRAS density fields is that of computing the dif-
ferential radial density contrast in shells with identical thick-
ness ∆s for both galaxy distributions:
ρ(s)
¯ρ
=
1
V ¯n
W (s,si, ∆s)
φ(si)
,
(12)
Np
Xi=1
where the window associated to the spherical shell Wδρ/ ¯ρ(s,si,
∆s) is the Heaviside function H(s + ∆s/2 - si)H(si + ∆s/2 - s),
si is the redshift of the object i, V is the volume used to com-
pute the mean number density ¯n and the sum runs over the Np
galaxies of the catalog. Fig. (6) shows the quantity ρ(s)/ ¯ρ for
1.2-Jy (light, dotted histogram) and PSCz (thick, continuous
histogram) galaxies within s = 15 000 km s- 1, computed in in
spherical shells of thickness ∆s = 200 km s- 1. Although the
general shapes of two radial density contrasts are similar (with
a bump at s ∼ 1 100 km s- 1 due to Virgo and Fornax clusters,
a dip at s ∼ 3 500 km s- 1 induced by the Sculptor Void and a
second prominent peak at s ∼ 5 000 km s- 1, at the same red-
shift as the Perseus-Pisces supercluster and the Great Attrac-
tor), a discrepancy between the two curves exists in the range
2 000 . s . 6 000 km s- 1 where the PSCz density is larger than
the 1.2-Jy one.
The two smoothed curves in Fig. (6) show the monopole
terms of the two overdensity fields δl=0,m=0 as a function of
radius, obtained from the spherical harmonics decomposition.
Both curves interpolate the histograms rather well (hence show-
ing the adequacy of our spherical harmonics analysis) and con-
firm the existence of a PSCz vs. 1.2-Jy density mismatch be-
tween the two catalogs.
5. THE IRAS PSCz AND 1.2-Jy PECULIAR VELOCITY FIELDS
To perform a similar analysis on the 1.2-Jy and PSCz model
velocity fields we have integrated Eqn. (5) using β = 0.6, a value
FIG. 6. -- Differential radial density contrast for PSCz (thick line histogram)
and 1.2-Jy galaxies (thin line histogram). The monopole of the PSCz and 1.2-
Jy density fields are shown with smooth solid and dashed lines, respectively.
Teodoro, Branchini & Frenk
7
FIG. 7. -- PSCz (top), 1.2-Jy (middle) and residual (bottom) radial velocity
fields in the spherical shell centred at s = 1 000 km s- 1. The thick-line contour
indicates δu = 0 or ∆ures = 0. The light- and dark-gray contours show regions
characterized by outflows or infall motions. In the top and middle panels the
contours are equally -- spaced by ∆u = 200 km s- 1. The bottom panel shows
the velocity residuals in steps of ∆δres = 25 km s- 1.
5.2. Velocity Residuals
To perform a more quantitative comparison between the two
radial velocity fields we have repeated the analysis of § 4.2 and
performed a linear regression between u1.2 and u0.6 by minimiz-
ing
Ngrid
Xi=1 (cid:0)u1.2 - A - Bσu,0.6(cid:1)2
u,0.6(cid:17)
(cid:16)u2
+ B2σ2
1.2
χ2 =
,
(13)
where Ngrid runs over the same points used for the density-
density comparison. Linear velocities are obtained by integrat-
ing the mass density fields over large scales causing them to
be correlated on scales larger than the radius of the Gaussian
kernel. This means that the number of independent gridpoint
Ne f f is smaller than the estimate given by Eqn. (11). Intrinsic
large scale smoothing also affects the velocity error estimate,
σu since uncertainties in modeling the galaxy distribution in the
unobserved areas now add to shot noise errors.
FIG. 8. -- Same as in Fig. (7). The shell is now centred at s = 3 000 km s- 1and
the contour spacings are ∆u = 200 km s- 1and ∆ures = 50 km s- 1
The shot noise errors have been computed using the same
bootstrap resampling technique described in § 4.2. Uncertain-
ties in the filling-in procedure have been evaluated from 20 in-
dependent PSCz and 1.2-Jy mock catalogs extracted from the
N-body simulations performed by Cole et al. (1997) simulating
a volume of ≈ 170 h- 1 Mpc in an Einstein-de Sitter universe
with a non-vanishing (ΩΛ = 0.7) cosmological constant and
density fluctuations normalized to the observed cluster abun-
dance (Eke, Cole & Frenk 1996 ). Model velocity fields from
each IRAS mock catalog were compared with velocities ob-
tained from ideal, all-sky IRAS mock catalogs. The variance
over 20 mocks at each gridpoint quantifies the uncertainties
in the filling procedure. Total uncertainties were obtained by
adding in quadrature the two errors.
The results of the linear regression are summarized in Tab. (2)
which shows the slope of the best fitting line B = 1.040 ± 0.019
and the value of the zero-point, A = 60.0 ± 5.9 km s- 1. While
the slope is still consistent with unity, at the 2.1-σ confidence
level, the zero point is significantly different from zero, hence
corroborating the evidence for a mismatch in the average densi-
ties of PSCz and 1.2-Jy catalogs. The dispersion around the fit
8
PSCz v.s. 1.2-Jy Velocity and Density Fields
FIG. 10. -- 1.2-Jy vs. PSCz radial velocities computed at 1883 randomly
selected gridpoint positions with b > 8◦ and s ≤ 8 000 km s- 1 inferred radial
velocity field (x-axis) versus that inferred from the 1.2-Jy survey (y-axis). The
parameters of the best fitting line (A, B) the scatter σu are indicated.
The monopole terms of PSCz (continuous line) and 1.2-
Jy (dashed line) velocity fields are plotted in Fig. (11) as a func-
tion of the redshift. The hatched and dashed regions represent
the 1-σ uncertainty strip, computed from the mock catalogs and
the bootstrap resampling procedure.
FIG. 11. -- Monopole coefficients of 1.2-Jy (dashed) and PSCz (continuous)
radial velocity fields. 1 - σ errors are represented by the hatched (shaded)
areas.
FIG. 9. -- Same as in Fig. (7). The shell is now centred at s = 5 000 km s- 1and
the contour spacings are ∆u = 200 km s- 1and ∆ures = 50 km s- 1.
is, σu = 88 km s- 1 is similar to the typical velocity error com-
puted from the IRAS mocks and the bootstrap resampling cat-
alogs (σu0.6 = 71 km s- 1 and σu1.2 = 140 km s- 1). The parameter
χeff/Ndof is less than unity which either indicates that errors
are overestimated or that the value Ne f f , taken from Eqn. (11),
does significantly overestimate the true number of degrees of
freedom.
5.3. Velocity Multipoles
The spherical harmonics decomposition allows us to bet-
ter investigate and characterize possible discrepancies between
1.2-Jy and PSCz velocity fields. Here we are only interested
in the first three (l = 0,1,2) components, corresponding to the
monopole, dipole and quadrupole terms, which have already
been investigated in previous analyses (e.g. BDSLS). It is worth
stressing that the monopole and dipole components at a given
redshift are only sensitive to the mass distribution within the
corresponding distance and thus can be directly related to the
structures within the volume considered, while the quadrupole
components are also sensitive to the mass distribution beyond
such a distance [see Eqn. (9) in Nusser & Davis 1994].
Teodoro, Branchini & Frenk
9
FIG. 12. -- The velocity dipole coefficients, u1m(s){m = - 1, 0, 1}, inferred
from PSCz (continuous) and 1.2-Jy (dashed) and their associated 1-σ uncer-
tainties. The three Galactic Cartesian components GX, GY , GZ are shown in
the top-left, top-right and bottom-right panel, respectively. The total ampli-
tude is shown in the bottom-left panel. The dotted line refers to the dipole of
PSCz galaxies with flux f60µm ≥ 1.2 Jy.
Both velocity monopoles display the same general features,
as expected given the similarities between the density monopoles
δ00(s) shown in Fig. (6). There is a significant radial in-
flow around s = 1 800 km s- 1 corresponding to the infall mo-
tion toward the Virgo-Fornax complex, an outflow in the range
1 800 . s . 4 000 km s- 1 due to the presence of the local void
and another infall beyond this radius, due to the Hercules, Hy-
dra, Centaurus and Perseus-Pisces superclusters. Differences
between the two velocity monopoles are detected with a signif-
icance level larger than 1-σ at very small radii (s < 1200 km s- 1)
and in the range 3500 km s- 1< s < 7500 km s- 1.
Fig. (12) shows the total amplitude (lower left panel) and
the three Cartesian components (remaining panels) of the two
IRAS velocity dipoles. The total amplitude of the PSCz dipole
(continuous line) is systematically smaller that the 1.2-Jy one
(dashed line) but the discrepancy is well within the expected
errors, in agreement with previous studies (e.g. Schmoldt et
al. 1999).
Finally, Fig. (13) shows the five components of the PSCz (con-
tinuous line) and 1.2-Jy (dashed line) velocity quadrupoles,
u 2m(s) (where m = - 2, ...,2) along with total amplitude (bottom-
left panel). As for the dipole case, the two sets of quadrupole
components agree to within 1-σ apart from a small disagree-
ment in the the m = +2 and m = - 2 components beyond s = 7000
km s- 1. Indeed, beyond s ≈ 6 000 km s- 1 the amplitude of the
PSCz quadrupole decreases, while the 1.2-Jy one increases
monotonically.
It is worth noticing that our 1.2-Jy velocity multipoles agree
FIG. 13. -- Quadrupole coefficients u 2m(s){m = - 2, ..., 2} derived from
PSCz (solid line) and 1.2-Jy (dashed) model velocity fields. Shaded regions in-
dicate the 1 - σ uncertainty regions. The dotted line represents the quadrupole
components for PSCz galaxies with flux f60µm ≥ 1.2 Jy.
with those computed by BDSLS except that their multipole
components exhibit sharper features at small radii (s < 500
km s- 1). This mismatch derives from our use of a Gaussian
smoothing kernel of radius 330 km s- 1 which is somewhat larger
than that applied by BDSLS.
6. DISCUSSION.
Perhaps the most surprising result of our analysis is the dis-
agreement between the PSCz and 1.2-Jy velocity monopoles in
the 1 800 . s . 4 000 km s- 1 range. Before investigating the
possible causes of this mismatch, it is worth noticing that none
of our IRAS monopoles is consistent with an outward flow of
magnitude ≈ 400 km s- 1 at a distance of 7 000 km s- 1 like the
one detected by Zehavi et al. (1998) using a sample of 44 Type
Ia supernovae. That result is also at variance with the velocity
model obtained from the distribution of ORS galaxies obtained
by BDSLS which means that, if confirmed, it would be difficult
to reconcile with the Gravitational Instability scenario.
A possible clue to understand the origin of the monopole
mismatch is provided by the analysis of BDSLS that returns
a velocity monopole somewhere between the PSCz and 1.2-
10
PSCz v.s. 1.2-Jy Velocity and Density Fields
Jy ones in the range of interest. This seems to suggest that
our monopole discrepancy originates from systematic errors in
one of the IRAS catalogs, possibly some incompleteness of the
largest catalog at faint fluxes.
The results of Tadros et al. (1999) seem to confirm that
PSCz catalog might be statistically incomplete at fluxes f60µm ≤
0.7 Jy. To test whether this is indeed the case we have extracted
a brighter PSCz subsample (called PSCz0.7) by discarding all
object with f60µm ≤ 0.7 Jy and have repeated our spherical har-
monics decomposition analysis. The results, shown in Fig. (14)
indicate that the PSCz0.7 velocity monopole (dotted line) basi-
cally coincide with that of the complete sample. It is only when
discarding galaxies with fluxes smaller than 1.2 Jy that the ve-
locity monopole (dashed line) agrees with the 1.2-Jy one.
Another possibility of understanding the monopole mismatch
is that of invoking a strong evolution of IRAS galaxies as a
function of the redshift, which we have neglected in our calcu-
lations so far. Roughly speaking evolution can be distinguish
into two different types. Pure density evolution changes only
the normalization of the luminosity function while its shape is
invariant with redshift. Pure luminosity evolution on the other
hand describes a possible change of the intrinsic luminosities of
the sources but not their number density. Both forms of evolu-
tion occur simultaneously in the Universe. For instance, galaxy
merging reduces the number density of galaxies causing thus
evolution. Besides, such events also cause different luminosi-
ties of the final systems and hence leading to luminosity evo-
lution. Other physical mechanism for galaxy evolution can be
devised. Some evidences for a sizable evolution of IRAS galax-
ies have been reported by a number of authors, although there
seem to be much controversy about the amplitude of the effect
(see Springel 1996 and references therein).
Canavezes et al. (1998) expressed evolution in terms of a
generalization of the selection function introduced by Yahil et
al. (1991) given by
φ(s)evol =( (cid:0) rs
1
⋆
rγ
⋆
r(cid:1)α(cid:16) rγ
+rγ
s
+rγ(cid:17)β/γ
, s ≥ H0rs,
, otherwise.
(14)
Springel (1996) showed that it indeed provides a rather accurate
modeling of both IRAS data-sets. The estimated parameters us-
ing a maximum-likelihood procedure are listed in Tab. (1) and
are denoted by PSCzevol. The effect of evolution are shown in
Fig. (14) in which the velocity monopole term derived from
the PSCz galaxy distribution in which evolutionary effects have
been accounted for through the new selection function φ(s)evol
(dot -- dashed line) is plotted against the unevolved model. The
evolutionary effects are very minor, as was somewhat expected
given the locality of our galaxy sample, and do not help in re-
ducing the PSCz vs. 1.2-Jy velocity monopole mismatch.
Overall, our analysis seems to indicate that the discrepancy
between 1.2-Jy and PSCz velocity monopoles cannot be as-
cribed to errors in the modeling of the selection function which,
as we have seen, would not change the monopole terms appre-
ciably. Catalog incompleteness at very low fluxes has to be
ruled out too. We can only conclude that either the PSCz in-
FIG. 14. -- Velocity monopoles for the 1.2-Jy (dark gray line) PSCz (light-
gray), PSCz0.7 (dot-dashed), PSCz1.2 (long-dashed) PSCzevol (dotted) sam-
ples.
completeness extend at objects with fluxes comparable to 1.2
Jy or that the monopole mismatch does reflect a genuine differ-
ence in the average density of faint vs. bright IRAS galaxies at
s ≤ 4000 km s- 1.
A second important result found in our analysis is that the
1.2-Jy and PSCz dipoles agree within the 1-σ errors. This
means that the use of the PSCz rather than the 1.2-Jy catalog
does not help in reducing the Mark III-IRAS dipole residuals
found by Davis, Nusser & Willick (1996). Indeed, both GX
and GY components of the PSCz velocity dipole are much too
large with respect to Mark III ones (see Fig 15 of Davis, Nusser
& Willick 1996). Similar results were also obtained from the
ORS vs. Mark III dipole velocity comparison (BDSLS). We
conclude that the discrepancy between the Mark III and 1.2-
TABLE 3
SELECTION FUNCTION PARAMETERS IN THE IRAS MOCK
CATALOGS.
Mock catalog
α
β
rs
[ h- 1 Mpc]
r⋆
[ h- 1 Mpc]
PSCz
1.2-Jy
0.53 1.90
0.48 1.79
10.90
5.00
86.40
50.40
Teodoro, Branchini & Frenk
11
ple in redshift space. Our model predictions are based on
the Nusser & Davis (1994) spherical harmonics decomposition
technique and assumes Zeld'ovich approximation and linear bi-
asing.
The IRAS overdensity fields are reconstructed with a smooth-
ing proportional to the 1.2-Jy inter-galaxy separation and ex-
hibit the same general features corresponding to the known cos-
mic structures in the local universe. The model velocity field,
computed in the LG frame, has a dipole -- like appearance, with
galaxies moving away from us toward Perseus-Pisces region
and approaching the LG from the opposite direction.
The analysis of the velocity multipoles has revealed a signif-
icant discrepancy between the PSCz and 1.2-Jy velocity mono-
poles in the range 1 800 . s . 4 000 km s- 1 in which PSCz radial
velocities are systematically larger than 1.2-Jy ones, that cannot
be explained by uncertainties in the IRAS selection functions
or incompleteness of the PSCz catalog at faint fluxes. Also,
the mismatch cannot be ascribed to errors derived from shot --
noise and treatment of unobserved areas since both have been
accounted for with the help of realistic mock IRAS catalogs and
the extensive application of bootstrap resampling techniques to
both IRAS datasets.
Both IRAS dipole and quadrupole moments are in good
agreement within the sampled volume apart from a small mis-
match in the u10, u2,- 2 and u22 quadrupole components beyond
s = 7 000 km s- 1. The agreement between the dipole moments
and their consistency with the velocity dipole computed by BD-
Jy velocity fields cannot be alleviated by reducing the shot noise
errors of the model velocity field.
The agreement between the PSCz, 1.2-Jy and ORS dipoles
is reassuring since it guarantees that uncertainties in model ve-
locity fields arising from within the galaxy samples that derive
from incompleteness in one (or more) catalogs, incorrect treat-
ment of non-linear velocities or non-linear non-uniform biasing
do not generate systematic errors apart from a the monopole
mismatch in the range 1 800 km s- 1≤ s ≤ 4 000 km s- 1.
Unlike the case of monopole and dipole moments, possible
differences between the 1.2-Jy and PSCz quadrupole compo-
nents can also be attributed to incorrect determination and di-
lute sampling of the density field beyond the sample. The fact
that 1.2-Jy and PSCz velocity quadrupole agree within the ex-
pected errors suggests that differences in the quadrupole com-
ponents of the two fields can be understood in terms of shot
noise errors and uncertainties in filling the empty regions.
With this respect it is worth investigating whether the denser
sampling of the PSCz catalog at large radii allows a better
modeling of the large scale contribution to the model veloc-
ity field. Indeed, when comparing the model 1.2-Jy velocity
field to Mark III velocities in the framework of the VELMOD
analysis, WSDK have found that velocity residuals exhibit a
quadrupole pattern of the form uQ = VQ r · r, where VQ is a
traceless, symmetric 3 × 3 tensor. They have attributed this
"VELMOD quadrupole" residuals to uncertainties in modeling
the IRAS density field beyond 3 000 km s- 1. In particular they
found that the most important sources of uncertainties were the
smoothing procedure (based on a Wiener Filtering technique)
and the shot noise errors, with the errors deriving from having
ignored mass inhomogeneities beyond 12,000 km s- 1 playing
a minor role. Using PSCz instead of 1.2-Jy catalog reduces the
shot noise errors in the density field beyond 3000 km s- 1 and
thus should improve the agreement with the Mark III velocities.
In other words one would expect the quadrupole of the resid-
ual velocity field u>
1.2, where one only takes in account
0.6
the mass beyond 3000 km s- 1, to be similar to the "VELMOD
quadrupole" of WSDK.
- u>
- u>
In Fig. (15) we compare the VELMOD quadrupole at s =
2000 km s- 1 (rescaled from β = 0.492 to β = 0.6, upper panel) to
the u>
1.2 quadrupole residual (lower panel). The VELMOD
0.6
quadrupole reaches its maximum value at (l ≈ 165◦, b ≈ 55◦)
and in the opposite direction of the sky. The u>
1.2 resid-
0.6
ual quadrupole exhibits a rather different pattern, reaching its
maximum amplitude at (l ≈ 156◦, b ≈ - 74◦). The value of the
quadrupole components of 1.2-Jy and PSCz as well as those of
1.2 quadrupole residuals computed at
the VELMOD and u>
0.6
a distance of 2000 km s- 1 are shown in Tab. (4). This com-
parison is at best qualitative since (i) VELMOD quadrupole is
computed in real space while both IRAS quadrupoles refer to
redshift space insted, and (ii) the smoothing scheme used in
WDKS is different from with ours.
- u>
- u>
7. CONCLUSIONS.
We have compared the model velocity and density fields of
the PSCz survey to those derived from the sparser 1.2-Jy sam-
- u>
FIG. 15. -- VELMOD (upper panel) and u>
1.2 (lower panel) quadrupole
0.6
residuals at s = 2 000 km s- 1. Stars denote outflowing from us (LG frame); cir-
cles denote infall to us. The size of the symbol is proportional to the amplitude
of the flow (notice the different velocity scales in the two panels). The max-
imum amplitude of the "VELMOD quadrupole" (top panel) is ≈ 181 km s- 1,
while for u>
0.6
1.2 is ≈ 23 km s- 1.
- u>
12
PSCz v.s. 1.2-Jy Velocity and Density Fields
SLS from the ORS catalog suggests that deviations from ho-
mogeneous linear biasing prescriptions have to be small and
cannot be advocated to explain the aforementioned mismatch
between the IRAS monopoles.
Another remarkable consequence of the agreement among
the various dipoles is that the 1.2-Jy-Mark III residuals cannot
be alleviated by using the PSCz catalog instead of the 1.2-Jy or
the ORS ones. In fact, this suggests that the origin of the 1.2-
Jy-Mark III dipole resides in the Mark III calibration procedure;
a suggestion corroborated by the fact that the dipole mismatch
disappears when adopting the alternative VELMOD procedure
to calibrate -Mark III velocities (WDSK). The VELMOD anal-
ysis requires a sizable external quadrupole component within
3000 km s- 1. This quadrupole accounts for uncertainties er-
rors in the 1.2-Jy model density field beyond that distance. We
found that using the PSCz sample provides part of the required
external tidal field.
As a final remark, we would like to stress that the present
analysis illustrates once more the importance of spherical har-
monics decomposition techniques to compare observed and
model density and velocity fields obtained from all-sky sam-
ples, to reveal possible data-model or model-model inconsisten-
cies and to assess model adequacy. These techniques will prove
very useful when the next generation of all-sky redshift surveys
and peculiar velocity catalogs [e.g. the 6dF (http:www.mso.anu.
edu.au/6dFGS) and 2MRS (http:cfa-www.harvard.edu/huchra/
2mass) datasets] will become available.
ACKOWLEDGEMENTS
We thank Adi Nusser for enlightening discussions. LT was
partly funded by FCT (Portugal) under the grants PRAXIS XXI
/BPD/16354/98, PRAXIS/C/FIS/13196/98 and POCTI/1999/
FIS/36285. LT aknowledges the hospitality of the Aspen Cen-
ter for Physics and the Kvali Institute for Theoretical Physics
where much of the work was completed.
TABLE 4
QUADRUPOLE COMPONENTS AT 2 000 km s- 1
VELMOD PSCz
[ km s- 1]
[ km s- 1]
1.2-Jy
[ km s- 1]
PSCz - 1.2-Jy
[ km s- 1]
VQ (x,x)...
VQ (y,y)...
VQ (z,z)...
VQ (x,y)...
VQ (x,z)...
VQ (y,z)...
45
44
-89
18
138
-29
57
-22
-35
-150
57
-106
69
-15
-54
-156
48
-102
-13
-7
20
6
10
-5
NOTE. -- (1) We have rescaled the "VELMOD quadrupole" compo-
nents presented in WSDK's Tab. (2) to the numerical value of β = 0.6.
REFERENCES
Baker, J. and Davis, M. and Strauss, M. A. and Lahav, O. and Santiago, B.,
1999, ApJ, 508, 6
Branchini, E. and Teodoro, L. and Frenk, C. S. and Schmoldt, I. and Efstathiou,
G. and White, S. D. M. and Saunders, W. and Sutherland, W. and Rowan-
Robinson, M. and Keeble, O. and Tadros, H. and Maddox, S. and Oliver, S.,
1999, MNRAS,308, 1
Bunn, E. F., 1995, PhD thesis, University of California, Berkeley
Canavezes, A. and Springel, V. and Oliver, S. J. and Rowan-Robinson, M. and
Keeble, O. and White, S. D. M. and Saunders, W. and Efstathiou, G. and
Frenk, C. S. and McMahon, R. G. and Maddox, S. and Sutherland, W. and
Tadros, H., 1998, MNRAS, 297, 777
Cole, S., Weinberg, D. H., Frenk, C. S., and Ratra, B.,1997, MNRAS, 289, 37
Colless, M. and Dalton, G. and Maddox, S. and Sutherland, W. and zes Norberg,
P. and Cole, S. and Bland-Hawthorn, J. and Bridges, T. and Cannon, R. and
Collins, C. and Couch, W. and Cross, N. and Deeley, K. and De Propris, R.
and Driver, S. P. and Efstathiou, G. and Ellis, R. S. and Frenk, C. S. and
Glazebrook, K. and Jackson, C. and Lahav, O. and Lewis, I. and Lumsden,
S. and Madgwick, D. and Peacock, J. A. and Peterson, B. A. and Price, I. and
Seaborne, M. and Taylor, K., 2001, MNRAS, 328, 1039
da Costa, L. N. and Bernardi, M. and Alonso, M. V. and Wegner, G. and
Willmer, C. N. A. and Pellegrini, P. S. and Rité, C. and Maia, M. A. G.,
2000, ApJ, 120, 95
Davis, M. and Nusser, A and Willick, J., 1996, ApJ, 473, 22
Dekel, A. and Bertschinger, E. and Yahil, A. and Strauss, M. A. and Davis, M.
and Huchra, J. P., 1993, ApJ, 412, 1
Dekel, A. and Eldar, A. and Kolatt, T. and Yahil, A. and Willick, J. A. and
Faber, S. M. and Courteau, S. and Burstein, D., 1999, ApJ, 522, 1
Eke, V., Cole, S., and Frenk, C.: 1996, MNRAS, 282, 263
Fisher, K. B. and Huchra, J. P. and Strauss, M. A. and Davis, M. and Yahil, A.
and Schlegel, D., 1995, ApJS, 100, 69
Fisher, K. B. and Lahav, O. and Hoffman, Y. and Lynden-Bel, D. and Zaroubi,
S., 1995, MNRAS, 272, 219
Haynes, M. P. and Giovanelli, R. and Salzer, J. J. and Wegner, G. and Freudling,
W. and da Costa, L. N. and Herter, T. and Vogt, N. P., 1999, ApJ, 117, 1668
Hernquist, L. and Katz, N., 1989, ApJS, 70, 419
Kogut, A., Linewater, C., Smoot, G., Bandey, C. B. A., N.W.Boggess, Cheng,
E., Amici, G. D., Fixsen, D., Hinshaw, G., Jackson, P., Janssen, M., Keegstra,
P., Loewenstein, K., Lubin, P., Mahter, J., Tenorio, L., Weiss, R., Wilkinson,
D., and Wright, E., 1993, ApJ, 419, 1
Lahav, O. and Fisher, K. B. and Hoffman, Y. and Scharf, C. A. and Zaroubi, S.,
Jackson, J., 1999, Classical Electrodynamics, New York: Wiley
Juszkiewicz, R. and Ferreira, P. G. and Feldman, H. A. and Jaffe, A. H. and
Davis, M., 2000, Science, 287, 109
Nusser, A. and Davis, M. , 1994, ApJ, 421, L1
Nusser, A. and da Costa, L. N. and Branchini, E. and Bernardi, M. and Alonso,
M. V. and Wegner, G. and Willmer, C. N. A. and Pellegrini, P. S., 2000,
MNRAS, 320, L21
Rowan-Robinson, M. and Lawrence, A. and Saunders, W. and Crawford, J. and
Ellis, R. and Frenk, C. S. and Parry, I. and Xiaoyang, X. and Allington-Smith,
J. and Efstathiou, G. and Kaiser, N., 1990, MNRAS, 247, 1
Saunders, W. and Sutherland, W. J. and Maddox, S. J. and Keeble, O. and
Oliver, S. J. and Rowan-Robinson, M. and McMahon, R. G. and Efstathiou,
G. P. and Tadros, H. and White, S. D. M. and Frenk, C. S. and Carramiñana,
A. and Hawkins, M. R. S., 2000, MNRAS, 317, 55
Sigad, Y. and Eldar, A. and Dekel, A. and Strauss, M. A. and Yahil, A., 1998,
ApJ, 495, 516
Springel, V.,1996, MSc thesis, Eberhard-Karls-Univerität Tübingen
Strauss, M. A.and Willick, J. A., 1995, PhR, 261, 271
Zaroubi, S. and Branchini, E. and Hoffman, Y. and da Costa, L. N., 2002,
MNRAS, 336, 1234
Tadros, H., Ballinger, W. E., Taylor, A. N., Heavens, A. F., Efstathiou, G.,
Saunders, W., Frenk, C. S., Keeble, O., MCMahon, R., Maddox, S. J., Oliver,
S., Rowan-Robinson, M., Sutherland, W. J., and White, S. D. M.: 1999,
MNRAS, 305, 527
Teodoro, L., 1999, PhD thesis, University of Durham
Willick, J. and Strauss, M. and Dekel, A. and Kolatt, T., 1997, ApJ, 486,629
Willick, J. and Courteau, S. and Faber, S. and Burstein, D. and Dekel, A. and
Strauss, M. A., 1997, ApJS, 109, 333
Willick, J. and Strauss, M. A., 1998, ApJ, 507, 64
Yahil, A. and Strauss, M. A. and Davis, M. and Huchra, J., 1991, ApJ, 372, 380
York, D. G. et al., 2000, ApJ, 120, 1579
Zehavi, I. and Riess, A. G. and Kirshner, R. P. and Dekel, A., 1998, ApJ, 503,
483
|
0801.2867 | 3 | 0801 | 2008-08-01T17:21:52 | The Helium and Heavy Elements Enrichment of the Galactic Disk | [
"astro-ph"
] | We present chemical evolution models for the Galactic disk. We also present a new determination of X, Y, and Z for M17 a Galactic metal-rich Hii region. We compare our models for the Galactic disk with the Galactic Hii regions abundances. The $\Delta Y/\Delta O$ ratio predicted from the galactic chemical evolution model is in very good agreement with the $\Delta Y/\Delta O$ value derived from M17 and the primordial helium abundance, Yp, taking into account the presence of temperature variations in this Hii region. From the M17 observations we obtain that $\Delta Y/\Delta Z = 1.97 \pm 0.41$, in excellent agreement with two $\Delta Y/\Delta Z$ determinations derived from K dwarf stars of the solar vicinity that amount to $2.1 \pm 0.4$ and $2.1 \pm 0.9$ respectively. We also compare our models with the solar abundances. The solar and Orion nebula O/H values are in good agreement with our chemical evolution model. | astro-ph | astro-ph |
Manuscript for Revista Mexicana de Astronom´ıa y Astrof´ısica (2007)
THE HELIUM AND HEAVY ELEMENTS
ENRICHMENT OF THE GALACTIC DISK
L. Carigi1,2 and M. Peimbert1
Draft version: November 13, 2018
RESUMEN
Presentamos modelos de evoluci´on qu´ımica para el disco de nuestra galaxia.
Tambi´en presentamos una nueva determinaci´on de X, Y , y Z para M17, una
regi´on H ii de nuestra galaxia rica en elementos pesados. Comparamos nue-
stros modelos del disco gal´actico con las abundancias de las regiones H ii. El
valor predicho por nuestro modelo para ∆Y /∆O es muy similar al valor que
obtenemos por medio de las observaciones de M17 y la abundancia primordial
de helio, Yp. A partir de M17 y Yp obtenemos que ∆Y /∆Z = 1.97 ± 0.41
resultado que concuerda con dos determinaciones de ∆Y /∆Z, obtenidas a
partir de observaciones de estrellas enanas K de la vecindad solar, que corre-
sponden a 2.1 ± 0.4 y 2.1 ± 0.9 respectivamente. Nuestros modelos ajustan
razonablemente bien el valor de O/H con el que se form´o el Sol.
ABSTRACT
We present chemical evolution models for the Galactic disk. We also present a
new determination of X, Y , and Z for M17 a Galactic metal-rich H ii region.
We compare our models for the Galactic disk with the Galactic H ii regions
abundances. The ∆Y /∆O ratio predicted from the galactic chemical evolution
model is in very good agreement with the ∆Y /∆O value derived from M17
and the primordial helium abundance, Yp, taking into account the presence
of temperature variations in this H ii region. From the M17 observations we
obtain that ∆Y /∆Z = 1.97 ± 0.41, in excellent agreement with two ∆Y /∆Z
determinations derived from K dwarf stars of the solar vicinity that amount
to 2.1 ± 0.4 and 2.1 ± 0.9 respectively. We also compare our models with
the solar abundances. The solar and Orion nebula O/H values are in good
agreement with our chemical evolution model.
Key Words: galaxies: abundances -- galaxies: evolution -- H II regions --
ISM: abundances -- ISM: individual: M17, Orion nebula -- Sun: abundances
1. INTRODUCTION
The main purpose of this work is to study the evolution of the helium
abundance with respect to the heavy elements as a function of time and posi-
tion in the Galactic disk. For this purpose we will use the Galactic chemical
evolution model by Carigi et al. (2005) that has been successful in explaining:
the observed O/H and C/H abundance gradients in the interstellar medium,
1Instituto de Astronom´ıa, UNAM, M´exico.
2Centre for Astrophysics, UCLan, UK.
1
2
CARIGI & PEIMBERT
ISM, the present gaseous distribution in the Galactic disk, the current star
formation rate, the stellar mass as a function of the Galactic radius, and most
of chemical properties of the solar vicinity.
To have useful observational constraints we need accurate X, Y , and Z
determinations or at least ∆Y /∆Z determinations to compare the models
with observations. In this paper we recompute the X, Y , and Z values for
the H ii region M17, the best Galactic H ii region for which it is possible
to compute an accurate enough Y value, for this purpose we make use of
the best observations available and the new He I atomic data needed for the
helium abundance determination. We also compare the chemical evolution
model with the ∆Y /∆Z determination derived from K dwarf stars of the
solar vicinity with metallicities similar or higher than solar by Jim´enez et al.
(2003) and Casagrande et al. (2007).
We also compare our model with the initial solar O/H value and with the
Orion O/H value. The initial solar O/H value is representative of the ISM,
4.5 Gyr ago when the Sun was formed, and the Orion nebula O/H value is
representative of the present day ISM.
The model together with the primordial helium determination is also used
to provide an equation between the Y enrichment and the O enrichment of
the ISM. This equation can be used to provide the initial Y values for those
stars for which we can derive their initial oxygen abundances. These initial Y
values provide meaningful initial abundances for a set of stellar evolutionary
models with different heavy elements content.
We adopt the usual notation X, Y , and Z to represent the hydrogen,
helium and heavy elements abundances by mass, respectively. Based on our
models we study the increase of helium, ∆Y , as a function of the increase of
C, O, F e, and Z by mass.
In Section 2 we discuss the general properties of the chemical evolution
models, we discuss inflow models for the Galaxy with two sets of stellar yields
and present the chemical abundances for the disk at 5 galactocentric distances.
For the two models discussed we present the increase of helium by mass ∆Y ,
relative to the increase of carbon, oxygen, iron and heavy elements by mass,
∆C, ∆O, ∆F e, and ∆Z. We also discuss the evolution of the Y and O
abundances for our models, and present an equation that predicts for the
Galaxy the Y enrichment as a function of the O enrichment of the ISM. In
Section 3 we present a new determination of the helium abundance for the
metal rich H ii region M17; this abundance is compared with our Galactic
chemical evolution models.
In Section 4 we compare abundances for the Orion nebula with those
of B stars of the Orion association.
In Section 5 we discuss the absolute
calibration of the O/H ratio in the local ISM, this ratio is one of the most
important observational constraints for Galactic chemical evolution models,
the absolute calibration of the O/H ratio is obtained based on the Orion nebula
and the solar abundances. In Section 6 we compare the solar initial helium
abundance, inferred from the standard solar models by Bahcall et al. (2006),
HE ENRICHMENT OF THE GALACTIC ISM
3
with our Galactic chemical evolution models considering the time since the
Sun was formed and the presence of gravitational settling and diffusion. The
conclusions are presented in Section 7.
Throughout this paper we will use the primordial helium abundance by
mass, Yp, derived by Peimbert et al. (2007) based on direct helium abun-
dance determinations of metal poor extragalactic H ii regions that amounts
to 0.2477 ± 0.0029. This result is in excellent agreement with the Yp deter-
mination by Dunkley et al. (2008) that amounts to 0.2484 ± 0.0003. This
determination is based on the Ωbh2 value derived from WMAP observations,
the assumption of standard big-bang nucleosynthesis, and the neutron life-
time, τn, of 885.7 ± 0.8 sec obtained by Arzumanov et al. (2000). Following
Mathews et al. (2005) the value of Yp derived from WMAP is revised down-
wards to 0.2468 ± 0.0003 by adopting the τn = 878.5 ± 0.8 sec derived by
Serebrov et al. (2005), and to 0.2475 ± 0.0006 by adopting for τn the new
world average that amounts to 881.9 ± 1.6 sec, average that includes the re-
sults by Arzumanov et al. (2000) and Serebrov et al. (2005).
2. CHEMICAL EVOLUTION MODELS
2.1. Model Parameters
We present chemical evolution models for the Galactic disk using the
CHEMO code (Carigi 1994) that considers the lifetime of each star until
it leaves the main sequence. The models have been built to reproduce the
present gas mass distribution and the present-day O/H values for each galac-
tocentric distance, the values were obtained from observations of H ii regions
in the galaxy (Esteban et al. 2005). Specifically, the characteristics of the
models are:
i) An inside-outside scenario with primordial infalls but without any type
of outflows. The infall rate as a function of time and galactocentric distance
r is given by IN F ALL(r, t) = A(r)e−t/τhalo + B(r)e−(t−1Gyr)/τdisk, where the
formation timescales are τhalo = 0.5 Gyr and τdisk = 6 + (r/r⊙ − 1)8 Gyr. We
assume the location of the solar vicinity is r⊙ = 8 kpc. The constants A(r)
and B(r) are chosen to match, first, the present-day mass density of the halo
and disk components in the solar vicinity, 10 and 40 M⊙pc−2, respectively, and
second, to reproduce the radial profile total mass in the Galaxy, M tot(r) =
50e−(r−8)/3.5 (Fenner et al. 2003).
ii) 13 Gyr as the age of the models, the time elapsed since the beginning
of the formation of the Galaxy.
iii) The Initial Mass Function (IMF) proposed by Kroupa, Tout & Gilmore
(1993), in the mass interval given by 0.01 < m/M⊙ < Mup, with Mup = 80
and 60 M⊙. This IMF is a three power-law approximation, given by IMF
∝ m−α with α = −1.3 for 0.01 - 0.5 M⊙, α = −2.2 for 0.5 - 1.0 M⊙, and
α = −2.7 for 0.5 - Mup.
Note that, our models were computed assuming an IMF with Mlow = 0.01
(1993) truncated their IMF at 0.08 M⊙because they
M⊙. Kroupa et al.
4
CARIGI & PEIMBERT
Fig. 1. Newly formed mass of a given element by massive stars, in M⊙, ejected to the
ISM. The initial heavy elements of the stars amount to Z = 0.02. Continuous lines:
high wind yields by Maeder (1992), dashed lines: low wind yields by Hirschi et al.
(2005).
considered only stars, but in our work we are assuming a non negligible amount
of substellar objects (0.01 < m/M⊙ < 0.08). In models with Mup = 60 M⊙,
the mass of objects with m < 0.08M⊙ is 12 % of the total Mstars, that
includes stars and remnants; this percentage is practically independent of
Mup. Even at present, the fraction of mass in substellar objects is unknown
and we consider that our predicted percentage might be realistic.
Due to the uncertainties in the current Mgas(r), Mstars(r), and SFR(r)
values we cannot discriminate between chemical evolution models assuming
Mlow = 0.01 M⊙and Mlow = 0.08 M⊙. The first ones predict smaller fractions
of massive stars and LIMS per single stellar generation. Therefore Mlow =
0.01 M⊙models with identical galaxy formation scenario, galactic age, and
stellar yields to those of Mlow = 0.08 M⊙models, require a higher SFR to
match the present-day O/H(r) values. The Mlow = 0.01 M⊙model with a
HE ENRICHMENT OF THE GALACTIC ISM
5
more efficient SFR predicts a lower Mgas and similar chemical abundances.
Such model with higher SFR and lower Mgas is also able to reproduce the
observational constraints (Carigi 1996).
iv) A star formation rate that depends on time and galactocentric distance,
that varies from almost constant and low (at large r values) to bursting and
high (at short r values), this SFR has been represented by the following re-
lation SF R(r, t) = νM 1.4
gas(r, t) (Mgas + Mstars)0.4(r, t), in order to reproduce
the current O/H gradient and the gas mass distribution of the Galactic disk
(Carigi 1996), where ν is a constant in time and space that is chosen in order
to reproduce the present-day radial distribution of the gas surface mass den-
sity. A ν value of 0.016 is required when the high-wind yields and Mup = 80
M⊙are adopted, the best model of Carigi et al. (2005), while ν values of 0.015
and 0.010 are required when the low-wind yields with Mup = 60 and Mup = 80
M⊙are adopted, respectively.
v) Two sets of stellar yields. Since the main difference between these
sets is the assumed mass-loss rate due to stellar winds by massive stars with
Z = 0.02, we will call them high-wind yields (HWY) and low-wind yields
(LWY), see Figure 1.
The HWY set is the one considered in the best model (model 1) of Carigi et al.
(2005). The HWY set includes: A) For massive stars (MS), those with
8 < m/M⊙ < 80, the following yields by: a) Chieffi & Limongi (2002) for
Z = 0.00; b) Meynet & Maeder (2002) for Z = 10−5 and Z = 0.004; c)
Maeder (1992) for Z = 0.02 (high mass-loss rate yields presented in his Table
6); d) Woosley & Weaver (1995) only for the Fe yields (Models B, for 12 to
30 M⊙; Models C, for 35 to 40 M⊙; while for m > 40 M⊙, we extrapolated
the m = 40 M⊙Fe yields). B) For low and intermediate mass stars (LIMS),
those with 0.8 ≤ m/M⊙ ≤ 8, we have used the yields by Marigo et al. (1996,
1998) and Portinari et al. (1998) from Z = 0.004 to Z = 0.02. C) For Type Ia
SNe we have used the yields by Thielemann et al. (1993). We have assumed
also that 5 % of the stars with initial masses between 3 and 16 M⊙ are binary
systems which explode as SNIa.
In the LWY set we have updated the yields of massive stars only for Z ∼ 0
and Z = 0.02 assuming the yields by Hirschi (2007) and Hirschi et al. (2005)
respectively. The rest of the stellar yields are those included in the high wind
set.
The main differences between the LWY set and the HWY set are due to the
contribution of massive stars at Z = 0.02. Therefore in Figure 1 we compare
the He, C, and O yields for Z = 0.02. The main difference between the HWY
and the LWY models is due to the stellar yields assumed for massive stars at
high Z. The HWY assume a relatively high mass-loss rate for massive stars
with Z = 0.02 (yields by Maeder 1992), while the LWY assume a relatively
low mass-loss rate for massive stars with Z = 0.02 (yields by Hirschi et al.
2005). These difference between a high and a low mass-loss rate produces
opposite differences in the C and O yields (see Figure 1), the reasons are the
following: a) a high mass-loss rate produces a high loss of C and consequently
6
CARIGI & PEIMBERT
Fig. 2.
Evolution of some common properties for all models of the Galactic
disk: gas mass surface density (M⊙pc−2), star formation rate and infall rates
(M⊙pc−2Gyr−1), at four galactocentric distances, 16, 12, 8, and 4 kpc (continu-
ous, long-dashed, short-dashed, and dotted lines, respectively).
a high C yield, and b) since C is needed to produce O, the high loss of C
reduces the O yield.
Since the solar vicinity and the Galactic disk contain stars and H ii regions
of a broad range of metallicities, our galaxy is a proper laboratory to study
the ∆Y /∆Xi behavior at high Z values and to try to observationally test the
predictions of the HWY models and the LWY models.
2.2. Results
The models presented in this paper reproduce the present stellar and gas
mass distributions in the Galactic disk, the current star formation rate as
a function of the Galactic radius, the O/H gradient evolution inferred from
PNe, the SN rates, the distribution of G-dwarf stars as a function of [Fe/H],
HE ENRICHMENT OF THE GALACTIC ISM
7
Fig. 3. Chemical evolution models for the Galactic disk and the solar vicinity (r = 8
kpc). The left panel shows the C/O evolution in the ISM of the solar vicinity with
O/H. The right panels show the present-day ISM abundance ratios as a function of
galactocentric distance. Continuous lines: high wind yields with Mup = 80 M⊙by
Carigi et al. (2005) , dashed and dotted lines: low wind yields with Mup = 60 and 80
M⊙, respectively (this paper). Filled circles: H ii regions, gas plus dust values; the
gaseous values from Esteban et al. (2005) have been corrected for the dust fraction.
Filled squares: Dwarf stars from Akerman et al. (2004). Open circle: Solar values
from Asplund et al. (2005).
the infall rate, and the evolution of [Xi/Fe] vs [Fe/H]. See Allen et al. (1998),
Carigi (1994), Carigi (1996), Carigi (2000).
In Fig. 2 we present some of those model properties, like gas mass surface
density, star formation rate, and infall rate for four galactocentric distances:
16, 12, 8, and 4 kpc. Infall rate follows the inside-outside scenario, the inner
parts of the Galaxy are formed faster than the outer parts. Since gas mass
comes from infall, mainly, M gas reflects the inside-outside scenario and the
SFR shows similar behaviour as the gas mass, due to the SFR is proportional
to the gas mass.
8
CARIGI & PEIMBERT
TABLE 1
PRESENT DAY VALUES FROM THE GALACTIC DISK MODELS
Galactocentric distance O(tf inal)(10−3) ∆Y /∆C ∆Y /∆O ∆Y /∆F e ∆Y /∆Z
4
8
12
16
4
8
12
16
High wind yields and Mup = 80 M⊙
5.91
8.89
6.38
6.68
4.26
7.01
8.89
4.01
3.29
1.95
3.23
Low wind yields and Mup = 60 M⊙
4.06
10.05
12.81
7.13
6.78
3.28
1.34
7.73
9.44
10.18
3.79
3.99
4.39
14.11
18.36
21.60
25.28
14.13
18.36
21.52
26.11
1.85
1.67
1.62
1.65
1.67
1.67
1.78
1.90
In Figure 3 we show O and C gradients in the Galactic disk and the
evolution of the C/O-O/H relation in the solar vicinity predicted by three
models that combine different yields and IMF Mup values. Models that assume
LWY fail to reproduce the C/O gradient and the C/O values for halo stars or
disk stars. The LWY model with Mup = 60 M⊙reproduces poorly the C/O
gradient and the C/O values in disk stars, and predicts C/O values for halo
stars higher than observed. The LWY model with Mup = 80 M⊙does not
reproduce at all the C/O gradient, matches partially the C/O values of disk
stars, and explains the observed C/O values for halo stars. The HWY model
with Mup = 80 M⊙reproduces successfully the C/O Galactic gradient and the
C/O values in the solar vicinity.
Based on Figure 3 we conclude that the LWY model with Mup = 60
M⊙reproduces the main behavior of C/O vs O/H in the solar vicinity but
cannot reproduce the C/O Galactic gradient. On the other hand, the HWY
model reproduces very well the C/O vs O/H relation in the solar vicinity and
the C/O Galactic gradient. Since the LWY model with Mup = 80 M⊙produces
the poorest fit to the C/O and C/H observed values it will not be considered
further.
Gibson et al (2006) considering the O yields by Arnett (1991), that are
lower than those by Maeder (1992) for MS with m < 30 M⊙, reproduce the
[O/Mg] values present in the Galactic Bulge. With the same yields Gibson
(1997) explains the [O/Fe] values in the intracluster medium and predicts a
small increase in the C/O evolution at late times for a massive elliptical galaxy.
By adopting the Arnett (1991) yields in our model we may obtain flatter C/O
gradients and might not be able to reach the C/O values observed in the H ii
regions and dwarf stars of the solar vicinity. The yields by Arnett (1991) do
not consider stellar winds. Chiappini et al. (2005) using LWY studied the C
and O evolution in the solar vicinity and the Galactic disk, they reproduce
HE ENRICHMENT OF THE GALACTIC ISM
9
also the C/O vs O/H behavior in the solar vicinity, but they predict flatter
C/O gradients than the observed ones and they cannot match the high C/O
values shown by the metal rich star in the solar neighborhood.
Recently, McWilliam et al. (2007) suggested that the strong metallicity-
dependent yields for massive stars by Maeder (1992), can explain the O/Mg
vs O-Mg/H and the O/Fe vs Fe/H relations in the Galactic bulge and in
the solar vicinity, in agreement with our results that favor the HWY model
over the LWY model. The HWY model includes low O yields at high Z
and therefore explains: a) the small O increase in the solar vicinity from the
time the Sun was formed until the present, and b) the flattening of the O
gradient in the direction of the Galactic center. Massive stars with high Z
values have strong winds, lose a considerable amount of C and produce high
C yields. With this C lost, the stars keep a small amount of C needed to
produce O and consequently their O yields are low. Therefore the C yields
for massive stars are important at high metallicities, while their O yields
are more important at low metallicities, as has been shown previously by
Akerman et al. (2004) and Carigi et al. (2005). By adding our previous results
(Akerman et al. 2004; Carigi et al. 2005) to those of McWilliam et al. (2007)
and of this paper we insist that the stellar winds with a high mass loss rate
are essential to reproduce the high C/O values observed in the disk stars of
the solar vicinity.
In the upper panel of Figure 4 we present the evolution of the model that
assumes HWY and Mup = 80 M⊙at different galactocentric radii (4, 8, 12,
and 16 kpc) that correspond to different final metallicities (Z = 0.028, 0.016,
0.009, and 0.004, respectively). The ∆Y /∆O increase at O > 4 × 10−3 present
in Figure 4 is due to the lower O yields for massive stars with Z = 0.02. In
the upper half of Table 1 we show the present-day O values and the ∆Y /∆Xi
values for each galactocentric radius. We note that ∆Y /∆Z increases slightly
with Z for large r values or low O values, while ∆Y /∆Z increases significantly
with Z for short r values or high O values.
In the lower panel of Figure 4 we show the results for the model with
LWY and Mup = 60 M⊙. This model does not predict an increasing ∆Y /∆O
value with increasing O for high O values. In the lower half of Table 1 we
show the present-day O values and the ∆Y /∆Xi values for each galactocentric
radius. With the LWY model we find higher final O values for lower r values
because the O yields are higher than than those of the HWY model at high Z.
Even if the HWY and LWY yields are identical for low Z, we get lower final O
values for higher r values for the LWY model because it does not include stars
with m > 60 M⊙. The increase or decrease of O is reflected on the ∆Y /∆O
and ∆Y /∆Z values because the final Y values are nearly independent of the
models.
The helium to oxygen mass ratio, ∆Y /∆O, is an important constraint in
the study of the chemical evolution of galaxies. We have studied the variation
of ∆Y /∆O as O increases in the HWY and LWY evolution models (see Figure
4). For the HWY model, the model that fits the C/O gradient, we have found
10
CARIGI & PEIMBERT
Fig. 4. Evolution of Helium vs Oxygen for the Galactic disk at four galactocentric
distances, 16, 12, 8, and 4 kpc (continuous, long-dashed, short-dashed, and dotted
lines, respectively). Upper panel: The chemical evolution model assumes Mup =
80 M⊙and high wind yields. Lower panel: The chemical evolution model assumes
Mup = 60 M⊙and low wind yields. See Table 1. Low metallicity H ii regions from
Peimbert et al. (2007) (filled triangles) and adopting Yp = 0.2477.
the following relations between Y and O:
Y = Yp + ∆Y = Yp + (3.3 ± 0.7)O,
(1)
for O < 4.3 × 10−3, and
Y = Yp + (3.3 ± 0.7)O + (0.016 ± 0.003)(O/4.3 × 10−3 − 1)2,
(2)
for 4.3 × 10−3 < O < 9 × 10−3.
For the LWY model with Mup = 60 M⊙we have found the following relation
between Y and O:
Y = Yp + (4.0 ± 0.7)O,
(3)
HE ENRICHMENT OF THE GALACTIC ISM
11
for 0 < O < 11 × 10−3.
Jim´enez et al. (2003) from a set of isochrones and observations of nearby
K dwarf stars found that ∆Y /∆Z = 2.1 ± 0.4. Casagrande et al. (2007)
found also that ∆Y /∆Z = 2.1 ± 0.9 from the newly computed set of Padova
isochrones and observations of nearby K dwarf stars. These observational
results are in very good agreement with the models presented in Table 1.
Assuming yields with low mass-loss rate due to stellar winds (similar to
that considered by Hirschi 2007), and yields without mass loss due to stellar
winds by Woosley & Weaver (1995), Chiappini et al. (2003) find for the solar
vicinity ∆Y /∆Z ∼ 2.4 and 1.5, respectively.
Since stellar winds change significantly the C and O yields, but not the
Y yields, we are interested to quantify the evolution of the Y contribution
due to massive stars and due to low and intermediate mass stars at different
metallicities. For that reason we show in Figure 5 the accumulative percentage
of Y for four galactocentric distances due to MS and LIMS obtained from the
HWY model, our successful model for Galactic disk.
The fraction of helium in the ISM due to MS and and to LIMS depends
strongly on time, but not on galactocentric radius or Z. At present about half
of the ∆Y in the ISM has been produced by MS and half by LIMS.
The strong dependency on time of the ∆Y contribution is due to the
lifetime of the stars. The ∆Y contribution of LIMS decreases less than 5
% from 4 to 16 kpc due mainly to the star formation history. In the inside-
outside scenario the SFR at 4 kpc is more intense and it is also higher at earlier
times than at larger distances (see the middle panel in Figure 2), therefore
the big number of LIMS formed in the first Gyrs at 4 kpc enrich the gas at
later times. This fact produces the O dilution that can be seen in the right
hand side panels of Figure 5, just after the model at 4 kpc reaches the value
of O = 3 × 10−3 for the first time.
3. THE HELIUM AND OXYGEN ABUNDANCES OF M17, A HIGH
METALLICITY GALACTIC H II REGION
We will compare the predictions of the HWY and the LWY models with
observations of Y and O in the Galactic disk. At present the best H ii region
in the Galaxy to derive the Y and O abundances is M17. The reason is that
the correction for the presence of neutral helium in the abundance determina-
tion is the smallest for the well observed Galactic H ii regions. This is due to
the high ionization degree of M17 (Peimbert et al. 1992; Esteban et al. 1999;
Garc´ıa-Rojas et al. 2007). Due to the large amount of neutral helium present
in the other well observed Galactic H ii regions, the error in the Y determi-
nation is at least two times larger than the error for M 17, therefore the Y
determinations for the other Galactic H ii regions will not be considered in
this paper.
To determine very accurate He/H values of a given H ii region we need to
consider its ionization structure. For objects of low degree of ionization it is
necessary to consider the presence of He0 inside the H+ zone, while for objects
12
CARIGI & PEIMBERT
Fig. 5. Cumulative percentage of He as a function of time and oxygen in the ISM,
produced and ejected by massive stars (MS) and low and intermediate mass stars
(LIMS) at four galactocentric distances on the Galactic disk (lines as Figure 2).
The model assumes Mup = 80 M⊙and high mass loss due to stellar winds, the HWY
model.
of high degree of ionization it is necessary to consider the possible presence of a
He++ zone inside the H+ zone. Peimbert et al. (1992), hereafter PTR, found
for M17 an upper limit of N (He++)/N (H+) of 8 ×10−5 a negligible amount;
alternatively they found differences with position of the N (He+)/N H(+) ratio
correlated with the sulphur ionization structure, result that implies that M17
is ionization bounded and the presence of a small but non negligible amount
of He0 inside the H+ zone. Therefore for this object the helium abundance is
given by
N (He0) + N (He+)
N (H+)
N (He)
N (H)
=
.
(4)
To minimize the effect of the correction for neutral helium we took into
account only regions M17-1, M17-2, and M17-3, from now on M17-123, that
show the highest degree of ionization of the observed regions by PTR, Esteban et al.
HE ENRICHMENT OF THE GALACTIC ISM
13
(1999), and Garc´ıa-Rojas et al. (2007), as well as the highest accuracy in the
line intensity determinations by PTR. Following PTR we will assume that He
is neutral in the regions where S is once ionized, that is
N (He0)
N (He)
=
N (S+)
N (S)
,
therefore
N (He)
N (H)
= ICF(He) ×
N (He+)
N (H+)
= (cid:20)1 +
N (S+)
N (S) − N (S+)(cid:21) ×
N (He+)
N (H+)
.
(5)
(6)
To estimate the ICF(He) value we recomputed the N (S+)/N (S) ratios derived
by PTR taking into account that the [S ii] λλ 4069 + 4076 lines are blended
with the O ii λλ 4069 and 4076 lines, the correction diminishes the [S ii] elec-
tron temperatures from about 12 000 K to about 7700 K (Garc´ıa-Rojas et al.
2007), the lower temperatures increase the N (S+)/N (S) ratios, and the av-
erage ICF (He) for M17-123 amounts to 1.035.
To obtain the N (He+)/N H(+) value for M17-123 we decided to recom-
pute the determinations by PTR based on their line intensities and the new
helium recombination coefficients by Porter et al. (2005), with the interpo-
lation formula provided by Porter et al. (2007).
In addition we used the
hydrogen recombination coefficients by Storey & Hummer (1995), and the
collisional contribution to the He i lines by Sawey & Berrington (1993) and
Kingdon & Ferland (1995). The optical depth effects in the triplet lines were
estimated from the computations by Benjamin et al. (2002). At the tem-
peratures present in M17 the collisional excitation of the hydrogen lines is
negligible and was not taken into account.
To determine the N (He+)/N (H+) value we took into account the following
He i lines λλ 3889, 4026, 4471, 4922, 5876, 6678, and 7065. We corrected the
4922 line intensity by considering that it was blended with the [Fe iii] 4924
line and that the contribution of the Fe line amounted to 5% of the total line
intensity (Esteban et al. 1999; Garc´ıa-Rojas et al. 2007). The M17-123 line
intensities adopted are presented in Table 2.
We did not correct the H and He line intensities for underlying absorp-
tion, the reasons are the following: a) the average observed equivalent width in
emission of H(β), EWem(Hβ), amounts to 668 A , b) the predicted EWem(Hβ)
for Te = 7000 K amounts to about 2000 A (Aller 1984), therefore about 1/3 of
the continuum is due to the nebular contribution and 2/3 to the dust scattered
light from OB stars, consequently the underlying stellar absorption only affects
two thirds of the observed continuum, d) considering the nebular contribu-
tion to the observations and based on the models by Gonz´alez Delgado et al.
(1999, 2005) for a model with an age of 2 Myrs as well as the observations by
Leone & Lanzafame (1998) for λ 7065, (since λ 7065 was not included in the
models by Gonz´alez Delgado et al. 1999, 2005), we estimated that the EWab
of the λλ 3889, 4026, 4471, 4922, 5876, 6678, and 7065 lines amount to 0.4,
0.4, 0.4, 0.1, 0.1, 0.2 A respectively, an almost negligible amount considering
14
CARIGI & PEIMBERT
TABLE 2
HE I LINE INTENSITIES RELATIVE TO Hβ FOR M17-123
He i Line
I
3889
4026
4471
4922
5876
6678
7065
7281
0.1738 ± 0.0086
0.0240 ± 0.0012
0.0508 ± 0.0013
0.0127 ± 0.0010
0.1549 ± 0.0038
0.0395 ± 0.0010
0.0499 ± 0.0025
0.0064 ± 0.0007
TABLE 3
PHYSICAL CONDITIONS AND CHEMICAL ABUNDANCES IN M17
Parameter
t2 = 0.000
t2 = 0.036 ± 0.013
T [O ii + O iii]
8300 ± 200
n
τ3889
N (He+)/N (H+)
ICF (He)
691 ± 246
9.5 ± 0.8
8300 ± 200
744 ± 247
11.0 ± 0.9
0.1014 ± 0.0014
0.0982 ± 0.0019
1.035 ± 0.010
1.035 ± 0.010
N (He)/N (H)
0.1049 ± 0.0017
0.1016 ± 0.0022
Y
∆Y
O
∆Y /∆O
Z
∆Y /∆Z
0.2926 ± 0.0034
0.0403 ± 0.0044a
0.00446 ± 0.00045
0.2837 ± 0.0044
0.0360 ± 0.0053b
0.00811 ± 0.00081
9.04 ± 1.35a
4.44 ± 0.79 b
0.0101 ± 0.0015
0.0183 ± 0.0027
4.00 ± 0.75a
1.97 ± 0.41 b
aWhere we adopted Yp = 0.2523 ± 0.0027
for t2 = 0.000 from Peimbert et al. (2007).
bWhere we adopted Yp = 0.2477 ± 0.0029
for t2 6= 0.000 from Peimbert et al. (2007).
the large EWem observed values (see Table 4 in Peimbert et al. 1992), e) the
weighted increase in the helium line intensities amounts to about 0.7%, again
an almost negligible amount, f) the Balmer lines also show underlying absorp-
tion and the average correction to the line intensities amounts to about 0.5%,
again a negligible amount that cancels to a first approximation the underlying
correction effect on the He/H line ratios.
HE ENRICHMENT OF THE GALACTIC ISM
15
TABLE 4
PRESENT DAY VALUES FROM THE GALACTIC DISK MODELS FOR
R = 6.75 KPC
Model
O(tf inal)(10−3) ∆Y /∆C ∆Y /∆O ∆Y /∆F e ∆Y /∆Z
HWY Mup = 80 M⊙
LWY Mup = 60 M⊙
7.16
8.28
6.54
7.26
4.47
3.73
17.04
17.04
1.70
1.62
To determine the helium physical conditions of the nebula simultaneously
with the N (He+)/N (H+) value we used the maximum likelihood implemen-
tation presented by Peimbert et al. (2002). This implementation requires as
inputs: a) the oxygen temperatures, T [O iii] and T [O ii], and the oxygen
ionization degree that provide us with the following restriction
T [O II + O III] =
N (O+)T [O II] + N (O++)T [O III]
N (O+) + N (O++)
,
(7)
and b) a large set of helium to hydrogen line intensity ratios. In addition an
estimate of the electron density, n, in the region where the He lines originate
is not required but it is useful. This implementation can determine the con-
ditions of the H ii regions either with the restriction of uniform temperature,
or relaxing this restriction.
From PTR we adopted T [O iii] = 8200 ± 200 K. From the I(3727/7325)
ratios for M17-123 by PTR and for M17-3 by Esteban et al. (1999), after
correcting the λ 7325 A lines for the recombination contribution (Liu et al.
2000), we obtained 8100 ± 1300 K and 9900 ± 1300 K respectively, therefore
we adopted for T [O ii] a value of 9000 ± 1000 K. From the T [O iii] and T [O ii]
values and the observations by PTR we find that N (O+)/N (H+) = 0.12 and
N (O++)/N (H+) = 0.88. Finally from the previous results and equation (7)
we obtained that T [O ii + O iii] = 8300 ± 200 K.
To estimate the electron density we used three determinations the n[S ii]
for M17-123 by PTR that amounts to 720 ± 250 cm−3, the n[O ii] for M17-3
by Esteban et al. (1999) that amounts to 790 ± 250 cm−3, and the n[Cl iii] for
M17-123 that we estimated from the observed I(5518)/I(5538) ratios by PTR
and the atomic physics parameters by Keenan et al. (2000) that amounts to
650 ± 450 cm−3. From the average of these three determinations we adopted
a value of n = 740 ± 250 cm−3 for M17-123.
By using as inputs for the maximum likelihood method T [O ii + O iii]
= 8300 ± 200 K, n = 740 ± 250 cm−3, and the helium line intensities pre-
sented in Table 2 we obtain for M17-123 the n, τ3889, and N (He+)/N (H+)
values presented in Table 3. The results for M17-123 are presented in Ta-
ble 3 for t2 = 0.000 (constant temperature over the observed volume) and
for t2 6= 0.000 (the temperature variations method). Without the restriction
of uniform temperature the maximum likelihood of the temperature fluctu-
ation parameter amounts to t2 = 0.036 ± 0.013. This t2 value is in good
16
CARIGI & PEIMBERT
Fig. 6. Predicted evolution of He vs Oxygen for r = 6.75 kpc. Continuous line:
model assumes Mup = 80 M⊙and high wind yield. Dotted line: model assumes
Mup = 60 M⊙and low wind yield. M17 H ii region at r = 6.75 kpc (filled circle,
t2 6= 0.000) (open circle, t2 = 0.00)
agreement with those for M17 derived by PTR, Esteban et al. (1999), and
Garc´ıa-Rojas et al. (2007) that are in the 0.033 to 0.045 range; these values
were determined with two different methods: a) combining the temperature
derived from the ratio of the Balmer continuum to the Balmer line intensi-
ties with the temperature derived from I(4363)/I(5007) [O iii] ratio, and b)
combining the O ii recombination line intensities with the λ 5007 [O iii] line
intensities.
From the mean values of N (He+)/N (H+) given in Table 3 and the ICF(He)
given by equation (6) we obtain N (He)/N (H) ratios for M17-123 of 0.1049
and 0.1016 for t2 = 0.000 and t2 = 0.036 respectively. These values are similar
to but more precise than those derived by PTR for M17-123, that amount to
0.106 and 0.100 for t2 = 0.000 and t2 = 0.040 respectively.
We obtained the ∆Y /∆O and the ∆Y /∆Z values presented in Table 3
HE ENRICHMENT OF THE GALACTIC ISM
17
based on the Yp determinations by Peimbert et al. (2007). The O abundance
presented in Table 3 includes both the gaseous and the dust contribution
and corresponds to the average of the values derived by PTR, Esteban et al.
(1999), and Garc´ıa-Rojas et al. (2007), these three values are in excellent
agreement.
M17 is located at a galactocentric distance of 6.75 kpc, under the assump-
tion that the Sun is located at a galactocentric distance of 8 kpc (Dias et al.
2002). In Figure 6 we have plotted the ∆Y /∆O value for t2 = 0.036, from
this figure it can be noted that this value is in good agreement, at about the
one σ level, with the Galactic chemical evolution model based on the HWY
set, and that the O value corresponds to the prediction by the models for a
galactocentric distance of 6.75 kpc (see also Table 4); alternatively the M17
∆Y /∆O value for t2 = 0.000 is considerably higher, by more than 3σ, than
the value predicted by the HWY model. Similarly in Figure 6 we compare
the M17 results with the LWY model, again the values for t2 = 0.036 are
in good agreement with the Galactic chemical evolution model while the val-
ues for t2 = 0.000 are not. Furthermore from Table 4 it is also found that
the t2 = 0.036 value for ∆Y /∆Z is within one σ of the models predictions,
while the t2 = 0.000 value for ∆Y /∆Z is about 3σ away form the models
predictions.
With the present accuracy of the ∆Y /∆O and ∆Y /∆Z determinations in
Galactic H ii regions it is not possible to distinguish between the HWY and
the LWY models.
4. COMPARISON OF STELLAR AND NEBULAR ABUNDANCES IN
ORION
To test the nebular abundances derived with different t2 values for the
Orion nebula we decided to compare them with those derived for B star abun-
dances of the Orion association. Cunha et al. (2006) obtained 12 + log O/H
= 8.70±0.09 from 11 B stars, while Lanz et al. (2008) obtained 12 + log Ar/H
= 6.66 ± 0.06 from 10 B stars. These results are in excellent agreement with
the nebular abundances derived by Esteban et al. (2004) for t2 6= 0.00 that
amount to 8.73 ± 0.03 and 6.62 ± 0.05 for O and Ar respectively. Alternatively
the values derived for t2 = 0.00 amount to 8.59 ± 0.03 and 6.50 ± 0.05 for
O and Ar respectively, values that are about 1σ and 3σ smaller than those
derived from B stars.
5. ABSOLUTE CALIBRATION OF THE O ABUNDANCE IN THE
SOLAR VICINITY
The predicted abundances by chemical evolution models are often com-
pared with stars and with H ii regions abundances. The most popular com-
parisons are made with the solar and the Orion nebula abundances. The
comparisons among the models, the Sun, and the Orion nebula are based on
the absolute abundances, therefore it is necessary to estimate not only the
18
CARIGI & PEIMBERT
statistical errors but also the systematic ones in the observational determina-
tions.
We will start by considering the solar O/H abundance. What we want
from the Sun is the O/H value when it was formed, the so called initial value,
and to keep in mind that it is representative of the ISM 4.5 Gyr ago when
the Sun was formed. We have also to consider that the photospheric and the
interior solar abundances might be different due to diffusion and gravitational
settling.
In Table 5 we present the most popular O/H photospheric determina-
tions of the last fifty years, the quoted values and the errors are the orig-
inal ones published by the authors. By looking at the differences among
the different determinations it is clear that for many determinations prob-
ably the errors represent only the statistical errors and that systematic er-
rors have not been taken into account, in short that the total errors have
been underestimated. The determinations, by Asplund et al. (2005) and by
Allende Prieto (2007), are qualitatively different to the previous five because
they are based on 3D models while the other five are based on 1D models.
The last two determinations included in Table 5, those by Caffau et al. (2008)
and Centeno & Socas-Navarro (2008), indicate that the possibility of a further
revision of the solar abundance is still open.
The abundances inferred from interior models of the Sun, that are based on
stereosismological data, are in disagreement with the 3D photospheric models
and predict heavy element abundances about 0.2 dex higher than the 3D
photospheric ones (e. g. Basu & Antia 2007, and references therein).
To compare with our models we will use as the low O/H value the 3D
photospheric determination by Asplund et al. (2005) and as the high O/H
value the 1D photospheric determination by Grevesse & Sauval (1998), that
is in good agreement with the helioseismic determination. The next step is
to have reliable stellar interior models to determine the initial O/H value,
for this purpose we will use the models by Bahcall et al. (2006) presented in
Table 6.
The models by Bahcall et al. (2006) indicate that the photospheric values
of Z and Y do not represent the initial values due to diffusion and gravitational
settling. By assuming that the O/Z ratio is not affected by these processes
the initial O/H values correspond to 12 + log O/H = 8.70 and to 8.89 for the
AGS05 and the GS98 surface abundances respectively. To obtain the present
day ISM value we have to consider that the Sun was formed 4.5 Gyr ago and
according to the Galactic chemical evolution model by Carigi et al. (2005) the
O/H ratio in the ISM has increased by 0.13 dex since the Sun was formed.
Therefore our determinations of the present day ISM 12 + O/H values based
on the AGS05 and the GS98 abundances amount to 8.83 and 9.02 respectively.
In Table 5 we also present the most popular O/H determinations for the
Orion nebula including the errors presented in the original papers. The pre-
dictions from the solar abundances and the chemical evolution models have to
be compared with the total abundances in the nebula that have to include gas
HE ENRICHMENT OF THE GALACTIC ISM
19
TABLE 5
12 + LOG(O/H)
Solar photosphere a
Year
t2 6= 0.00
t2 = 0.00
Year
Orion nebula b
8.96
8.77 ± 0.05
8.84 ± 0.07
8.93 ± 0.035
8.83 ± 0.06
8.736± 0.078
8.66 ± 0.05
8.65 ± 0.03
8.86 ± 0.07
8.76 ± 0.07
1960(1)
1968(2)
1976(3)
1989(4)
1998(5)
2001(6)
2005(7)
2007(8)
2008(9)
2008(10)
8.79 ± 0.12
...
8.75 ± 0.10
8.52 ± 0.10
...
8.49 ± 0.08
8.72 ± 0.07
8.55 ± 0.07
...
...
8.51 ± 0.08
8.49 ± 0.06
8.73 ± 0.03
8.59 ± 0.03
1969(11)
1977(12)
1992(13)
1998(14)
2000(15)
2003(16)
2004(17)
a
1 Goldberg et al.
(1960).
(1998).
(2001).
8 Allende Prieto (2007).
2
(1968).
3 Ross & Aller
4 Anders & Grevesse (1989).
6
7 Asplund et al.
9
10
Lambert
(1976).
5 Grevesse & Sauval
Holweger
(2005).
Centeno & Socas-Navarro (2008).
Caffau et al. (2008).
b 11 Peimbert & Costero (1969).
12
(1977).
Peimbert & Torres-Peimbert
14
13 Osterbrock et al.
Esteban et al.
in-
(1998),
cludes the fraction of O tied up in dust
15
grains that amounts to 0.08 dex.
Deharveng et al. (2000).
16 Pilyugin
(2003).
17 Esteban et al. (2004), this
value includes the fraction of O tied up in
dust grains that amounts to 0.08 dex.
(1992).
this value
and dust. With the exception of the determinations by Esteban et al. (1998,
2004) that take into account the dust fraction all the other determinations only
include the gaseous content. The other difference is that there are two possible
sets of values: a) those that assume constant temperature over the observed
value given by the 4363/5007 ratio of [O iii], the t2 = 0.00 case, where t2 is
the mean square temperature variation (Peimbert 1967), or b) those based on
the O ii recombination lines, that are in agreement with those derived from
the 5007/Hβ ratio taking into account the presence of temperature variations
over the observed volume, and consequently that t2 6= 0.00.
20
CARIGI & PEIMBERT
TABLE 6
STANDARD SOLAR MODELS A
Values
initial X
initial Y
initial Z
initial O
initial O/H
initial ∆Y /∆Z
initial ∆Y /∆O
surface X
surface Y b
surface Z
surface O/H b
GS98
0.70866
0.27250
0.01884
0.00879
8.89
1.32
2.82
0.7410
AGS05
0.72594
0.26001
0.01405
0.00582
8.70
0.88
2.11
0.7586
0.2420 ± 0.0072
0.2285 ± 0.0067
0.0170
0.0122
8.83 ± 0.17
8.66 ± 0.17
heavy
element
The GS98
aStandard solar models by Bahcall et al.
(2006).
and AGS05
correspond to models with
columns
de-
the
rived from photospheric
observations
(1998) and by
by Grevesse & Sauval
Asplund et al. (2005) respectively.
bThe errors are the conservative ones
adopted by Bahcall et al. (2006).
abundances
Esteban et al. (2004) obtain for the Orion nebula that 12 + O/H = 8.73
for t2 6= 0.00, value that includes the dust correction. By taking into account
the presence of the O/H Galactic abundance gradient that amounts to −0.044
dex kpc−1 (Esteban et al. 2005) we obtain a value of 12 + O/H = 8.75 for
the local ISM. This value is smaller than the values estimated for the local
ISM based on: the solar photospheric abundances by AGS05 and the GS98,
the standard solar models by Bahcall et al. (2006), and the chemical evolution
of the Galaxy that amount to 8.83 and 9.02 respectively. Similarly from the
results by Esteban et al. (2004) for the Orion nebula for t2 = 0.00 and the
observed Galactic gradient we obtain that 12 + O/H = 8.51 for the local
ISM. From the previous discussion it follows that the best agreement for the
derived O/H ISM value is given by the t2 6= 0.00 result from Orion and the
AGS0 result from the Sun. From these two determinations we recommend for
the present day local ISM the value 12 + log O/H = 8.79 ± 0.08.
From the previous discussion it follows that the best agreement between
the solar and the Orion nebula abundances is obtained for the high nebular
abundances, that are derived from the t2 6= 0.00 values and that include the
fraction of atoms tied up in dust grains, and the AGS05 solar value.
HE ENRICHMENT OF THE GALACTIC ISM
21
6. THE SOLAR HELIUM ABUNDANCE
We want to compare also our Galactic chemical evolution model with the
helium abundance when the Sun was formed, the initial Y value. Basu & Antia
(2004) based on seismic data have derived the Y value in the solar convective
envelope and amounts to 0.2485±0.0034. To derive the initial value we need a
model of the solar interior that takes into account helium and heavy elements
diffusion and that agrees with the helium abundance of the envelope.
Again we have at our disposal the solar interior models by Bahcall et al.
(2006) presented in Table 6. The GS98 model agrees with the Y value in the
envelope derived by Basu & Antia (2004), but not with the O/H value in
the envelope derived by Asplund et al. (2005). On the other hand the AGS05
model agrees with the O/H value in the envelope derived by Asplund et al.
(2005) but not with the Y value derived by Basu & Antia (2004). Based
on the discussion of the previous section we conclude that the Orion nebula
O/H value agrees with the O/H value predicted by the Galactic chemical
evolution model and the photospheric value by Asplund et al. (2005) and not
with the photospheric value by Grevesse & Sauval (1998). The discrepancy
between the photospheric abundances by Asplund et al. (2005) and the Y
value derived from helioseimological data is a very important open problem,
an excellent review discussing this issue has been presented by Basu & Antia
(2007).
Bahcall et al. (2006) present the initial Y , Z, and O solar values for
their standard solar models, see Table 6. By adopting the Yp value by
Peimbert et al. (2007) for t2 6= 0.00 it is also possible to obtain the ∆Y /∆Z
and the ∆Y /∆O values for the GS98 and the AGS05 standard solar models.
The values so derived are in fair agreement with the predictions of the HWY
and LWY Galactic chemical evolution models. To try to make a more rigorous
comparison between the solar interior and Galactic chemical evolution models
it is necessary to estimate the errors in the initial solar values by Bahcall et al.
(2006), a task which is beyond the scope of this paper.
7. CONCLUSIONS
Based on the HWY model we find the following equation to estimate the
initial helium abundance with which stars form in the Galactic disk
for O < 4.3 × 10−3, and
Y = Yp + (3.3 ± 0.7)O,
Y = Yp + (3.3 ± 0.7)O + (0.016 ± 0.003)(O/4.3 × 10−3 − 1)2,
for 4.3 × 10−3 < O < 9 × 10−3.
The increase of ∆F e/∆Z has to be taken into account in order to deter-
mine the ∆Y /∆Z value based on the [Fe/H] abundances of stars in the solar
vicinity.
22
CARIGI & PEIMBERT
High mass loss rates due to stellar winds should be adopted in the evolu-
tionary stellar models for massive stars of high metallicity, because only the
Galactic chemical evolution models with HWY can reproduce simultaneously
the O/H and C/O Galactic gradients, the C/O versus O relation in the solar
vicinity, and the ∆Y /∆O value in the inner Galactic disk.
Based on the O/H value of the Orion nebula and the solar photospheric
value together with a chemical evolution model of the Galaxy we recommend
for the present day local ISM a value of 12 + log O/H = 8.79 ± 0.08, where
both the gaseous and the dust components of O are taken into account.
By comparing the O/H value of the Orion nebula with the solar value
we find that the nebular ratio derived using O recombination lines, that is
equivalent to the use of the forbidden O lines under the adoption of a t2 6= 0.00,
is in considerably better agreement with the initial solar value than the Orion
nebula value derived adopting t2 = 0.00.
The stellar O/H and Ar/H abundance ratios derived by Cunha et al. (2006)
and Lanz et al. (2008) for B stars of the Orion association are in excellent
agreement with the nebular abundance ratios derived from the t2 6= 0.00 val-
ues for the Orion nebula.
The ∆Y /∆Z = 1.97 ± 0.41 value derived from observations of M17 for
t2 = 0.036 is in very good agreement with the 2.1 ± 0.4 and the 2.1 ± 0.9 values
derived by Jim´enez et al. (2003) and Casagrande et al. (2007) from K dwarf
stars of the solar vicinity. On the other hand the value ∆Y /∆Z = 4.00 ± 0.75
derived from observations of M17 for t2 = 0.000 is not.
Both Galactic chemical evolution models with the HWY set and the LWY
set are in agreement with the observed ∆Y /∆Z for t2 = 0.036 but not with
the ∆Y /∆Z for t2 = 0.000. Higher accuracy determinations of ∆Y /∆Z for
high metallicity objects are needed to discriminate between the HWY model
and the LWY model predictions.
We are grateful to Brad Gibson and Antonio Peimbert for several fruitful
discussions. We are also grateful to the anonymous referee for some excellent
suggestions. This work was partly supported by the CONACyT grants 46904
and 60354.
REFERENCES
Akerman, C. J., Carigi, L., Nissen, P. E., Pettini, M., & Asplund, M. 2004, A&A,
414, 931
Allen, C., Carigi, L., & Peimbert, M. 1998, ApJ, 494, 247
Allende Prieto, C. 2007, 14th Cambridge Workshop on Cool Stars, Stellar Systems
and the Sun, ed. G. van Belle, ASP Conference Series, in press, astro-ph/0702429
Aller, L. H. 1984, Physics of Thermal Gaseous Nebulae (Dordretch: Reidel)
Anders, E., & Grevesse, N. 1989, Geochimica et Cosmochimica Acta, 53, 197
Arnett, D. 1991, in: Frontiers of Stellar Evolution, ed. D. L. Lambert, ASP Confer-
ence Series, 20, 389
HE ENRICHMENT OF THE GALACTIC ISM
23
Arzumanov, S., et al. 2000, Physics Letters B, 483, 15
Asplund, M., Grevesse, N., & Sauval, A. J. 2005, in: Cosmic Abundances as Records
of Stellar Evolution and Nucleosynthesis, ed. F. N. Bash & T. G. Barnes, ASP
Conference Series, 336, 25
Bahcall, J. N., Serenelli, A. M., & Basu, S. 2006 ApJS, 165, 400
Basu, S., & Antia, H. M. 2004, ApJ, 606, L85
Basu, S., & Antia, H. M. 2007, Physics Reports, in press, arXiv:0711.4590
Benjamin, R. A., Skillman, E. D., & Smits, D. P. 2002, ApJ, 569, 288
Caffau, E., Ludwig, H.-G., Steffen, M., Ayres, T. R., Bonifacio, P., Cayrel, R., Frey-
tag, B., Plez, B. 2008, A&A, submitted, arXiv0805.4398
Carigi, L. 1994, ApJ, 424, 18
Carigi, L. 1996, RevMexAA, 32, 179
Carigi, L. 2000, RevMexAA, 36, 171
Carigi, L., Peimbert, M., Esteban, C., & Garc´ıa-Rojas, J. 2005, ApJ, 623, 213
Casagrande, L., Flynn, C., Portinari, L., Girardi, L. & Jim´enez, R. 2007, MNRAS,
382, 1516
Centeno, R., & Socas-Navarro, H. 2008, ApJ, 682, L61
Chiappini, C., Matteucci, F., & Ballero, S. K. 2005, A&A, 437, 429
Chiappini, C., Matteucci, F., & Meynet, G. 2003, A&A, 410, 257
Chieffi A., & Limongi M. 2002, ApJ, 577, 281
Chiosi, C., & Matteucci, F. 1982, A&A, 105, 140
Cuhna, K., Hubeny, I., & Lanz, T. 2006, ApJ, 647, L143
Deharveng, L., Pena, M., Caplan, J., & Costero, R. 2000, MNRAS, 311, 329
Dias, W. S., Alessi, B. S., Moitinho, A., & L´epine, J. R. D. 2002, A&A, 389, 871
Dunkley, J., et al. 2008, ApJ, submitted, arXiv0803.0586
Esteban, C., Garc´ıa-Rojas, J., Peimbert, M., Peimbert, A., Ruiz, M. T., Rodr´ıguez,
M., & Carigi, L. 2005, ApJ, 618, L95
Esteban, C., Peimbert, M., Garc´ıa-Rojas, J., Ruiz, M. T., Peimbert, A., &
Rodr´ıguez, M. 2004, MNRAS, 355, 229
Esteban, C., Peimbert, M., Torres-Peimbert, S., & Escalante, V. 1998, MNRAS, 295,
401
Esteban, C., Peimbert, M., Torres-Peimbert, S., & Garc´ıa-Rojas J. 1999,
RevMexAA, 35, 65
Fenner, Y., & Gibson, B. K. 2003 PASA, 20, 189
Garc´ıa-Rojas, J., Esteban, C., Peimbert, A., Rodr´ıguez, M., Peimbert, M., & Ruiz,
M. T. 2007, RevMexAA, 43, 3
Gibson, B. K. 1997, MNRAS, 290, 471
Gibson, B. K. , MacDonald, A. J., Sanchez-Blazquez, P. Carigi, L. 2006, in The
Metal-Rich Universe, eds. G. Israelian & G. Meynet, (Cambridge Univ. Press),
in press (arXiv:astro-ph/0611879)
Goldberg, L., Muller, E. A., & Aller, L. H. 1960, ApJS, 5, 1
Gonz´alez Delgado, R. M., Cervino, M., Martins, L. P., Leitherer, C., & Hauschildt,
P. H. 2005, MNRAS, 357, 945
Gonz´alez Delgado, R. M., Leitherer, C., & Heckman, T. M. 1999, ApJS, 125, 489
Grevesse, N., & Sauval, A.J. 1998, Space Sci. Rev. 85, 161
Hirschi, R. 2007, A&A, 461, 571
Hirschi, R., Meynet, G., & Maeder, A. 2005, A&A, 433, 1013
Holweger, H. 2001,
in Solar and Galactic Composition, ed. R. F. Wimmer-
Schweingruber, AIP Conference Proceedings, 598, 23
24
CARIGI & PEIMBERT
Jim´enez, R., Flynn, C., MacDonald, J., & Gibson, B. K. 2003, Science, 299, 1552
Keenan, F. P., Aller, L. H., Ramsbottom, C. A., Bell, K. L., Crawford, F. L., &
Hyung, S. 2000, P. Natl. Acad. Sci. USA, 97, 4551
Kingdon, J., & Ferland, G. 1995, ApJ, 442, 714
Kroupa, P., Tout, C. A., & Gilmore, G. 1993, MNRAS, 262, 545
Lambert, D. L. 1968, MNRAS, 138, 143
Lanz, T., Cunha, K., Holzman, J., Hubeny, I. 2008, ApJ, 678, 1342
Leone, F., & Lanzafame, A. C. 1998, A&A, 330, 306
Liu, X.-W., Storey, P. J., Barlow, M. J., Danziger, I. J., Cohen, M., & Bryce, M.
2000, MNRAS, 312, 585
Maeder, A. 1992, A&A, 264, 105
Marigo, P., Bressan, A., & Chiosi, C. 1996, A&A, 313, 545
Marigo, P., Bressan, A., & Chiosi, C. 1998, A&A, 331, 580
Mathews, G. J., Kajino, T., & Shima, T. 2005, Physical Review D, 71, 021302
McWilliam, A., Matteucci, F., Ballero, S., Rich, R. M., Fulbright, J. P., & Cescutti,
G. 2007, AJ, submitted (arXiv:0708.4026)
Meynet G., & Maeder A. 2002, A&A, 390, 561
Osterbrock D. E., Tran H. D., & Veilleux S. 1992, ApJ, 389, 196
Peimbert, A., Peimbert, M., & Luridiana, V. 2002 ApJ, 565, 668
Peimbert, M. 1967, ApJ, 150, 825
Peimbert, M., & Costero, R. 1969, Bol. Obs. Tonantzintla y Tacubaya, 5, 3
Peimbert, M., Luridiana, V., & Peimbert, A. 2007, ApJ, 666, 636
Peimbert, M., & Torres-Peimbert, S. 1977, MNRAS, 179, 217
Peimbert, M., Torres-Peimbert, S., & Ruiz, M. T. 1992, RevMexAA, 24, 155 (PTR)
Pilyugin, L. S. 2003, A&A, 401, 557
Porter, R. L., Bauman, R. P., Ferland, G. J., & MacAdam, K. B. 2005, ApJ, 622L,
73
Porter, R. L., Ferland, G. J., & MacAdam, K. B. 2007, ApJ, 657, 327
Portinari, L., Chiosi, C., & Bressan, A. 1998, A&A, 334, 505
Ross, J. E., & Aller, L. H. 1976, Science, 191, 1223
Sawey, P. M. J., & Berrington, K. A., 1993, Atomic Data and Nuclear Data Tables,
55, 81
Serebrov, A., et al. 2005, Physics Letters B, 605, 72
Storey, P. J., & Hummer, D. G. 1995, MNRAS, 272, 41
Thielemann, F. K., Nomoto, K., & Hashimoto, M. 1993, in Origin and Evolution of
the Elements eds. N. Prantzos et al., Cambridge University Press, p. 297
Woosley, S. E., & Weaver, T. A. 1995, ApJS, 101, 181
Both authors: Instituto de Astronom´ıa, Universidad Nacional Aut´onoma de
M´exico, Apdo. Postal 70-264, M´exico 04510 D.F., M´exico (carigi, peim-
[email protected]).
Leticia Carigi: Centre for Astrophysics, University of Central Lancashire, Pre-
ston, Lancashire, PR1 2HE, United Kingdom ([email protected]).
|
astro-ph/9804177 | 2 | 9804 | 1998-06-22T10:00:19 | Assisted inflation | [
"astro-ph",
"gr-qc",
"hep-ph"
] | In inflationary scenarios with more than one scalar field, inflation may proceed even if each of the individual fields has a potential too steep for that field to sustain inflation on its own. We show that scalar fields with exponential potentials evolve so as to act cooperatively to assist inflation, by finding solutions in which the energy densities of the different scalar fields evolve in fixed proportion. Such scaling solutions exist for an arbitrary number of scalar fields, with different slopes for the exponential potentials, and we show that these solutions are the unique late-time attractors for the evolution. We determine the density perturbation spectrum produced by such a period of inflation, and show that with multiple scalar fields the spectrum is closer to the scale-invariant than the spectrum that any of the fields would generate individually. | astro-ph | astro-ph | Assisted inflation
Andrew R. Liddle, Anupam Mazumdar and Franz E. Schunck
Astronomy Centre, University of Sussex, Falmer, Brighton BN1 9QJ, U. K.
(August 4, 2018)
In inflationary scenarios with more than one scalar field, inflation may proceed even if each of the
individual fields has a potential too steep for that field to sustain inflation on its own. We show
that scalar fields with exponential potentials evolve so as to act cooperatively to assist inflation, by
finding solutions in which the energy densities of the different scalar fields evolve in fixed proportion.
Such scaling solutions exist for an arbitrary number of scalar fields, with different slopes for the
exponential potentials, and we show that these solutions are the unique late-time attractors for the
evolution. We determine the density perturbation spectrum produced by such a period of inflation,
and show that with multiple scalar fields the spectrum is closer to the scale-invariant than the
spectrum that any of the fields would generate individually.
PACS numbers: 98.80.Cq
Sussex preprint SUSSEX-AST 98/4-3, astro-ph/9804177
8
9
9
1
n
u
J
2
2
2
v
7
7
1
4
0
8
9
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
I. INTRODUCTION
The idea of cosmological inflation [1,2] is an attrac-
tive one, solving a range of otherwise troubling prob-
lems. Inflation is normally achieved by a period of the
Universe's evolution during which the energy density is
dominated by the potential energy of a scalar field. Al-
though quite probably the early Universe contained sev-
eral scalar fields, it is normally assumed that only one of
these fields remained dynamically significant for a long
time, with the others rapidly finding their way into the
minima of their respective potential energies.
In this paper we consider scalar fields with exponential
potentials. These are already known to have interesting
properties; for example, if one has a universe containing
a perfect fluid and such a scalar field, then for a wide
range of parameters the scalar field 'mimics' the perfect
fluid, adopting its equation of state [3,4]. These scaling
solutions are attractors [5] at late times. The behaviour
of such a field during an inflationary epoch has also been
considered [5].
What was not considered in Ref. [5] is the effect of in-
troducing a scalar field with an exponential potential on
the other scalar field. The simplest example would be if
the other field also possessed an exponential potential.
Then the behaviour of both fields will be modified, since
they feel only their own potential gradient, but experi-
ence, via the expansion, the frictional effect of all scalar
fields present.
where mPl is the Planck mass. Note that there is no
direct coupling of the fields, which influence each other
only via their effect on the expansion. The equations of
motion are
H 2 =
8π
3m2
Pl
φi = −3H φi −
m
Xi=1(cid:20)V (φi) +
dV (φi)
.
dφi
1
2
φ2
i(cid:21) ;
(2)
(3)
Our fields are combined additively; this is different from
soft inflation [6], where an exponential potential multi-
plies the potential of another scalar field.
If there is only a single scalar field, this leads to the
well-known power-law solution [7]
a(t) ∝ tp .
(4)
This is inflationary only if p > 1, i.e. for sufficiently shal-
low exponentials. The power-law solution also applies for
any p in the range 1/3 to 1, where it is non-inflationary.
For p < 1/3, the asymptotic solution is that of a free
scalar field, with a ∝ t1/3 regardless of the value of p in
the range (0, 1/3).
We first find a particular solution where all the scalar
fields are equal: φ1 = φ2 = · · · = φm. We shall later show
it is the unique late-time attractor. With this ansatz the
equations become
H 2 =
8π
3m2
Pl
m (cid:20)V (φ1) +
dV (φ1)
1
2
φ2
1(cid:21) ;
.
(5)
(6)
dφ1
II. DYNAMICS
φ1 = −3H φ1 −
For simplicity, we begin by considering m scalar fields,
φi, which each have an identical potential
V (φi) = V0 exp(cid:18)−r 16π
p
φi
mPl(cid:19) ,
(1)
1
These can be mapped to the equations of a model with
a single scalar field φ by the redefinitions
φ2
1 = m φ2
1
;
V = m V
;
p = mp ,
(7)
so the expansion rate is a ∝ t p, provided that p > 1/3.
The expansion becomes quicker the more scalar fields
there are. And in particular, potentials with p < 1, which
for a single field are unable to support inflation, can do so
as long as there are enough scalar fields to make mp > 1.
Note also that this solution does not require p to exceed
1/3, only the product mp. If mp is less than one third
then the solution will instead be that of a free scalar field.
Although the solution with all scalar fields equal is a
particular one, it is in fact the generic late-time attrac-
tor. To see this, keep φ1 but replace the rest with the
redefined fields
ψi = φi − φ1 ,
i = 2, . . . , m .
These fields obey the equation
ψi + 3H ψi =
p
V0
mPlr 16π
×(cid:20)exp(cid:18)−r 16π
exp(cid:18)−r 16π
mPl(cid:19) − 1(cid:21) ,
ψi
p
p
φ1
mPl(cid:19)
(8)
(9)
i = 2, . . . , m .
This is the equation of a scalar field in an effective po-
tential
ψi + 3H ψi = −
∂Veff(φ1, ψi)
∂ψi
,
(10)
with
p
Veff = V0r 16π
×(cid:20)r p
16π
p
exp(cid:18)−r 16π
exp(cid:18)−r 16π
φ1
mPl(cid:19)
mPl(cid:19) +
ψi
p
ψi
mPl(cid:21) .
(11)
The minimum in the ψi direction is always at ψi = 0,
regardless of the behaviour of φ1, so the late-time solution
has all the φi equal. The length of time to reach this
attractor will depend on the initial separation (the value
of ψi) and the extent to which friction, coming from the
expansion rate H, is important.
III. DENSITY PERTURBATIONS
It is now well known how to calculate the density per-
turbation produced in multi-scalar field models. Sasaki
and Stewart [8] (see also [9]) quote the result
PR =(cid:18) H
2π(cid:19)2 ∂N
∂φi
∂N
∂φj
δij ,
(12)
where PR is the spectrum of the curvature perturbation
R in the usual units [2], N is the number of e-foldings of
inflationary expansion remaining, and there is a summa-
tion over i and j. Since N = −R H dt, we have
φi = −H ,
∂N
∂φi
Xi
(13)
2
where in our case each term in the sum is the same,
yielding
PR =(cid:18) H
2π(cid:19)2 1
m
H 2
φ2
1
.
(14)
Note that this last expression only contains one of the
scalar fields, chosen arbitrarily to be φ1. This expression
looks as if it is m times smaller than the usual formula for
a single scalar field (see e.g. Ref. [2]); however, remember
that the presence of multiple fields has modified both H
and φ1.
Of particular interest is the spectral index n. This is
given by [8]
n − 1 = 2
H
H 2
− 2
∂N
∂φi (cid:16) 8π
m2
Pl
φi φj
H 2 −
∂N
δij
∂φi
Pl
m2
8π
∂N
∂φj
V,i,j
V (cid:17) ∂N
∂φj
, (15)
where there is a summation over repeated indices and
the commas indicate derivatives with respect to the cor-
responding field component. Under our assumptions, the
complicated second term on the right-hand side of the
above equation cancels out, and Eq. (15) reduces to the
simple form
1 − n = −2
H
H 2 =
m2
Pl
8π " ∂V (φ1)
V (φ1)#
∂φ1
2
=
2
mp
.
(16)
This result shows that the spectral index also matches
that produced by a single scalar field with p = mp. The
more scalar fields there are, the closer to scale-invariance
is the spectrum that they produce. Note however that if
the fields have such steep potentials as to be individually
non-inflationary, p < 1, then many fields are needed be-
fore the spectrum is flat enough (say n > 0.7) to have the
possibility of explaining the observed structures. Large
numbers of scalar fields are predicted by some theories,
for example the 70 scalar fields, with unknown poten-
tials, of the low-energy compactified superstring effective
action [10].
IV. POTENTIALS WITH DIFFERENT SLOPES
We now generalize the above discussion, by considering
each potential to have a different slope pi
Vi(φi) = V0 exp(cid:18)−r 16π
pi
φi
mPl(cid:19) .
(17)
Notice that we keep the same V0 for each field; since
changing V0 is equivalent to shifting the scalar field defi-
nition by a constant, this serves to fix the zero values of
the fields.
We conjecture that scaling solutions exist, where the
energy densities of the different fields attain fixed ratios
at late times, relative to an arbitrarily chosen field φ1:
φ2
i
φ2
1
=
Vi(φi)
V1(φ1)
= Ci .
which generalizes Eq. (9) to
(18)
To guess the appropriate form of Ci, we note that the
slow-roll approximation gives
which suggests
φ2
i
≃
2
3pi
,
V 2
i (φi)
Pi Vi(φi)
Ci =
pi
p1
.
(19)
(20)
We stress though that the slow-roll approximation is not
needed in what follows.
Integrating the kinetic part of Eq. (18) then gives
ψi + 3H ψi =
pi
V0
mPlr 16π
p1 (cid:20)exp(cid:18)−r 16π
exp(cid:18)−r 16π
mPl(cid:19) − 1(cid:21) ,
ψi
p1
pi
pi
×
φ1
mPl(cid:19)
(27)
i = 2, . . . , m .
As before, the effective potentials for the ψi fields have
a unique minimum at ψi = 0 for all i = 2, . . . , m. Our
scaling solution is therefore the unique late-time attrac-
tor.∗
The calculation of the spectral index follows the same
lines as before, yielding
1 − n =
2
p
.
(28)
φi =r pi
p1
φ1 + αi ,
(21)
This reduces to Eq. (16) when the slopes of the potentials
are same.
where αi are the integration constants. Ensuring the
potentials also scale as in Eq. (18) requires the constants
to have values
αi = −r pi
16π
mPl ln
pi
p1
.
(22)
To see that this solution will solve the full dynamical
equations, we generalize the scaling argument of Section
II. Eq. (21) reduces us to a single degree of freedom φ1 in
a manner consistent with the equations of motion, which
become
H 2 =
8π
3m2
Pl Pm
i=1 pi
p1
(cid:20)V1(φ1) +
φ1 = −3H φ1 −
dV1(φ1)
.
dφ1
1
2
φ2
1(cid:21) ;
(23)
(24)
Using the scaling of the potential from Eq. (18), the other
scalar wave equations all match the latter of these, con-
firming consistency of the ansatz. Note in particular that
Eq. (21) brings all the exponentials into the same form.
These can then be turned into the equations of a model
with a single scalar field via the redefinitions
φ2
1 = P pi
p1
φ2
1
;
V1 = P pi
p1
V1
;
pi , (25)
p =Xi
of which Eq. (7) is a special case. This result is exact,
not requiring a slow-roll approximation, and once more
shows that the presence of multiple scalar fields increases
the expansion rate. The expansion law is a ∝ t p, and is
valid provided that p > 1/3.
The scaling construction shows that this solution ex-
actly solves the multi-scalar field equations. One can
show that this solution is an attractor by generalizing
the argument of Section II, via the ansatz
V. CONCLUSION
Although the early Universe is likely to contain many
scalar fields, a common assumption when analyzing in-
flation is that all but one of these fields has become dy-
namically irrelevant. However, for scalar fields with ex-
ponential potentials the late-time behaviour is for the
energy densities of the different fields to scale with each
other, as had already been noted for the case of a scalar
field with an exponential potential plus a barotropic fluid
[3 -- 5], even if the fields have no direct coupling to each
other and if their potentials have different slopes.
Such multiple scalar fields can act cooperatively to
drive a period of inflation, even if the individual fields
have potentials which are too steep in their own right;
the expansion law in the scaling solution is t p, where
p =P pi with the pi being the power-law expansion rates
that the individual fields would drive in isolation. The
reason for this behaviour is that while each field experi-
ences the 'downhill' force from its own potential, it feels
the friction from all the scalar fields via their contribution
to the expansion rate.
We have also studied the density perturbation spec-
trum produced, which has a spectral index n matching
that of power-law inflation driven by a single field at rate
p. The spectrum is therefore brought closer to scale-
invariance the more fields participate in the inflationary
expansion. A perturbation spectrum close to scale in-
variance is preferred by current observations, and this
phenomena may offer assistance to supergravity-based
inflation models which often predict spectra which are
not all that close to scale invariance [11].
ψi = φi −r pi
p1
φ1 − αi ,
i = 2, . . . , m .
(26)
∗We have also confirmed these solutions as late-time attrac-
tors numerically for a wide range of values of pi.
3
ACKNOWLEDGMENTS
A.R.L. is supported by the Royal Society, A.M. by the
Inlaks foundation and the ORS, and F.E.S. by the Euro-
pean Union TMR Marie Curie programme. We thank Ed
Copeland, Jim Lidsey and David Wands for discussions,
and acknowledge use of the Starlink computer system at
the University of Sussex.
[1] A. Guth, Phys. Rev. D 23, 347 (1981); E. W. Kolb and
M. S. Turner, The Early Universe, Addison -- Wesley, Red-
wood City (1990).
[2] A. R. Liddle and D. H. Lyth, Phys. Rep. 231, 1 (1993).
[3] C. Wetterich, Nucl. Phys. B302, 668 (1988).
[4] D. Wands, E. J. Copeland and A. R. Liddle, Ann. N. Y.
Acad. Sci. 688, 647 (1993).
[5] E. J. Copeland, A. R. Liddle and D. Wands, Phys. Rev.
D 57, 4686 (1998).
[6] A. L. Berkin, K. Maeda and J. Yokoyama, Phys. Rev.
Lett. 65, 141 (1990); G. Abolghasem, A. Burd, A. Coley
and R. van den Hooden, Phys. Rev. D 48, 557 (1993).
[7] F. Lucchin and S. Matarrese, Phys. Rev. D 32, 1316
(1985).
[8] M. Sasaki and E. D. Stewart, Prog. Theor. Phys. 95, 71
(1996).
[9] A. A. Starobinsky and J. Yokoyama, Report. No. gr-
qc/9502002 (1995); J. Garc`ıa-Bellido and D. Wands,
Phys. Rev. D 53, 5437 (1996).
[10] E. Cremmer and B. Julia, Phys. Lett. B80, 48 (1978);
Nucl. Phys. B156, 141 (1979).
[11] D. H. Lyth, Lancaster preprint hep-ph/9609431.
4
|
astro-ph/9907197 | 1 | 9907 | 1999-07-14T23:30:14 | Warm Gas and Ionizing Photons in the Local Group (ApJL version) | [
"astro-ph"
] | Several lines of argument suggest that a large fraction of the baryons in the universe may be in the form of warm (T\sim 10^5-10^7 K) gas. In particular, loose groups of galaxies may contain substantial reservoirs of such gas. Observations of the cosmic microwave background by COBE place only weak constraints on such an intragroup medium within the Local Group. The idea of a Local Group corona dates back at least forty years (Kahn & Woltjer 1959). Here we show that gas at T\sim 2-3 X 10^6 K (the approximate virial temperature of the Local Group) -- extremely difficult to observe directly -- can in principle radiate a large enough flux of ionizing photons to produce detectable H alpha emission from embedded neutral clouds. However, additional constraints on the corona -- the most stringent being pulsar dispersion measures towards the Magellanic Clouds, and the timing mass -- rule out an intragroup medium whose ionizing flux dominates over the cosmic background or the major Local Group galaxies. A cosmologically significant coronal gas mass could remain invisible to H alpha observations. More massive galaxy groups could contain extensive coronae which are important for the baryon mass and produce a strong, local ionizing flux. | astro-ph | astro-ph |
Warm Gas and Ionizing Photons in the Local Group
Philip R. Maloney1 and J. Bland-Hawthorn2
ABSTRACT
Several lines of argument suggest that a large fraction of the baryons in the universe
may be in the form of warm (T ∼ 105 − 107 K) gas. In particular, loose groups of
galaxies may contain substantial reservoirs of such gas. Observations of the cosmic
microwave background by COBE place only weak constraints on such an intragroup
medium within the Local Group. The idea of a Local Group corona dates back at
least forty years (Kahn & Woltjer 1959). Here we show that gas at T ∼ 2 − 3 × 106
K (the approximate virial temperature of the Local Group) -- extremely difficult to
observe directly -- can in principle radiate a large enough flux of ionizing photons to
produce detectable Hα emission from embedded neutral clouds. However, additional
constraints on the corona -- the most stringent being pulsar dispersion measures
towards the Magellanic Clouds, and the timing mass -- rule out an intragroup medium
whose ionizing flux dominates over the cosmic background or the major Local Group
galaxies. A cosmologically significant coronal gas mass could remain invisible to Hα
observations. More massive galaxy groups could contain extensive coronae which are
important for the baryon mass and produce a strong, local ionizing flux.
Subject headings: Local Group -- cosmic microwave background -- intergalactic medium
-- diffuse radiation
1.
Introduction
The standard Big Bang cosmological model makes remarkably precise predictions for
the abundance of baryons in the universe:
in terms of the critical density parameter Ω, the
prediction is Ωb ≃ (0.068 ± 0.012)h−2, where we take the present-day Hubble constant to be
H0 = 100h km s−1 Mpc−1 (e.g., Olive et al. 1991; Schramm & Turner 1998). An inventory of
baryons observed at high redshift (z ∼ 2 − 3), chiefly in the form of the low-column density
Lyα-forest clouds, gives an estimate of Ωb which, although subject to substantial, systematic
uncertainties, is in reasonable agreement with the standard prediction of Big Bang nucleosynthesis
1Center
for Astrophysics and Space Astronomy, University of Colorado, Boulder, CO 80309-0389;
[email protected]
2Anglo-Australian Observatory, P.O. Box 296, Epping, NSW 2121, Australia; [email protected]
-- 2 --
(see the summary in Fukugita, Hogan, & Peebles 1998). However, as has been noted repeatedly
(e.g., Persic & Salucci 1992; Fukugita et al. 1998), at z ≈ 0 only a small fraction of the expected
number of baryons has been observed, suggesting that there is a substantial, even dominant
reservoir of baryons which has not yet been characterized.
A plausible suggestion for one reservoir of baryons is that loose groups of galaxies contain
substantial masses of warm (T < 107 K) ionized gas, an idea which appears to have originated
with Kahn & Woltjer (1959; see also Oort 1970; Hunt & Sciama 1972). X-ray observations of
poor groups of galaxies frequently detect intragroup gas at T ∼ 1 keV (e.g., Pildis, Bregman, &
Evrard 1995; Mulchaey et al. 1996). In general, only groups dominated by ellipticals are detected;
spiral-rich groups tend to show only emission from individual galaxies. Although this may be due
to the absence of gas in such groups, it is also plausible that the gas has not been seen because
its temperature is too low: the velocity dispersions characterizing groups dominated by spiral
galaxies are significantly smaller than those of compact, elliptical-dominated groups, and imply
temperatures T ∼ 0.2 − 0.3 keV (Mulchaey et al. 1996), making detection even at relatively soft
X-ray wavelengths very difficult.
Most recently, Blitz et al. (1998) have suggested that the majority of high-velocity clouds
(HVCs; for a review, see Wakker & van Woerden 1997) are not associated with the Galactic ISM,
but represent remnants of the formation of the Local Group (LG), as material continues to fall
into the LG potential. In this scenario, some fraction of these infalling clouds will collide in the
vicinity of the LG barycenter and shock up to the virial temperature, T ∼ 2 × 106 K, producing a
warm intragroup medium.
In this paper, we explore the possibility that the Local Group contains such a reservoir of
warm ionized gas. In particular, we examine whether significant constraints can be placed on the
amount of gas through the detection of recombination lines from neutral gas within the Local
Group. In the next section we briefly recapitulate the existing constraints on such an intragroup
medium, and in §3 we estimate the flux of ionizing photons. §4 discusses the implications and
additional constraints which can be imposed, in particular, mass flux due to cooling and the
timing mass of the Local Group.
2. COBE and X-Ray Constraints on Local Group Gas
Suto et al. (1996) suggested that a gaseous LG halo could significantly influence the CMB
quadrupole moment observed by COBE. Assume the Local Group contains an isothermal plasma
at temperature Te whose electron number density is (for core density no and core radius ro)
ne(r) = no
r2
o
r2 + r2
o
cm−3;
(1)
i.e., the nonsingular isothermal sphere. Since we allow ro as well as no to vary, the parameterization
of equation (1) includes density distributions ranging from ne ≈ constant to ne ∝ r−2. As in Suto
-- 3 --
et al. (1996), we calculate the resulting Sunyaev-Zeldovich temperature decrement as a function
of angle, expand in spherical harmonics and average over the sky to obtain the monopole and
quadrupole anisotropies. The COBE FIRAS data (Fixsen et al. 1996) imply that the Compton
y-parameter y = T0,sz/2 < 1.5 × 10−5 (95% CL), which imposes the constraint
noroTkeV < 7.4 × 1021θ−1
o
R
ro (cid:18)
y
1.5 × 10−5(cid:19) cm−2,
where θo ≡ tan−1(R/ro). Similarly, the COBE quadrupole moment requires
noroTkeV < 1.6 × 1020QµK
R
ro "θo − 3(cid:18) ro
R(cid:19) + 3θo(cid:18) ro
R(cid:19)2#−1
(2)
(3)
cm−2
where the rms quadrupole amplitude QRMS = 10−6QµK K; the observed value QµK ≈ 6 (e.g.,
Bennett et al. 1996). Suto et al. (1996) argued that a LG corona which satisfied equation (2) could
significantly affect the quadrupole term, as equation (3) is more restrictive than (2). However,
Banday & Gorski (1996) showed there is no evidence for a LG corona in the COBE DMR skymaps.
In addition, Pildis & McGaugh (1996) pointed out that the typical values of noroTkeV observed
in poor groups of galaxies, resembling the Local Group, are well below the limit (2), generally
no more than a few ×1020 cm−2. Furthermore, spiral-rich groups usually reveal no evidence for
intragroup gas at X-ray energies; Pildis & McGaugh give upper limits of a few ×1019 cm−2 for an
assumed temperature Tkev ∼ 1.
Thus, although the COBE constraints on a LG corona are in fact quite weak3, analogy
with similar poor groups suggests that the LG is unlikely to have a significant gaseous X-ray
corona. However, as noted in §1, the lower temperature expected for the gas in spiral-rich
groups significantly relaxes the X-ray constraints on warm gas in groups similar to the LG. If the
product noroTkeV in a LG corona is typical of that seen in more compact groups, merely at lower
temperature, the mass in baryons can still be very substantial: for the density distribution (1),
scaling noroTkeV to 1020 cm−2, the mass inside radius r is approximately (assuming r/ro
>∼ a few)
M (r) ≈ 7 × 1011(cid:18) ro
100 kpc(cid:19)2(cid:18) r
ro(cid:19)(cid:18) noroTkeV
1020 cm−2(cid:19)(cid:18) TkeV
0.2 (cid:19)−1
M⊙ ;
(4)
this could be a substantial fraction of the mass of the Local Group (see §4).
Direct detection of emission from gas at such temperatures is exceedingly difficult. Using
deep ROSAT observations, Wang & McCray (1993) (WM) find evidence for a diffuse thermal
component with TkeV ∼ 0.2 and ne ∼ 1 × 10−2 x−0.5
kpc cm−3 (assuming primordial gas) where xkpc
is the line-of-sight depth within the emitting gas in kiloparsecs. In the next section we consider
an indirect method of detection: the recombination radiation from neutral gas embedded in the
corona, due to the ionizing photon flux generated by the corona gas.
3There is some confusion in the literature regarding the interpretation of the COBE limits. In evaluating eq. (9)
of Suto et al. (1996), there is no numerical fudge factor suggested by Pildis & McGaugh (1996). Moreover, in Fig. 4
of Suto et al. , the 6µK curve is displaced downwards by a factor of 3.
-- 4 --
3.
Ionizing Photon Flux from a Local Group Corona
We assume the density distribution (1). Approximating the surface of a cloud as a
plane-parallel slab, the normally incident flux on the inner (facing r = 0) cloud face is
φi(r) ≈
πn2
oro
(1 + r2/r2
o)1.5 ξih0.8 + 1.3(r/ro)1.35i phot cm−2 s−1
(5)
where ξi is the frequency-integrated ionizing photon emissivity and the term in brackets is
accurate to 10% for 10−3 ≤ r/ro ≤ 12. (For r/ro
>∼ 2, the flux on the outer face of the cloud is
insignificant.) To calculate ξi, we have used the photoionization/shock code MAPPINGS (kindly
provided by Ralph Sutherland). Models have been calculated for metal abundances Z = 0.01, 0.1,
and 0.3 times solar, and for equilibrium and nonequilibrium ionization. For 104 < T < 107 K,
3 × 10−15 ≤ ξi ≤ 3 × 10−14 phot cm−3 s−1 sr−1. Scaling to physical values,
φi(r) ≈ 104n2
−3r100(cid:18) ξi
10−14(cid:19) (cid:2)0.8 + 1.3(r/ro)1.35(cid:3)
(1 + r2/r2
phot cm−2 s−1
(6)
o)1.5
where the central density no = 10−3n−3 cm−3 and the core radius ro = 100r100 kpc. Poor groups
show a very broad range of core radii, from tens to hundreds of kpc (Mulchaey et al. 1996), and
typical central densities no ∼ a few ×10−3 cm−3 (Pildis & McGaugh 1996).
In Fig. 1, we plot φi as a function of core radius ro for densities no = (1, 3, 10) × 10−3 cm−3,
for a metallicity Z = 0.1 times solar; results differ by <∼ 20% for the other values of Z. The
value of φi is evaluated at r = 350 kpc, the assumed distance rMW of the Galaxy from the center
of the LG (solid lines), and at r = 0 (dashed lines). The fluxes can be very large, exceeding
106 phot cm−2 s−1. However, for ro ≪ rMW, equation (6) shows that the incident flux at rMW is
greatly diminished compared to the peak value of φi.
At distances r ∼ 2ro or less, the ionizing flux produced by a LG corona could be large enough
for detection in Hα: the emission measure is related to the normally incident photon flux by
Em = 1.25 × 10−2(φi/104) cm−6 pc. However, to produce a significant flux, no must be so large
<∼ 109 years. Even though the LG may be,
that the cooling time tc within r ∼ ro is short, tc
dynamically, considerably younger than a Hubble time, such a short cooling timescale makes it
necessary to consider explicitly the fate of cooling gas.
To estimate the mass cooling flux M , we assume that the flow is steady, spherical, and
subsonic, and that any gradients in the potential are small compared to the square of the sound
M = 4πρvr2; v is
speed. In this case the pressure is constant, and mass conservation requires that
the inflow velocity. The cooling radius rc is set by the condition tc ∼ tLG, the Local Group age.
c / M , where ρc is the gas density at rc. We assume
The flow time from rc is tf ∼ rc/v ∼ 4πρcr3
tf ∼ tc, so that the gas has time to cool before reaching r = 0. This sets v ∼ rc/tLG at rc. If
the cooling function Λ does not vary rapidly with T , the density and temperature within rc scale
nearly as ρ ∝ rc/r, T ∝ r/rc (Fabian & Nulsen 1977). We have used these scalings to calculate M
and φi, including the variation of ξi and Λ with radius. Fig. 2 shows several models. The ionizing
-- 5 --
M if ro is large and no is low, but in many cases
M is prohibitively
flux can be large for small
large, ruling out any such coronae. However, there are several important caveats. Unless the LG is
very old, it is unlikely that a steady-state flow has been established (e.g., Tabor & Binney 1993),
especially as infall of gas into the LG is likely to be ongoing. (If a steady-state flow existed with
M , one would expect the line luminosity -- e.g., Hα -- from the cooled gas to be high:
substantial
M is sensitive to the assumed density distribution. For a
see Donahue & Voit 1991.) Furthermore,
given metallicity (and therefore Λ(T )) and LG age, there is a unique value of no at which tc = tLG
and M → 0. As no is raised above this value M increases rapidly, since rc increases and M ∝ r2
c .
M also depends on Z, since the reduced Λ for low Z means
The value of φi at a given value of
that no is larger for a fixed tc. Given these uncertainties, it is not clear that the estimated values
of
M should be regarded as serious constraints.
4. Discussion
The results of the previous section show that a warm Local Group corona could in principle
generate a large enough ionizing photon flux to produce detectable Hα emission from neutral
hydrogen clouds embedded within it. This would offer an indirect probe of gas which is extremely
difficult to observe in emission. Whether the flux seen by clouds at distances comparable to the
offset of the Galaxy from the center of the Local Group is high enough for detection depends to a
large extent on the core radius characterizing the gas distribution, due to the dropoff in flux for r
substantially greater than ro. As shown in Fig. 1, for sufficiently large values of ro and no, φi can
be detectably large even at a few hundred kpc from the LG barycenter.
These large−no, large−ro models run into insurmountable difficulties, however, when we
examine the additional constraints which can presently be imposed on a LG corona. In Fig. 3 we
show, shaded in gray, the range in (ro, no) for which the resulting ionizing photon flux is between
φi = 104 and φi = 105 phot cm−2 s−1, for radial offsets r = 0 (lower region) and r = rMW = 350
kpc (upper region). The cosmic background is probably φi,cos ∼ 104 phot cm−2 s−1 (Maloney &
Bland-Hawthorn 1999: MBH). We also plot the following constraints:
(1) The assumption that any LG intragroup medium is "typical" (Mulchaey et al. 1996; Pildis
<∼ 1.5 × 1021 cm−2, assuming Tkev ∼ 0.2. This is
& McGaugh 1996) constrains the product noro
plotted as the short-dashed line in Fig. 3. Any corona which is not unusually rich must lie to the
left of this line. This restriction alone rules out any significant contribution to φi at rMW.
(2) Assuming that the relative velocity of approach of the Galaxy and M31 is due to their mutual
gravitational attraction, one can estimate the mass MT of the Local Group (Kahn & Woltjer
1959; q.v., Zaritsky 1994). This 'timing mass' depends somewhat on the choice of cosmology; we
take MT = 5 × 1012 M⊙ within r = 1 Mpc of the LG center. The timing mass constraint (using
equation [10]) is shown as the solid line in Fig. 3. As plotted, it is barely more restrictive than the
COBE quadrupole constraint (the long-dashed line), and is only more stringent than restriction
-- 6 --
(1) for large core radii. However, realistically the MT constraint is much more severe, as the Milky
Way and M31 undoubtedly dominate the mass of the Local Group, and so the timing mass curve
in Fig. 3 should be moved downward in density by a factor of at least ∼ 5 − 10.
(3) We possess some information on (more precisely, upper limits to) the actual electron densities
at r ∼ rMW. Constraints on ne(rMW) come from two sources. Observations of dispersion measures
Dm toward pulsars in the LMC and the globular cluster NGC 5024 (Taylor, Manchester & Lyne
1993) require a mean n−3 ∼ 1; this is a slightly weaker constraint than provided by MT . However,
most of this column must be contributed by the Reynolds layer, and some fraction of the Dm
toward the LMC pulsars presumably arises within the LMC, so probably <∼ 10% can be due
to a LG corona. Second, a mean density of no more than n−3 ∼ 0.1 is allowed by models of
the Magellanic Stream; otherwise, the Stream clouds would be plunging nearly radially into the
Galaxy (Moore & Davis 1994). This limits the central density to n−3 ≈ 0.1 + (rMW/ro)2. The
hatched region in Fig. 3 indicates the portion of (ro, no) space in which ne(rMW) ≤ 10−4 cm−3.
(4) As noted earlier (§2), WM found evidence for a thermal soft X-ray component at Tkev ≈ 0.2.
If this emission arises in a LG corona, then the corresponding electron density as derived from
the emission measure Em is ne ∼ 3 × 10−4 x−0.5
Mpc cm−3, where x is the extent of emitting region
along the line of sight; the density would be ∼ 3 times smaller for gas of solar rather than zero
metallicity. This density constraint is comparable to the Dm constraint plotted in Fig. 3.
Some of these constraints can be avoided if the corona gas is clumped. The estimates of mass
(equation[4]) and φi assume a smooth density distribution. However, if the actual densities are
a factor C higher than the mean (smoothed) density at a given radius, φi can be kept constant
while reducing both the gas mass and Dm by 1/C. This is ad hoc, but if the LG halo is being
fueled by ongoing infall, it would not be at all surprising for the gas distribution to be nonuniform.
However, the WM X-ray determination is unaffected by clumping, as it is derived from Em.
The constraints on a LG corona shown in Fig. 3 rule out a significant contribution to the
ionizing flux at r ∼ rMW. If the core density no is high, the core radius ro must be small; conversely,
for large ro, no must be low. LG coronae within the allowed region of parameter space can produce
fluxes φi ≫ φi,cos, but only on scales of a few tens of kpc, at best. Thus the maximum volume
in which a corona ionizing flux exceeds φi,cos is only of order 1% of the LG volume, comparable
to the volume which can be ionized by galaxies (MBH). This has important implications for the
model of Blitz et al. (1998), in which most HVCs are remnants of the formation of the Local
Group. If HVCs are at megaparsec distances, φi will be dominated by the cosmic background. The
resulting emission measures will be small: barring unusually favorable geometries, the expected
Hα surface brightnesses ( <∼ 10 mR) are at the limit of detectability. Any HVCs which are truly
extragalactic and detectable in Hα would need to lie close to the dominant spiral galaxies (within
their "ionization cones": Bland-Hawthorn & Maloney 1999a,b) or the LG barycenter.
In summary, a warm LG corona which significantly dominates the UV emission within the
Local Group is ruled out, although such a corona could contain a cosmologically significant
-- 7 --
quantity of baryons. More massive galaxy groups could well contain coronae that are both
cosmologically important and dominate over the ionizing background. Such coronae could have
major impact on the group galaxies through ionization and ram pressure stripping4. Finally, we
note that, four decades later, the observational limits on a LG corona have yet to improve on the
values suggested by Kahn & Woltjer (1959).
PRM is supported by the Astrophysical Theory Program under NASA grant NAG5-4061.
Bandy, A.J., & G´orski, K.M. 1996, MNRAS, 283, L21
REFERENCES
Bennett C.L., Banday A.J., G´orski K.M., Hinshaw, G., Jackson, P., Keegstra, P., Kogut, A.,
Smoot, G.F., Wilkinson, D.T., & Wright, E.L. 1996, ApJ, 464, L1
Bland-Hawthorn, J., & Maloney, P.R. 1999a, ApJ, 510, L33
Bland-Hawthorn, J., & Maloney, P.R. 1999b, in The Stromlo Workshop on High Velocity Clouds,
ed. B.K. Gibson & M.E. Putnam (San Francisco: ASP), 212.
Blitz, L., Spergel, D.N., Teuben, P.J., Hartmann, D., & Burton, W.B. 1998, ApJ, 514, 818
Donahue, M., & Voit, G.M. 1991, ApJ, 381, 361
Fabian, A.C., & Nulsen, P.E.J. 1977, MNRAS, 180, 479
Fixsen, D.J., Cheng, E.S., Gales, J.M., Mather, J.C., Shafer, R.A., & Wright, E.L. 1996, ApJ, 473,
576
Fukugita, M., Hogan, C.J., & Peebles, P.J.E. 1998, ApJ, 503, 518
Hunt, R., & Sciama, D.W. 1972, MNRAS, 157, 335
Kahn, F.D., & Woltjer, L. 1959, ApJ, 130, 705
Maloney, P.R., & Bland-Hawthorn, J. 1999, in The Stromlo Workshop on High Velocity Clouds,
ed. B.K. Gibson & M.E. Putnam (San Francisco: ASP), 199.
Moore, B., & Davis, M. 1994, MNRAS, 270, 209
Mulchaey, J.S., Davis, D.S., Mushotzky, R.F., & Burstein, D. 1996, ApJ, 456, 80
4We note that, in principle, observations of the O VI doublet at 1032 and 1038 A are extremely sensitive to the
presence of such a corona: for the maximum allowed coronae of Fig. 3, the expected line fluxes could be as large as
F ∼ a few × 10−10 erg cm−2 s−1. However, the observational difficulties (absorption and scattering of the photons
within the ISM of the Galaxy and the very large spatial extent of the source for a LG corona) are severe.
-- 8 --
Olive, K.A.; Steigman, G., & Walker, T.P. 1991, ApJ, 380, L1
Oort, J.H. 1970, A& A, 7, 381
Persic, M., Salucci, P. 1992, MNRAS, 258, 14P
Pildis, R.A., & McGaugh, S.S. 1996, ApJ, 470, L77
Pildis, R.A., Bregman, J.N., & Evrard, A.E. 1995, ApJ, 443, 514
Schramm, D. N., & Turner, M. S. 1998, Rev. Mod. Phys., 70, 303
Suto, Y., Makishima, K., Ishisaki, Y., & Ogasaka, Y. 1996, ApJ, 461, L33
Tabor, G., & Binney, J. 1993, MNRAS, 263, 323
Taylor, J.H., Manchester, R.N., & Lyne, A.G. 1993, ApJS, 88, 529
Wakker, B.P., & van Woerden, H. 1997, ARA&A, 35, 217
Wang, Q.D., & McCray, R.M. 1993, ApJ, 409, L37
Zaritsky, D. 1994, in The Local Group: Comparative and Global Properties, 3rd CTIO/ESO
Workshop, ed. A. Layden, R.C. Smith, & J. Storm (Garching: ESO), p. 187
This preprint was prepared with the AAS LATEX macros v4.0.
-- 9 --
Fig. 1. -- Normally incident ionizing photon fluxes φi from a Local Group corona, for core densities
n0 = (1, 3, 10) × 10−3 cm−3 (bottom to top), as a function of core radius ro. The solid lines are
for r = 350 kpc from the center of the corona, and the dashed lines are for r = 0. A metallicity
Z = 0.1Z⊙ has been assumed.
Fig. 2. -- Ionizing photon fluxes φi versus the mass cooling rate M , for a steady-state cooling flow.
A core radius ro = 50 kpc has been assumed. From top to bottom, the curves are for assumed
corona ages tLG = 1, 2, 4, and 8 Gy.
Fig. 3. -- Constraints on a Local Group corona in the (ro, no) plane. Coronae within the gray-
shaded regions produce ionizing photon fluxes between φi = 105 and 104 phot cm−2 s−1 (upper and
lower edges) at radii r = 0 (lower region) and r = 350 kpc (upper region) with respect to the LG
center. The long-dashed line is the COBE quadrupole constraint, the short-dashed line assumes
the LG medium is "typical", the solid line is the timing mass constraint, and the hatched region
satisfies ne ≤ 10−4 cm−3 at r = 350 kpc. See §4 for discussion.
-- 10 --
Fig. 1. --
-- 11 --
Fig. 2. --
-- 12 --
Fig. 3. --
|
astro-ph/0701608 | 2 | 0701 | 2007-07-12T13:11:17 | Could we identify hot Ocean-Planets with CoRoT, Kepler and Doppler velocimetry? | [
"astro-ph"
] | Planets less massive than about 10 MEarth are expected to have no massive H-He atmosphere and a cometary composition (50% rocks, 50% water, by mass) provided they formed beyond the snowline of protoplanetary disks. Due to inward migration, such planets could be found at any distance between their formation site and the star. If migration stops within the habitable zone, this will produce a new kind of planets, called Ocean-Planets. Ocean-planets typically consist in a silicate core, surrounded by a thick ice mantle, itself covered by a 100 km deep ocean. The existence of ocean-planets raises important astrobiological questions: Can life originate on such body, in the absence of continent and ocean-silicate interfaces? What would be the nature of the atmosphere and the geochemical cycles ?
In this work, we address the fate of Hot Ocean-Planets produced when migration ends at a closer distance. In this case the liquid/gas interface can disappear, and the hot H2O envelope is made of a supercritical fluid. Although we do not expect these bodies to harbor life, their detection and identification as water-rich planets would give us insight as to the abundance of hot and, by extrapolation, cool Ocean-Planets. | astro-ph | astro-ph | I09772: rev ised vers ion, accepted by Icarus
Could we identify hot Ocean-Planets with CoRoT, Kepler
and Doppler velocimetry?
--------
version 21 11 Jul. 07,
F. Selsis (1), B. Chazelas (2), P. Bordé(3,*), M. Ollivier (2), F. Brachet(2), M. Decaudin(2), F.
Bouchy(4), D. Ehrenreich(4), J.-M. Griessmeier(5), H. Lammer(6), C. Sotin (7), O. Grasset(7), C.
Moutou(8), P. Barge(8), M. Deleuil(8), D. Mawet(9), D. Despois (10), J. F. Kasting(11), A. Léger(2)
(1) Ecole Normale Supér ieure de Lyon, Cen tre de Recherche As tronomique de Lyon , 46 allée d'Italie,
F-69364 Lyon Cedex 07 , France ; CNRS, UMR 5574 ; Université de Lyon 1, Lyon , France.,
franck.selsis@ens-lyon .fr
(2) Institu t d ’As trophysique. Spa tia le, bat 121 , Université Paris-Sud and CNRS (UMR 8617); Un iv.
Paris-Sud, F-91405 Orsay; Fr, bruno [email protected]; Marc [email protected] .fr;
Frank.Brache t@ias .u-psud .fr ; miche l.decaudin@ias .u-psud.fr ; Alain [email protected]
(3) Harvard-Smithson ian Center for Astrophysics, 60 Garden Stree t, Cambridge , MA 02138 , USA,
[email protected] .edu
(*) Michelson Postdoctoral Fellow. Curren t address : Michelson Science Cen ter, Ca lifornia Institu te of
Technology, 770 S W ilson Avenue, MS 100-22, Pasadena, CA 91125 , USA
(4) Ins titut d’Astrophysique de Paris ; CNRS (UMR 7095) ; Université Pierre & Marie Cur ie ; 98, bis
boulevard Arago, F-75014 Paris, France ; bouchy@iap .fr; ehrenrei@iap .fr ;
(5) LESIA, CNRS-Observatoire de Paris , 92195 Meudon, France ,
jean-math [email protected]
(6) Space Research Institu te , Aus trian Academy of Sciences, Schmiedlstr . 6, A-8042, Graz , Austria ,
helmu t.lammer@oeaw .ac .at
(7) Géophysique , Un iversité de Nantes, F-44321 Nantes cedex 3 , Fr, sotin@chimie .univ-nantes .fr ;
Olivier .Grasset@chimie .univ-nantes .fr
(8) Labora toire d’Astrophysique de Marseille (LAM/OAMP), CNRS, – BP 8 - Traverse du Siphon ,
13376 Marseille Cedex 12 , C laire .Moutou@oamp .fr ; p [email protected]; magali.deleuil@oamp .fr
(9) université de Liège , 17 allée du 6 Août, 4000 Sart-Tilman , Belguim, dimitri.mawet@ulg .ac.be
(10) Observatoire de Bordeaux ( INSU/CNRS) , B.P. 89 , F-33270 Flo irac, Fr, D idier [email protected]
bordeaux1.fr
(11) Dept. of Geosciences , The Pennsylvania State University, University Park , Pennsylvania 16802,
USA, [email protected] .edu
ABSTRACT
Planets less massive than abou t 10 MEarth are expec ted to have no massive H-He atmosphere and a
cometary composition (~ 50% rocks, 50% wa ter, by mass) provided they formed beyond the
snowline o f protoplanetary disks. Due to inward migration , such planets cou ld be found a t any
distance between their forma tion site and the star . If migra tion s tops within the habitable zone, this
may produce a new kind of p lane ts, ca lled Ocean-Planets . Ocean-p lane ts typica lly consist in a
silicate core , surrounded by a thick ice mantle, itse lf covered by a 100 km – deep ocean . The
possible exis tence of ocean-plane ts raises impor tan t astrobiological questions: Can life origina te on
such body, in the absence of continen t and ocean-silicate inter faces? Wha t would be the nature of
the atmosphere and the geochemica l cycles ?
In th is work, we address the fa te of Ho t Ocean-Plane ts produced when migra tion ends at a closer
distance . In this case the liqu id/gas inter face can disappear, and the hot H2O envelope is made o f a
supercritical fluid . Although we do no t expect these bodies to harbor life, the ir de tection and
identifica tion as water-rich planets wou ld give us insight as to the abundance o f hot and, by
extrapola tion, cool Ocean-Planets .
The wa ter reservoir of these planets seems to be weakly affected by gravita tiona l escape , provided
that they are loca ted beyond some minimum dis tance, e .g. 0 .04 AU for a 5-Earth-mass plane t
around a Sun-like star . The swelling o f their wa ter atmospheres by the h igh stellar flux is expected
not to sign ificantly increase the planets’ rad ii. We have studied the possibility o f detec ting and
characterizing these Hot Ocean-Planets by measur ing their mean densities using transit missions in
space – CoRoT (CNES) and Kepler (NASA) – in combina tion with Dopp ler velocime try from the
ground – HARPS (ESO) and possible future ins truments. We have de termined the domain in the
[stellar magnitude , orbital dis tance ] plane where discr imination be tween Ocean-Plane ts and rocky
plane ts is possible with these instrumen ts.
The brigh test s tars of the mission target lists and the plane ts closest to their stars are the most
favorable cases. Full advan tage o f h igh precision photometry by CoRoT, and par ticularly Kepler,
can be ob tained on ly if a new genera tion of Doppler ins truments is built.
JID:YICARAID:8261/FLA[m5+;v1.73;Prn:12/06/2007;8:22]P.2(1-16)2F.Selsisetal./Icarus•••(••••)•••–•••1.Introduction1.1.Water-richvssilicate-richplanetsOcean-planets(OPs)representanewkindofplanets,re-centlyproposedbyLégeretal.(2004,2003)andKuchner(2003),whichnaturecanbedescribedasfollows:(1)Theirformationtakesplacebeyondthesnowline,giv-ingthemacometary-likebulkcomposition:silicatesandwater,inroughlyequalamountsbymass.Planetsthataccretedplanetesimalsformedatlessthanabout3AUfromasolartypestarhaveamuchlowerwatercontent(w≡H2O/silicate,bymass):wis5×10−4fortheEarthandremainsbelow10%foraplanetentirelymadeofcarbonaceous-chondriticmaterial.Raymondetal.(2006b,2007)simulatedtheformationofhabitableplanetsintheabsenceofmigrationandfoundw<5%attheendoftheaccretion.Bysimulatingplanetaryformationafterthemi-grationofanoutergiantplanettoaclose-inorbit,Raymondetal.(2006a)predictedthatthemixingofmaterialfromtheinnerandouterregionscanproduceplanetsinthehabit-ablezoneconsistingofupto∼30%water.ThismightbeanotherwaytoformOPs.(2)UnlikeUranusandNeptune,OPsdonotreachasufficientmasstoaccreteagaseousenvelopedirectlyfromthepro-toplanetarydisk.Theyarethuslessmassivethanabout10MEarthandassumedtohavenoH2–Heenvelope.Thisas-sumptionisrobustforplanetsof6MEarthandbelow,andhastobediscussedfurtherforplanetsintherange6–10MEarthforwhichsignificantgasaccretionmaystillhaveoccurred(Alibertetal.,2006;Rafikov,2006).(3)TobecomeanOP,therocky–icyplanethastomigratetotheinnerpartoftheplanetarysystemthroughtype-Imigration(Papaloizouetal.,2006).Ifmigrationstopsinthehabit-ablezone(HZ)ofitsstar,theplanetbecomeswhatLégeretal.(2004,2003)callanocean-planet(OP).Whenmigra-tionstopsatshorterorbitaldistances,itcangivebirthtoplanetswithathickandhotH2Oenvelopewithnoliquid–atmosphereinterface.(4)PlanetsofafewMEarthareexpectedtobemainlymadeofsilicateandwater,exceptwhenfoundaroundstarswithaC/Oratiosignificantlyhigherthansolar,wherecarbon-planetsmayform(KuchnerandSeager,2005).Althoughoneshouldkeepinmindthispossibilityforsomeplane-tarysystems,weconsiderinthispaperthatplanetscanbecharacterizedwithasinglenumber:thewatercontentw.Inthiscase,internalstructuremodelsprovidedifferentplan-etaryradii,foragivenmass,asafunctionofthiswaterratio.Fora6-Earth-massplanet,Légeretal.(2004,2003)foundRpl=2.00REarthforanOPandRpl=1.63REarthforarockyone.Sotinetal.(2007)generalizedthesecalcula-tionsformassesinthe(0.01–10MEarth)rangeandvariablewatercontent(Fig.1).Independently,Valenciaetal.(2006)alsoderivedmass–radiuscurvesforplanetswithvariouswatercontents.Thesetwoworksareinverygoodagree-Fig.1.Relationbetweenthemassandtheradiusofocean-planets(OPs)(m50%rocks,50%water)uppercurve,androckyplanets(m100%rocks)lowercurve,asobtainedfrommodelsoftheirinternalstructurebySotinetal.(2007)whenplanetsarelocatedinthehabitablezoneoftheirstar,orfurther.Foragivenmass,themeandensitiesandtheradiioftheplanetsaresignificantlydifferentaccordingtotheirnature.Planetarymassandradiusareaccessibletoobserva-tionalmeasurementsbyDopplervelocimetryandplanetarytransitphotometry,respectively.Thesemeasurementshavethepotentialcapabilityofdiscrimi-natingbetweenthetwotypesofplanets.ItshouldbenotedthatsomeoftheGalileansatellitesofJupiterareexamplesofOPs,butfortheirlowermasses.ment.Bothfoundthat,inallcases,acoupleofvalues(Mpl,Rpl)areindicativeofaplanetarycomposition.1.2.Onthevergeofdetectingocean-planets?Thankstoremarkableprogress,radialvelocity(RV)mea-surementsarenowunveilingapopulationofplanetswithamassbetween10and20MEarththatarefoundatshortorbitaldistancesandasfarasthehabitablezoneoftheirstar.ThisiswellillustratedbythetriplesystemHD69830a,b,c(Lovisetal.,2006).Modelsreproducingthissysteminvoketheformationoficy–rockycoresbeyondthesnowline,followedbyinwardmi-gration(Alibertetal.,2006;TerquemandPapaloizou,2006).Becauseoftheirhighermass,thesehotNeptunesdifferfromOPsbytheirH–HeenvelopeofseveralMEarth.Coresabove10MEarthareindeedexpectedtoaccreteasignificantfractionoftheirmassashydrogen-richgas.Fortworeasons,thesediscoveriesareextremelypromisingforthesearchofOPs:First,thesamemechanismthatcom-monlyproduceshotNeptunesshouldalsogenerateOPsforlessmassivemigratingrocky–icycores.Second,theincreasedaccu-racyinRVmeasurementsshouldsoonallowustodetectplanetsbelow10MEarth,intherangeofmassesofOPs.TheCoRoTmission(http://CoRoT.oamp.fr;Rouanetal.,1999)wassuccessfullylaunchedbytheendof2006andKeplerisscheduledforlaunchin2008(http://kepler.nasa.gov;Kochetal.,2006).Theirtransitsearchprogramswilldeterminetheradiiofdiscoveredplanets.TheDopplerfollow-up,wheneverJID:YICARAID:8261/FLA[m5+;v1.73;Prn:12/06/2007;8:22]P.3(1-16)Identificationofhotocean-planets3Fig.2.ProbabilityofobservinganexistingplanetaroundatargetstarbyCoRoT,asafunctionofitsorbitalperiod.Assumingarandomorientationofplanetaryorbits,thegeometricalprobabilitythatatransitisobservedfromtheEarthispg=Rst/a,theratioofthestellarradiustothatoftheplanetdistancetoitsstar,inthelimitRpl%Rst.UsingtheKeplerianrelationbetweenaandtheplanetaryperiodTorb(foragivenstellarmass,T2orb/a3=const)andthetimeprobabilitythat!3eventsoccurwithina150dayduration,p3,oneob-tainstheprobabilitythatagivenplanetaroundaGVstarisdetected.Thisisarapidlydecreasingfunctionoftheplanetaryperiod,whichpointsoutthatplanetdetectionisbiasedtowardsshortperiodobjects.possible,willdeterminetheirmasses.Foreachdetectedplanet,theseobservationswilldetermineapointanditsboxerrorinthe(Mpl,Rpl)plane.ModelsoftheirinternalstructurepredictRpl(Mpl)curvesinthatplane,whichvarywiththeplanetarycomposition(Fig.1).Assumingthatthemodelsarecorrect,therelativepositionsoftheerrorboxandthecurveswilllead,ornot,tothecharacterizationofthenatureoftheplanet.ThisapproachwassuccessfullyappliedtotransitinghotJupiters,amongwhichwecandistinguishacore-dominatedplanet(HD149026b;Satoetal.,2005)fromtheothergas-dominatedplan-ets(forinstanceTrEs-1;Alonsoetal.,2004).Inreturn,theobservationshaveputseriousconstraintsonthemodelsandthestructureofhotJupiters,orPegasides.Forinstance,thelargeradiusofHD209458bisnotyetwellexplained.1.3.PlanetorbitalperiodsaccessibletoCoRoTandKeplerCoRoTwasinitiatedasaCNES“smallmission.”Itislo-catedinalow-altitudeEarthorbit(∼900km).Asitmustnotpointtoofarfromtheanti-solardirection,itcanobserveagivenstellarfieldcontinuouslyonlyduring5months(150days).Itisconsideredthatthedetectionofaplanetarycandidatebythetransitmethodrequirestheobservationof3transitsormore.Thus,onlyplanetswithperiodP<75dayscanbedetected(Fig.2).Forasolar-typestar,thiscorrespondstoadistancetoitsstar,a,lessthan0.35AU(circularorbit)andablackbodytemperature(albedo=0,nogreenhouseeffect)Tbb>460K.PlanetswithP<75dayscanbehabitableiftheirhost-starislessmassivethanaK5star(M∼0.7MSun).However,theselow-massstarsrepresentonly1.5%ofthestarsaccessiblewithCoRoT.Withlessthan200oftheminitsfieldandatransitprobabilityof∼1%,CoRoTisunlikelytodetectthetransitofahabitableplanet.Statistically,theprobabilitythat3transitsormoreofagivenplanetaredetectedwithin150daysisarapidlydecreasingfunc-tionofitsorbitalperiod(Fig.2).Ifplanetswereuniformlydistributedindistancesaroundtheirstars,andifCoRoThadthesensitivitytodetectallofthem,thehistogramofdetectionswouldbeproportionaltothisprobability,stressingthestrongbiastowardsshortperiods.Inaddition,thelargerthenumberoftransitsthelargerthedetectionS/Nforsmall-sizeplanets.Asaconsequence,itisimportanttoaddressthefateofOPswhentheyareclosetotheirstars.ThesituationfortheKeplermissionismorefavorable(http://kepler.nasa.gov;Kochetal.,2006).Itisalargermis-sionthatwilloperateonanEarth-trailingheliocentricorbit,continuouslyfor4years.Itcandetectplanetswithperiod"1.33yraroundsolar-typestars,correspondingtosemimajoraxisa"1.21AUandTbb!250K,whichincludesEarth-likeplanets,themaingoalofthemission.However,evenforKepler,thedetectionofinnerplanetswillbeeasierandmoreaccurate.AccordingtoBordéetal.(2003),CoRoTcoulddetectsev-eraltensofplanetswithRpl∼2REarthanda×(L/LSun)1/2<0.35AU,whereListhestellarluminosity,ifeachstarhasoneplanetwithintheserangesofsizeandlocation.Thishypothesisisarbitrary,butitindicatesthattheactualnumberofdetectionscouldtellustheabundanceoftheseplanetsifthemissionca-pacitiesareasexpected.ForKepler,theprospectsarehigherandstartfromplanetswitha"1.21AUforSun-likestars.1.4.OutlineofthisstudyDuetotheobservationmethodsdescribedabove,exoplanetswiththeshortestperiodsaretheeasiesttocharacterizebytheirmassandradius.Ontheotherhand,andasalreadydiscussedbyKuchner(2003),thestrongstellarirradiationcanaffecttheevo-lutionofthewaterreservoirofclose-inplanets(byatmosphericescape)andoftheirradius.Insummary,OPsthathavekeptenoughwatertobedistinguishedfromsilicate-dominatedplan-etsmaybefoundonlyatorbitalperiodthatwillnotbeavailablesoontoaccuratemassandradiusmeasurements.Thisquestionprovidesthegeneraloutlineofourpaper.InSection2,weaddressthefateofwater-richplanetsthathavemigratedveryclosetotheirstar,theirswellingduetothestrongexternalheating,and,especially,thelossoftheirwa-terthroughthermalandnon-thermalevaporationinducedbystellarXandEUVradiation,windandcoronalmassejections(CMEs).InSection3,weestimatetheaccuracyoftheradiusandmassdetermination,asafunctionoftheorbitalperiodandstellarproperties,thatcanbeobtainedwithCoRoT,Kepler,andRVfollowup.Then,onthebasisofthisaccuracyandofthetheo-reticalRpl(Mpl,w)curves,weevaluatethecapabilitytodistin-guishawater-richfromasilicate-richpopulationofplanet.InSection4,wediscusswhattherealisticdiversityof1–10MEarthshort-periodplanetscouldbe.Inparticularweaddressthepossibledegeneracyof(Rpl,Mpl)couplesofvaluesduetotheexistenceofanaccretedatmosphereofH2–He,andtotheJID:YICARAID:8261/FLA[m5+;v1.73;Prn:12/06/2007;8:22]P.4(1-16)4F.Selsisetal./Icarus•••(••••)•••–•••existenceofplanetswithadifferentbulkcompositionsuchascarbon-planets(KuchnerandSeager,2005).OurconclusionsarepresentedinSection5.2.Structureandfateofashortperiodocean-planetOPsaredefinedasplanetsthatformfartherawaythanthesnowlineandmigrateinwards.Whentheorbitaldistancesta-bilizeswithinthehabitablezone,andassumingthattheplanethascooledtothethermalequilibriumwiththestellarirradi-ation,aglobalandthickoceanwouldcondenseaboveanicymantle(Légeretal.,2004).Forshorterorbitaldistancesnogas/liquidinterfacecanbesustained.Assumingthatwateristhemainconstituentoftheatmosphere,thishappenswhenthereisnomorephasetransitionbetweenthevaporphaseandtheliquidone,i.e.whenthetemperaturedistributionwithintheplanetT(P)runsabovethecriticalpointofwater(Tc=647K,Pc=22MPa∼220bars)asillustratedinFig.3a.Whenanoceanispresent,theT(P)curvecrossesthegas–liquidtransi-tionline,whichdeterminestheoceansurfacelocation(Fig.3b).Intheabsenceofliquid–gasinterface,thewholefluidlayercanbecalledenvelopeasingiantplanets.Thewholewaterenve-lopeissupercritical.Thesehotplanetscouldbenamed“planetswithsupercriticalwaterenvelope”or“sauna-planets.”Forthesakeofsimplicity,weshallkeepthenameofhotocean-planets.ThepossibilitythattransitmissionsdetectandcharacterizesomeOPsdependsontheirexistence/abundanceandthein-strumentcapabilitiestodetectsmallplanets,saywithRpl∼2REarth,atsmalldistancesfromtheirstars,asdiscussedinSec-tion3.Inthepresentsection,wediscusswhethertheRpl(Mpl)relationestablishedbySotinetal.(2007)(Fig.1)intheHZismodifiedbyaswellingoftheplanetduetostrongerirradiation(Section2.1)andhowclosetotheirstarOPscanbewithoutlosingmostoftheirwater(Section2.2).2.1.Estimatedswellingofanocean-planetclosetoitsstarSotinetal.(2007)andValenciaetal.(2006)modeledtheinternalstructureofOPsandestablishedtheRpl(Mpl)relationforOPsintheHZ.Whentheseplanetsareclosertotheirstars,thehigherstellarfluxinducesahighertemperatureattheouterboundary,whichmayresultinanenhancedradius.Inordertotestthesensitivityoftheradiustothisouterboundarytem-perature,Sotinetal.(2007)havecalculatedthat,inthecaseofa5MEarthOP,anincreaseof1000Kresultsina0.9%in-creaseoftheradius.However,thiscomputationwasdoneonlyforthecondensed(solidandliquid)partoftheplanet.Amoresignificantswellingisexpectedfromtheexpansionofthehotwatervaporatmosphere.InsidetheHZ,theatmosphericwatervaporcontentisafunctionoftheorbitaldistance:aroundthepresentSun,intheabsenceofothergreenhousegasesandas-sumingalargeoceanatthesurface,thepartialpressureofH2Owouldbe1barat0.93AUandwouldreachthecriticalpressure(220bar)at0.84AU(Kasting,1988).Atcloserdistances,theouterpartoftheplanetisathicksupercriticalwaterenvelope,or“steam-ocean,”boundedbythehigh-pressureicemantle,orbythesilicatemantleifthetemperatureimposesthatallthewa-terisfluid.Thedepthofanhot“steam-ocean”wouldbelargerthanthedepthofliquid–solidwatermantleofthesamemass,whilethesubcriticalatmosphereaboveP=220barexpandsquasi-linearlywithitstemperatureandmightthusresultinaswelling.Letusfirstconsidertheswellingduetotheexpansionoftheopaqueatmosphere.ThepressureintheatmospheredependsFig.3.WaterphasediagramandqualitativeT(P)profilesfortheouterlayersofawater-richplanet.Planet(a)isinthehabitablezone(HZ)ofitsstarandhasanocean–atmosphereinterface.Planet(b)isatanorbitaldistanceslightlyshorterthantheinneredgeoftheHZandplanet(c)hasaveryhotequilibriumtemperature(>500K).Planets(b)and(c)havenoocean–atmosphereinterfacebutathickfluidenvelopeofsupercriticalH2O.JID:YICARAID:8261/FLA[m5+;v1.73;Prn:12/06/2007;8:22]P.5(1-16)Identificationofhotocean-planets5uponthealtitude,z,as(1)P=P0e−z/H,H=CkT/µg,whereHistheatmosphericheightscale,P0thecriticalpres-sureatthelowerboundary,ktheBoltzmannconstant,µthemeanmolecularweight,gthelocalgravity,andCisthecom-pressibilityfactor,whichdependsonthepressureandtempera-ture,C=1foraperfectgasandC<1foradensegas.Theradiusdeterminedbyatransitobservationisthedis-tancez,totheplanetarycenter,ofalightraygrazingtheplanetaryatmospherewithanopticalthicknessofunityattherelevantwavelength.Wavelength-dependentradiioftransitingEarth-sizeplanetshavebeencalculatedbyEhrenreichetal.(2006).Theypresenttransmittedspectraoftransitingocean-planetswithorwithoutclouds.Forahotocean-planet,watercloudsarenotexpectedbecausethegasisatatemperature!1000K(seebelow),whichisabovethecriticalpointofwaterandpreventsthecondensationofliquiddroplets.Assumingthattherearenootheraerosolsorgrains(Feorsilicates),intheup-peratmosphere,Rayleighscatteringisexpectedtobethemaincontributortothetotalextinction.Anapproximationofthecol-umndensityNacrosstheatmosphericlimbalongthelineofsightisN=n(2πRplH)1/2,wheren=P/(kT)isthevolumicdensity(Fortney,2005).ThecorrespondingopticaldepthhasbeencalculatedbyEhrenreichetal.(2006)andcanexpressedas(2)τ(λ)=√2πP(z)kTR1/2plH1/2σRayl(l),whereσRaylistheRayleighcross-sectionofthegasmolecules(herewater).Requiringthatτ=1at0.6µm,acentralwave-lengthofthephotometricmeasurements,impliesavalueforPandthenzfromEq.(1).Normalizingtoareferencecase,usingσRayl(H2O,0.6µm)=2.32×10−27cm2,onereads:(3)P=0.38¯T1/2¯M1/2¯R−3/2bar,with¯T=T/103K,¯M=Mpl/6MEarth,¯R=Rpl/6REarth.Forthereferencecase,onefindsz−z0=13kmwherezandz0arethealtitudeswhereτ=1at0.6µm(P=0.38bar),andP0=1bar,respectively.Theeffectivelengthleffofthelighttravelintheatmosphereisthen1000km,correspondingtoasustainingangle2θ=4.3◦.Forcomparison,foratransitingEarth,usingσRayl(air,0.6µm)=3.16×10−27cm2andne-glectingtheabsorptionsduetoO3andtheaerosols,onegets:P=0.24barandz−z0=11.4km,leff=640km,2θ=5.8◦.Thesevalueshavetobecomparedwiththemuchhigherattenu-ationoftheSunatsunsetwhenthevalueofPis1barandonlyhalfofthegrazingtravelisperformedbythelight.InthefollowingestimatesoftheswellingofahotOP,weconsiderthattheplanetaryradiusdeterminedbyatransitobser-vationcorrespondstoP∼0.4bar,andwedeterminetheheightoftheatmosphericlayerboundedbythe0.4and220barlevels.AttheorbitaldistanceofahotJupiter(0.05AU),thetem-peratureofthe0.4barlevelofanOPwouldbeclosetoitsequilibriumtemperature,thatis1100K.InahotJupiterofwithasimilarequilibriumtemperature,thetransitionbetweentheouterradiativelayerandtheconvectiveregionoccursintheH2–Heenvelopeatpressuresandtemperaturesoftheorderof500–1000bar,2000–3000K(Barmanetal.,2005).Althoughwewouldneedaself-consistentmodeltocomputetheT–Pprofileofawaterenvelope,asafunctionoftheorbitaldistanceandtakingintoaccountthepreciseopacitiesofH2O,wecanassumethattheatmospherictemperatureincreasesfromabout1100Kintheupperandoptically-thinlayers,totypically2000–2500Katthetransitionbetweentheradiativeandconvectiveregion,thatwearbitrarilyfixat200barshere.Byestimatingtheheightofa1000and2500Kisothermalatmosphereathy-drostaticequilibrium,wecanbrackettheheightoftheradiativelayerthatisopaquetothetransitobserver.Fora6MEarthplanetandthesetwoatmospherictemperatures,the0.4barlevelisreached,respectively,at260and700kmabovethe220barlevel.Theseestimatestakeintoaccountthenon-idealproper-tiesofthehighpressurewatervapor(NISTsteamtablesforthecompressibilityfactor),butonecannotethatusingtheperfectgasapproximationyieldssimilarvalues(210and700km).Theintrinsicswellingoftheconvectiveenvelopeunderneaththeradiativeatmosphereislowerbutitaffectsalargefractionoftheradius.Theproblemtoestimatethedepthofthislayerasafunctionofthetemperatureoftheouterlayersisdouble.Firstwearerapidlylimitedbythelackofavailabledatafortheequa-tionofstateofwaterathightemperatureandpressure.Then,asthecoolingoftheplanetsisdeterminedbythegradientoftem-peratureintheouterlayers,ahighequilibriumtemperatureoftheplanetcanleadtoasignificantlydecreasedheatfluxandveryhotinterior.Addressingthisissuewouldrequireaconsis-tentformationmodelprovidinginitialthermalconditionsandanevolutioncode.Verypreliminaryestimateusingextrapola-tionsofavailabledatashowthatthedepthoftheconvectiveenvelopeofanhot6MEarthcanincreaseby∼300–500km.Comparetothe2REarth(12,800km)radiusitwouldhaveintheHZ,ahot6MEarthOP(at0.05UAfromitsstar)wouldexhibitaradiusincreasedby5–10%:+200–700kmfortheatmosphere,+300–500kmforthefluidwaterenvelope,+<100kmforthesilicate+metalinterior.Forsmallerplan-etsandplanetsstillclosertotheirstars,theswellingcanbesomewhatlarger.For1MEarth,at0.05AU,theswellingshouldbe>10%.Rockyplanetscouldalsoexperiencesuchaswellingoftheiratmosphere,inprinciplemakingthemdifficulttodistin-guishfromOPs.However,attheorbitaldistanceswheretheswellingcouldbesignificant,silicate-dominatedplanetswouldloserapidlytheirvolatilecontentsthroughthermalandnon-thermalatmosphericlossprocesses,whileOPshaveanalmostinexhaustiblereservoirofwater(seebelow).Hence,whenanOPisclosetoitsstar,itmayundergosomeswellingthatwouldhelptodiscriminateitfromarockyplanet.However,astheseestimatesfortheswellingareveryprelimi-naryandconcernonlyvery-short-periodplanetswewillassumeconservativelythat,ifitsurvives,anOPwillhavearadiussim-ilartowhatitwouldhaveifitwerelocatedwithintheHZ.Duringtheplanetformationprocess,somehydrogenmaybeaccreted,andwaterislikelytobeincontactwithreducingFe2+JID:YICARAID:8261/FLA[m5+;v1.73;Prn:12/06/2007;8:22]P.6(1-16)6F.Selsisetal./Icarus•••(••••)•••–•••ions,possiblyevenmetalliciron;hence,onecanexpectachem-icalreactionwithahydrogenoutput.ThepresenceofH2wouldimplyalargerswellingoftheplanet.However,theatmosphericerosionprocessesthatareconsideredinthenextsectionsaremuchmoreefficientforhydrogenthanfortheothergasesandwillrapidlytakehydrogenawayfromtheatmosphere,espe-ciallyattheorbitaldistancesrelevantforCoRoT.Thereafter,weconsiderthatOP’satmospheresaremainlyfreeofhydrogen,buttheactualhydrogencontentofOPsmeritsfurtherinvestiga-tions.2.2.Evaporationofshortperiodocean-planetsOcean-planetsclosetotheirstarsmaylosetheirwatertospaceandbecomerockyplanets.Theaimofthepresentsec-tionistoestimatehowclosetotheirstarswecanexpectOPstokeeptheirwaterandpossiblybeidentifiedbyCoRoTandKepler.TheprecisemodelingoftheerosionofanOPiscom-plex,andthecorrespondingworkisstilltobedone.Wecan,however,estimatetheupperlimitonatmosphericlossesbasedonmaximumenergydepositionintheupperatmosphere.Usingenergy-limitedescape,whichislikelytobesignificantlylargerthanactuallossrates,wecandeterminetheminimumperiod(ororbitaldistance)atwhichmostofthewaterreservoirshouldsurviveduringagiventime,forinstance5Gyrs.DeterminingthisminimumorbitaldistanceisimportantinthecontextofCoRotsearchesthatwillfocusonclose-inplanets.Note,how-ever,thatplanetsfoundatshorterperiodscouldstillhavekeptmostoftheirwater,dependingonwhetheractuallossesarelim-itedbymechanismsotherthanenergydeposition.Themainprocessesforatmosphericescapethatcanleadeventuallytotheexhaustoftheplanetarywaterreservoirare:(1)thermalescapedrivenbyexosphereheatingbyX-raysandExtremeUV(orXUV:0.1–100nm),(2)non-thermalescapedrivenbytheactionofparticlesfromthestar.2.2.1.WatererosionbythermalescapeTheheatingoftheupperatmospherebyXUVphotonscanresultintheescapeofgasestospace.EscapeaffectsmainlylightspecieslikeHandHe,butheavierspecieslikeOcanalsobecarriedawaybythehydrodynamicflowwhenitishighenough(Chassefiére,1996):suchaneffecthasbeenobservedinthecaseofthehotJupiterHD209458b(at0.05AUfromitsstar)forwhichnotonlyH,butalsoCandO,havebeende-tectedintheescapingupperatmosphere(Vidal-Madjaretal.,2003,2004).Anupperlimittothemasslossisgivenbytheenergy-limitedescaperate(Lammeretal.,2003),whichwouldbereachedifalltheenergyabsorbedatλ<100nmwouldbeconvertedintogravitationalenergy,asexpressedinthefollow-ingequation(4)εFXUVπR2pl(a/1AU)2=GMplmH2ORpl.ThelefttermisthefractionoftheXUVfluxthatisinterceptedbytheplanetandavailablefordrivingthethermalescape,andtherighttermisthevariationofgravitationalenergyin-ducedbythemass-lossofwater.FXUVistheXUVenergyflux[Wattm−2]receivedat1AU,M,Randaarethemass,radiusandorbitaldistanceoftheplanet,respectively;thefactorεcor-respondstotheefficiencyoftheconversionofincidentXUVenergyintoeffectiveescapeofgas,andcontainsallthephys-icalcomplexityoftheescapeprocess.Themassofwaterlostoveratimetis(5)mH2O=επR3plGMpl(a/1AU)2t!0FXUV(t()dt(.Wecanconsiderεastheproductofaheatingefficiencyεh,whichgivesthefractionoftheincidentenergythatisnotre-radiatedtospace,andanescapeefficiencyεesc,whichisthefractionofthedepositedenergythatiseventuallylostthroughescapinggas.Theheatingefficiency,εh,istypicallylowerthan0.2,especiallyathighXUVirradiation.Yelle(2004)usedadetailedphotochemicalmodelandfoundavalueof∼0.1forHD209458b.Here,weuseavalueof0.2inordertoobtainanupperlimitonthelossrate.Intheprecedingrelations,RplissometimesreplacedbytheradiusatwhichtheXUVradiationisabsorbedRXUV(e.g.Lammeretal.,2003).ThiscomesfromWatsonetal.(1981)hydrodynamicmodelingoftheescapeinwhichalltheincom-ingXUVradiationisassumedtobeabsorbedinaninfinitelythinlayeratRXUV.Inthisapproach,RXUVcanbeaslargeasseveraltimesRpl,forterrestrialplanets.Thisexpansionoftheupperlayercanimplyextremelyhighescaperates.Recenthydrodynamicmodels(steadystate:Tianetal.,2005;timede-pendent:Penzetal.,2006)includingarealisticdepositionoftheXUValongtheescapingflow,showthatthethinabsorb-inglayerapproachoverestimatestheescaperatebyordersofmagnitude,especiallyforhighincomingXUVflux.OneofthereasonsforthisoverestimationistheXUVself-shielding:someXUVphotonsthatareabsorbedwithintheescapingoutflowcontributetoheatingtheoutflowbutnottoenhancingthelossrate.WhentheXUVdepositionisjusthighenoughtodrivehy-drodynamicescape,theescapingoutflowisnotdenseenoughtoabsorbasignificantfractionoftheincomingXUVenergy,whichismostlydepositeddeeperwherepartofitiseventuallyconvertedintogravitationalenergy.Inthisphase,theescaperateincreasesnearlylinearlywiththeXUVflux,butwithin-creasingirradiation,escapereachesapointatwhichthecolumndensityofescapingatomscannolongerbeconsideredtranspar-enttoXUV:theoutflowbecomeshotterbuttheefficiencyεescdecreases.TheescapeefficiencywasestimatedbyPenzetal.(2006)bymodelingthehydrodynamicflow.ForHD209458btheyfoundanescapeefficiencyofabout0.5at0.045AUfromthepresentSun(conditionsassumedforHD209458b)thatde-creasesrapidlyforincreasingXUVfluxes(εesc=0.15iftheXUVfluxismultipliedby2).Therefore,thetotalefficiencyε=εhεescvariesfrom0.02to0.1intherangeoforbitaldis-tancesthatareobservablebyCoRoT.Fig.4givesthetimerequiredtolosethewholewaterreser-voirofanOPthroughthisenergy-limitedmass-lossasafunc-tionoftheorbitaldistancea(solidlines).ThisestimatetakesintoaccounttheevolutionoftheXUVstellarluminosityac-JID:YICARAID:8261/FLA[m5+;v1.73;Prn:12/06/2007;8:22]P.7(1-16)Identificationofhotocean-planets7Fig.4.Minimumlifetimeofthewatercontentofa10,5and1Earth-massOPasafunctionofitsorbitaldistance.Solidlinesgivethelifetimeofthewaterreservoirerodedbythermalescapeatitsenergy-limitedrate(determinedbythedepositionofstellarXUVradiation).Dashedlinesgivetheenergy-limitedlifetimefornon-thermalescapeinducedbythestellarwind(determinedherebythedepositionofprotons’kineticenergy).SeeSection2.2fortheassumptionsontheefficiencyofthethermalescape,onstellarXUVandwindproperties.Arounda5Gyroldstar,OPswouldbeweaklyaffectedatorbitaldistances!0.03AU(10MEarth),0.04AU(5MEarth)and0.15AU(1MEarth).cordingtoRibasetal.(2005)andthevariationofεwiththeXUVfluxfromPenzetal.(2006).ForOPswithamassbelow6Earthmasses,thelossislimitedbytheXUVenergydeposi-tion.Fora5Earth-massOP,Fig.4showsthatthethermallossofthewaterreservoir(HandO)wouldtakemorethan5Gyrsata!0.04AU.2.2.2.WatererosionbythestellarwindandcoronalmassejectionsNow,thequestioniswhethertheerosionbythestellarwindcansignificantlychangethelowerlimitofthewaterreservoirlifetime.Thisprocessisefficientwhentheupperatmosphereofaplanetisnotprotectedbyanintrinsicmagneticfield,whichisstrongenoughtodeflectthestellarplasmaflowatlargedis-tancesabovetheexobase.OPsclosebytheirstarsareexpectedtohavetheirspinrota-tiontidallylockedontheirorbitalrotation.Fromsimplethe-oreticalmodels,severalanalyticalscalinglawswerederivedwhichallowonetoestimatetheplanetarymagneticdipolemo-mentfromtheplanetarycharacteristics(e.g.density,rotationrate,sizeofthedynamoregion).ThesemodelsaresummarizedandcomparedinGriessmeieretal.(2004,2005).Accordingtothesemodels,theexpectedmagneticmomentsofslowlyrotat-ingexoplanets,e.g.tidallylockedplanets,aresmallerthanthatoffastrotatingplanetliketheEarth.Byapplyingthesescal-ingrelationstoanOPof6.0Earthmassesand2.0Earthradii(Légeretal.,2004),whichistidallylockedat0.2AUaroundastarwith0.2solarmasses,weobtainamaximumintrinsicmag-neticmomentof0.7timesthatofthepresentEarth.Becauseoftherelativelyweakmagneticmoment(consideringthelargeplanetaryradius),theplanetarymagnetospheremaynotbeeffi-cientatprotectingtheplanetaryatmosphere.InsuchaVenus-likeinteraction(e.g.Teradaetal.,2002)theplanetaryneutralgasintheupperatmospherecanbeionizedbyelectronimpact,X-raysandExtremeUVradiations.Af-terchargeexchangeduringcollisionswiththeincidentstellarplasma,itispickedupbythisstellarplasmaflowandescapesfromtheplanet.Inthetwonextsectionsweestimatethenon-thermalat-mosphericlossduetotheinteractionwiththemeanstellarwindandwiththesporadiccoronalmassejections(CMEs),inte-gratedoverthewholehistoryofthesystem.StellarwindAsanupperlimittotheerosionduetothestel-larwind,wecalculatedthefluxofprotonsinterceptedbytheplanetintheabsenceofmagneticdeflectionandthecorrespond-ingdepositionofkineticenergyintotheupperatmosphere.Byanalogywiththeenergy-limitedthermallossratedescribedinSection2.2.1,wecaninferanenergy-limitedescapeinducedbythestellarwindprotons.Themaindifficultywiththisap-proachisthepoorlyknownevolutionofthestellarwind.Usingindirectobservations,Woodetal.(2002,2005)showedthatthemasslossofSun-likestarsiscorrelatedwiththeirX-raylu-minosityforagesabove0.7Gyr.Youngerstarsdonotseemtoexhibitahighermassloss.Inordertomakeanupperesti-mateofthepossiblelosses,weusedconstantwindpropertiescorrespondingthehighestmasslossmeasuredbyWoodetal.(2005).Thisallowsustooverestimatetheenergydeposition.Animprovementofthisassumptionwouldrequiremoredataonstellarwinds,especiallyonthoseofyoungstars.DashedlinesonFig.4showtheresultwefindfortheminimumlifetimeofJID:YICARAID:8261/FLA[m5+;v1.73;Prn:12/06/2007;8:22]P.8(1-16)8F.Selsisetal./Icarus•••(••••)•••–•••thewaterreservoir.Atadistancea!0.04AU,a5Earth-massOPwouldkeepitswaterformorethan5Gyrsifstellarwinderosionisconsideredalone.Thisisanupperlimitforthenon-thermallossthroughmean-stellar-windprotons,buttheeffectofcoronalmassejections(CMEs)fromthestarhasalsotobeaddressed.CoronalmassejectionsByusingtheparametersestimatedbyKhodachenkoetal.(2007)oneobtainsat0.05AU,anaver-ageCMEplasmavelocityof∼500kms−1andplasma(proton)densities"5×104cm−3.Lammeretal.(2006)studiedmaximumratesofO+ionpickupbyCMEformagnetizedandnon-magnetizedEarth-likeexoplanets.Theycalculatedtheplasmaflowaroundthemagne-topause/ionopauseforXUVfluxvalues10–100timeshigherthanthatofthepresentSun,asexpectedforayoungstar(Ribasetal.,2005).ThetotalionpickuplossratewasobtainedbycalculatingtheO+productionratecausedbychargeexchange,electronimpactandphotonsalongthestreamlinesintheplasmaflowaroundtheplanetaryobstacle.TheyfoundthatforstrongCMEsandfortheestimatedmin-imumplanetarymagneticmoment,thesituationwouldresultinaVenus-likeplasmainteraction(e.g.Teradaetal.,2002).PuttinginthetypicalnumbersfortheCMEprotondensityandtheaveragewindspeedmentionedabove,wefindthat,duringthecriticalfirst0.5GyrforanOPat0.05AU,thewaterlossescouldbeupto"200Earthoceans(bymass),whichissignifi-cantlylessthanthelossesbystellarwindsandcanbeneglected.Oneshouldnotethattheseestimatesarealsoupperlim-itsbecausethemoremassiveOPconsideredinourstudyhasahighergravitationalpotentialthantheEarth-massplanetsstudiedbytheprecedingauthors,whichwillholdbackthethermosphere–exosphereenvironmentexpansion,resultinginlowerlossrates.Furthermore,theevaporatinghydrogenwillformacoronaaroundtheplanetthatalsodecreasesthenon-thermallossofoxygenandotherheavyspeciesbyprotectingthem.Thoseupperestimatesofnon-thermallossprocessesfromweaklymagnetizedOPsdonotcomeclosetoexhaustingtheimportantwaterreservoirofanOPthatislocatedfartherthan0.05AU.Forinstance,200Earthoceans(200×1.45×1021kg)representonly1.5%ofthemassofa6Earth-massOP.2.2.3.Howclosetoitsstarcanweexpecttofindanocean-planet?AccordingtoFig.4,non-thermalescapeinducedbythestel-larwindseemstodominatetheatmosphericlossforMpl"5MEarth.Forlargermasses,thesurvivalofthewaterreservoirislongerthan5Gyrs.Intheabsenceofmoredetailedmodelsforthestellarwind-inducedlosses,ourassumptions(100%ef-ficiency,veryintensestellarwind,nointrinsicmagneticfield)certainlyoverestimatesignificantlythesolarwind-inducedloss(Section2.2.2).Withtheseovermaximizingassumptions,find-ingnon-thermallossesmuchhigherthanthethermalescapewouldhavepreventedusfromprovidingafirmconclusiononthesurvivalofOPs.Fortunately,wefindanupperlimitonthesolarwindinducedlossthatislower,oronlyslightlyabove,thethermalloss.Wecanthusderiveanestimatefortheclosestdis-tanceatwhichOPswillnotbeaffectedbywaterlosstospace:asatypicalresult,a6MEarthOPwillkeepmostofitswaterwhenitisfartherthan0.04AUfromitsstar.ThisconclusionisverysimilartotheresultsobtainedbyKuchner(2003),whoalreadystudiedthesurvivalofthevolatilecontentofshortperiodOPs:onFig.4ofhispaperonecanseethatthelifetimeof5MEarthvolatile-richplanetat0.04AUis5Gyr.ButKuchneralsopointedout2potentialproblemsaf-fectinghiscalculationsattheshortestorbitalperiods.ThefirstoneisrelatedtotheRXUV/Rplratio(discussedinSection2.2.1)thatstronglyoverestimatesthelossrateathighXUVirradiationinWatson’snumericalapproachofthehydro-dynamicescape.ToestimatetheefficiencythelossathighXUVirradiation,weusedrecenthydrodynamiccalculations(Penzetal.,2006)andfoundlossesmuchbelowwhatisgivenbyWat-son’sscheme.Itshould,however,benotedthatthevalueofthisefficiencyhasastrongimpactonthedeterminationofthemini-mumdistance,andthathydrodynamicmodelingofatmosphericescapeisafieldinprogress.Differentestimatesofthiseffi-ciencymightbeobtainedinthefuture.AsecondpointraisedbyKuchneristhattheatmosphericlosscaninduceanexpansionofthewholegaseousenvelope.Becauseofpositivefeedbackonthelossrate,thiseffectcouldleadtoarunawayescapeandtothelossofthewholevolatilereservoirwithinamuchshortertimescalethanwhatisfoundherebyassumingnocouplingbetweenthehydrostaticstruc-tureandthemassloss.ThisrunawaylosswasstudiedforgasgiantsbyBaraffeetal.(2004)andwasshowntooccurwhenthetimescaleofthelossm/mbecomesshorterthanthether-mal(Kelvin–Helmholtz)timescalerequiredforthermalread-justmentoftheenvelopetoamodificationofitsmass.Thisconditioncanbewritten:(6)mm<Gm2RplLpl.Here,misthemassofthewaterenvelopeandLplistheplanetluminosity,whichcanbewrittenas4πR2plσTeffbyassumingameanbrightnesstemperatureTeff,or2πR2plσTeffifthereisnoredistributionoftheincomingenergybetweenthedayandnighthemispheres.Therunawayconditioncanbewritten(7)2σT4effεFXUV<1orTeff<"εFXUV2σ#1/4,whereεistheconversionefficiency.Fortheclosestorbitaldis-tancesweconsidered,thisconditionfortherunawayimpliesTeff<240K.Attheseshortorbitaldistances,realisticeffectivetemperaturescannotdropbelow500K,andwecanassumethatthermalreadjustmentareinstantaneouscomparedtothelosstimescale,whichiscompatiblewithourquasi-static/hydrostaticapproach.Weconcludethattherunawaylossofwaterisnotex-pected.However,westressagainthecentralinfluenceoftheefficiencyε.HigherefficienciesthantheoneobtainedbyPenzetal.(2006)couldtriggerarunawaylossandthelossofthewholewaterreservoirwithinshorttimescales,asdescribedinBaraffeetal.(2004).JID:YICARAID:8261/FLA[m5+;v1.73;Prn:12/06/2007;8:22]P.9(1-16)Identificationofhotocean-planets93.CouldRplandMplmeasurementsbeaccurateenoughtodiscriminateocean-planetsfromrockyplanets?Inordertoknowwhetheranexoplanetisanocean-planetorarockyone,itsmassandradiusmustbemeasuredwithhighenoughaccuracysothatthepositionoftheerrorboxinthe(Mpl,Rpl)planeallowsadistinctionbetweenthepredictionsforthosetwotypesofplanets,assumingthatthemodelisadequate(Fig.1).Forsmallmassesandradii(Mpl<10MEarth,Rpl"2REarth),thisisanotatrivialtask.Hereafter,weestimatethedifferentuncertaintiesthatdefinetheerrorboxes.3.1.Whatarethedifferentsourcesofuncertainty?Aroundagiven(Mpl,Rpl)point,thephysicalquantitythattheplanetarymodelspredicttobedifferentisthemeanden-sity.Thisdensity,)ρ*=(3/(4π))MplR−3plcanbeobtainedbyproperobservations.TheplanetarymassresultsfromDopplervelocimetry.Theplanetaryradiusisobtainedfromthetargetstarradiusestimateandtherelativestellarfluxdropduringthetransit,i.e.Rpl=Rst(F/F)1/2,ifthestellarlimbdarkeningcanbeneglected.Theuncertaintiesonthemass,thestellarradiusestimateandfluxmeasurementsbeingindependent,theirvariancesaddinquadrature.Weestimatetheuncertaintyonthe1/3powerofthedensity,becausethisquantityhasarelativeuncertaintythatissmallerthanthatonthedensity,whichismoreappropriateforlinearexpansions(Protanov,2002):(8)"σρ1/3ρ1/3#2="σMpl3Mpl#2+"σRstRst#2+"σF/F2F/F#2.ConsideringthecaseMpl=6MEarththatcorrespondstoabigtelluricplanetbutthatisstillfarfromthelimitwherethehy-drogengasislikelytohavebeenaccreted,theinternalstructuremodelingbyLégeretal.(2004)yieldsradiiofR1=2.0REarthandR2=1.63REarthforocean-planetsandrockyplanets,re-spectively.Therelativedifferenceforthecorrespondingρ1/3is:(9)ρ1/3ρ1/3=RplRpl=2.0−1.631.8=20.6%.AssumingaGaussiandistributionoferrors,a2σcharacteriza-tionofρ1/3/ρ1/3wouldresultina95%confidencediscrimi-nationbetweenthetwokindsofplanets.Therefore,weconsiderthatthestandarddeviationshouldsatisfy:(10)σρ1/3ρ1/3"10.3%.Next,thedifferentcontributionsinrelation(8)areestimated.3.2.UncertaintyontheplanetarymassThesemi-amplitudeoftheradialvelocitywobbleduetothepresenceofaplanetaroundasolartypestaris(11)K=0.09"MplMEarth#"MstMSun#−1/2"a1AU#−1/2ms−1.Thecorrespondinguncertaintyinthemassis(12)σMplMpl=$"σKK#2+"σMst2Mst#2%1/2.Theuncertaintyinthestellarmassesofthetargetsstarsthatareconsidered(mv<13)canbeestimatedfromthatinthestellarradius,σRst/Rst∼5%(Section3.3),andM(R)empir-icalrelation(Cox,2000):log(R)=0.92log(M)+const.Theresultisanambiguityof∼6%.Thisisaconservativevaluebecause,inpractice,themassisdeducedfromobservablequan-titiesusingatmosphericmodels,whereastoobtaintheradiusanadditionaloperationisneededthatimpliesstellarevolutionmodeling.ThisresultisinagreementwithCharbonneauetal.(2006),whoestimatethattheuncertaintyonthestellarmasscanbeashighas5%.Anyhow,thistermhasanegligiblecon-tributiontotheplanetarymassuncertaintybecauseitleadstoσMst/2Mst"3%,whereasσK/K∼20%(seebelow),andthehierarchy(σMst/2Mst)2%(σK/K)2alwaysstands.OneoftheverybestDopplerinstrumentspresentlyavail-ableistheHARPSspectrometerattheESO-LaSilla3.6mtelescope.Forfaintstars(mv>10),itsuncertaintyisclosetothefundamentalshotnoiselimit.Itwillbethemostaccurateinstrumentforthefollow-upofplanetarytransitcandidatesde-tectedbyCoRoT.Inthatprogram,theDopplermeasurementswillbeperformedinfavorableconditionsbecausetheplanetaryephemeriswillbedeterminedbeforehandbytransitphotome-try(transitperiodandepochs).Theonlyunknownswillbethesemi-amplitudeofthesinusoidalradialvelocitycurveKandthecenter-of-massvelocityv0,ifacircularorbitoftheplanetisassumed.TheshotnoiseuncertaintyofHARPSdependsnotonlyuponthestellarmagnitudeandtheintegrationtime,butalsouponthespectraltypeandtherotationalbroadeningvsinioftheobservedstar(Bouchyetal.,2001).Forthemostappro-priatespectraltypes(fromG5toK5)andstarswithlowrotation(vsini<2km/s),theexpectedshotnoiseuncertaintyonindi-vidualradialvelocitymeasurements,σV,isgivenbyMayoretal.(2003)andPepeetal.(2005):mV=14→σV=6m/s(1h),mV=12→σV=2m/s(1h).Forsuchstars,theuncertaintiesduetoshotnoisearelargerthaninstrumentalsystematicerrors(estimatedto0.8m/sonHARPS).Wesupposeherenon-activestarswithsurfaceveloc-itiesatthelevelof2m/srms,whichisthecaseofthemoststablesolar-likestarsmonitoredwithHARPS.FromσVandthenumberofobservations,theerrorσKonKcanbecomputedasfollows.Incaseofacircularorbit,theradialvelocityreads(13)v(t)=v0+Kcos"2πtTorb+ϕ#,wheretheorbitalperiodTorbandthephaseϕareknown,thankstotransitphotometry.Therefore,onlytwoparameters,thesys-temicvelocityv0andthesemi-amplitudeK,shouldbedeter-minedfrommeasuredvelocities.Least-squaresfittingwouldbethestandardprocedureforthatpurpose.JID:YICARAID:8261/FLA[m5+;v1.73;Prn:12/06/2007;8:22]P.10(1-16)10F.Selsisetal./Icarus•••(••••)•••–•••FollowingtheproofinSection15.2ofPressetal.(2002),andassumingthatallmeasureshavethesameerrorσv,onecanshowthatthevarianceofKis(14)σ2K=σ3v&Ni=1cos2’2πtiTorb+ϕ(−1N’&Ni=1cos’2πtiTorb+ϕ((2.Ifthemeasurementsareuniformlydistributedalongtheorbitalperiod,thefirstsumcanbeapproximatedbyNtimesthein-tegralofasquaredcosineoveritsperiod,i.e.,N/2,andthesecondsumcanbeapproximatedbythemeanvalueofacosineoveritsperiod,i.e.,zero.Therefore,(15)σ2K≈σ2vN/2.However,ifwetakefulladvantageoftheknowledgeofthephaseφandperformthemeasurementsinequalnumbersatthemaximaandminimaofthevelocity,thenthefirstsumisequaltoN(becausethesquaredcosineisalwaysequalto1),whilethesecondsumisstillzero.Hence(16)σ2K≈σ2vN.Inpractice,itmightbedifficulttotakemeasurementsonlyattheextremaofthevelocitycurve.ArealcaseiswellillustratedbyFig.2inO’Donovanetal.(2006),whoperformedtheirmea-surementsclosetotheextrema.Apossiblevalueis(17)σ2K≈σ2vN/1.2.Weassumethatforanimportantproject,amaximumintegra-tiontimecanbeabout40h,correspondingtoacumulativetimeofabout5nights.InordertoexplicitlyestimatethedependenceofσVuponthedifferentparameters(mv,t,Φtel),wecompute(σV)0foragivensetofparametersandforareferencecase,andweexpressσVasafunctionofitandtheparameters.Intheshotnoise-limitedregime,theuncertaintyontheradialvelocityreads(18)σv(σv)0=100.2(mv−mv0)"tt0#−1/2"ΦtelΦ0#−1,whereΦtelisthetelescopediameter,tthecumulativeintegra-tiontime,andtheindex“0”correspondstothereferencecase.Thereferencecasethatweconsideris:Mpl,0=6MEarth,a0=0.10AU,Mst=MSun.Relation(11)givesK0=1.7m/s.Anmv=12starobservedwithHARPSatthe3.6mESOtele-scopeduringt0=40h,leadsto(σK)0=0.35m/s(Eq.(17))and(1/3)(σK/K)0=0.068.Usingrelations(18),theuncer-taintyontheplanetarymass(relation(12))reads(19)13σMplMpl=6.9%A"a0.10AU#1/2100.2(m−12),withA="Mpl6MEarth#−1"MstMSun#1/2"t40h#−1/2"Φtel3.6m#−1.Fig.5.RelativeerroronthestellarradiusplottedagainstthemagnitudeoftheCoRoTtargets.ThedotsshowtheerrorestimatedfromtheVandKactualmagnitudesfromrelation(22)inKervellaetal.(2004),whentheuncertaintyonthereddening,ordistance,isadded.Thefulllineshowsthefittothedata.3.3.UncertaintyonthestellarradiusThestellarradiusofCoRoTtargetswillbeestimatedbyin-depthfollow-upobservationsoftheparentstar,includinghigh-resolutionhigh-SNRspectroscopy.Afirstideais,however,obtainedfromcalibratedrelationshipsofcolorindiceswithstel-larangulardiameters(Kervellaetal.,2004).Apparentmagni-tudesintheVandKfiltersaremeasuredwithaprecisionoftypically0.03magnitudesintheCoRoT/exoplanetstellarcata-logue.Apreliminaryspectralclassificationisthenderivedfrombroadbandspectrophotometryinthevisibleandnear-infrared,withanestimationofthereddening(Moutouetal.,inprepa-ration).High-resolutionspectroscopywillthenallowamoreaccurateestimationoftheeffectivetemperature,witherrorbarsoftypically50to100K(2–3%ofthetemperature).Thiswillfurtherconstrainthereddening,attheleveloftypicallyanad-ditional0.05magnitudeuncertainty,whichincludestheerrorondistanceestimation.ThefinalerroronthelinearradiusoftheCoRoTtargetsisthenobtainedbycombininganerrorontheangulardiameter(derivingEq.(22)inKervellaetal.,2004)withanerroronthedistance(fromspectraltype,reddening,ab-soluteluminosity).Fig.5showsthecalculatedfinalerroronthestellarradiusforCoRoTtargetsinthemagnituderange12–14.Notethatothermethodsofradiusestimationgiveresultsthatareconsistentwithin5%(DiBenedetto,2005).Forcomparison,thestellarradiusofstarsinknowntran-sitingsystemsisestimatedwitherrorsoftypically5–7%forOGLEsystems(e.g.Santosetal.,2006).TheexpectationsforCoRoTtargetsisoftheorderofwhatisachievedonOGLEtar-getsofsimilarmagnitudes,asathoroughfollow-upobservationprogramisforeseentoconstrainatbestthestellarparameters.WeconcludethatthestandarddeviationonRstisestimatedas(20)σRst/Rst∼5%forstarsbrighterthanmv∼13,whichimpliesaverycarefulphotometricandspectroscopicanalysisofthetarget,asinter-JID:YICARAID:8261/FLA[m5+;v1.73;Prn:12/06/2007;8:22]P.11(1-16)Identificationofhotocean-planets11stellarextinctionwillbepresentinthelineofsightofCoRoTtargets.3.4.UncertaintyontheamplitudeofthetransitTheamplitudeoftherelativefluxvariationduringatransit,F/F,is(Rpl/Rst)2,or:(21)FF=0.84×10−4"RplREarth#2"RstarRSun#−2.ForCoRoT,theuncertaintyonthetransitdepthmeasurementisexpectedtobe1.2timestheshotnoiseforstarswithmv=11–14(Fridlundetal.,2006).Theshotnoisedependsontheinstrument,thestellarmagnitude,thenumberkoftransitsandtheirindividualdurationtr.Foratotaldurationobservation,trun,anorbitalperiodTorb,andanimpactparameter(theratioofthedistanceoftheprojectedplanettrajectorytothestellarcenterandthestellarradius)of0.5,thecumulativetransitingtime,ttot,is(22)ttot=ktr,withtr=3.56(a/0.10AU)1/2(Rst/RSun)(Mst/MSun)−1/2h,Torb=11.5(a/0.10AU)3/2(Mst/MSun)−1/2days,k=trun/Torb.Fisobtainedbymeasuringthedepthofthetransitinthestellarlightcurve,F=Ftr−Fbg.Thebackgroundterm,Fbg,canbeaccuratelyestimatedbyaveragingoverlongdurations;therefore,weconsiderthattheshotnoiseonFismainlythatonFtr.Itisdeterminedbycalculatingthenumberofphotoelec-tronsatthefluxlevelFtr∼F(notatthelevelF)duringthetotaltransitingtime.TheuncertaintyonF,innumberofpho-toelectrons,is(23)σF=1.2(Fttot)1/2.TherelativeuncertaintyonFbeingmuchweakerthanthatonF,aftersomealgebra,onereads(24)σF/FF/F≈1.2(F/F)(Fttot)1/2.CoRoThasa0.27mentrancepupilandobservesduring150-dayruns.Itproducesa2.6×104ph-els−1fluxfora12thmagnitudestar(http://CoRoT.oamp.fr).Consideringasaref-erencecase,thetransitofaRpl=2REarthplanetinacircularorbitaroundasolartypestar(F/F=3.36×10−4),atdis-tance0.10AUobservedwithCoRoT,theuncertaintyis(25)"12σF/FF/F#0=2.7%.Keplerhasa0.95mentrancepupilandobservesduring4years.Fora6.5hdurationtransitanda12thmagnitudestar,thedifferentialphotometryaccuracyisestimatedat20ppm,andismainlyshotnoiselimitedforfainterstars(D.Koch,per-sonalcommunication).Forourplanet-starreferencecase,forthewholemissionduration,onefinds(26)"12σF/FF/F#Kepler=0.25%.Ingeneral,foramainlyshotnoiselimitedobservation,theun-certaintyis(27)12σF/FF/F=2.7%B"a0.10AU#1/2100.2(mv−12)withB="Rpl2REarth#−2"RstRSun#3/2"trun150days#−1/2"Φtel0.27m#−1.Itisremarkablethatthedependencesofbothuncertaintiesontheplanetarymass(Eq.(19))andtransitamplitude(Eq.(27))upontheorbitalradius,a,andstellarmagnitude,mv,arethesame.3.5.Capabilitiestoidentifythenatureofthetelluric-likeplanetsThedifferenttermsinEq.(8)arenowestimated.Using(19),(20)and(27),theconditionfordiscrimination(Eq.(10))reads(28)(0.05)2+)(0.069A)2+(0.027B)2*"a0.10AU#100.4(mv−12)"(0.103)2.ConsideringthecaseofaplanetwithM=6MEarth,R=2REarthinfrontofaSun-likestar,observedwith:•CoRoTduring150daysandwith40hofHARPSfollow-upona3.6mtelescope,onereadsA=B=1.InthesquarebracketofEq.(28),thetermcontainingA(theuncertaintyonthemass)dominatesthatwithBbyafactor5.AlthoughwehavealreadyconsideredaverylongintegrationtimefortheDopplermeasurement,thebottleneckremainstheun-certaintyontheplanetarymass,evenwhenusingthesmalltelescopeofCoRoT;•Keplerduring4yrsanda40hfollow-upwithanHARPS-likespectrometerona3.6m(8m)telescope,onereadsA=1(0.45),B=0.09.Now,inEq.(28),theuncertaintyonthemassfullydominatesthatontheradiusbyafac-tor∼800(160).Thewaytoimprovethediscriminationbetweenthetwotypesofplanetsisthentoimprovetheac-curacyofthemassbyusinglargertelescopes.Thisisundertheassumptionthat,forfaintstars,theuncer-taintiesonMplandF/Faredominatedbyshotnoise,ratherthanbytheastronomicalstellarnoiseortheinstrumentcapa-bilities.Arealisticdetectionexercise,includingstellarnoisesinCoRoTsimulatedlightcurves,hasshownthatstellarmicro-variabilitylimitsthedetectiononlyinextremecases(Moutouetal.,2005);however,theimpactofstellarmicro-variabilityontransitparameterestimationisstillunderstudy.Fig.6showstheregionsinthe(mv,a)planewherethedis-criminationbetweenanocean-planetandarockyplanetwith6JID:YICARAID:8261/FLA[m5+;v1.73;Prn:12/06/2007;8:22]P.12(1-16)12F.Selsisetal./Icarus•••(••••)•••–•••Fig.6.Regionsinthe(targetmagnitudemv,planettostardistancea)planewherethediscriminationbetweenanOPandarockyplanetcanbemadeat2σ(95%confidence)whentheirmassesareMpl=6or10MEarth,aroundGVstars.Thecorrespondingradiiofocean-planetsareROP=2.0,or2.3REarth,respectively.TheplanetaryradiusisdeterminedusingeithertheCoRoTinstrument(a)ortheKeplerone(b).ThemassdeterminationismadewithaDopplervelocimeterbasedontheperformancesoftheHARPSspectrometermountedona3.6m(40hofcumulatedobservations)or,speculatively,onan8mtelescope.Itisassumedthatbothtransitamplitudeandstellarreflexvelocityaccuraciesareshotnoise-limited.Theuppershadingindicatesaregionthatisnotaccessibleasaresultofinstrumentalconstraints(toolongorbitalperiods).ThelowershadingsindicatearegionwhereOPsmay,ormaynot,havekepttheirwater,becauseourestimatesofwatererosionareonlyupperlimits.Panels(c)and(d)arethehistogramsofthetargetstarsforCoRoTandKepler,respectively.Iftheswellingofahotocean-planetwasconfirmedtobeof10%(seeSection2.1),itwouldsignificantlyhelptodiscriminatebetweenOPandrockyplanets.Forinstance,theboundaryforthe6MEarthplanetwouldapproximatelymovetothatofthe10MEarthcase.Earth-massora10Earth-massplanetaroundaGVstarcanbemadeat2σ(95%confidence).Twotransitinstrumentsarecon-sideredforthetransitmeasurements,thatoftheCoRoTandKe-plermissions.TheDopplerdeterminationofthemassisbasedontheperformancesoftheHARPSspectrometermountedona3.6-mtelescope(thepresentsituation)oran8-mtelescope(aspeculativesituation).Itappearsthatthetargetstarsoftran-sitmissionsarefaintobjects,whichisaserioushandicapfortheDopplermeasurements.Asaresult,thenatureofsmallplan-etscanbeonlybedeterminedforthebrighteststars(mv<13)andtheshortest-periodplanets,atleastforCoRoT.ThispointsouttheimportanceofdetermininghowclosetotheirstarsOPsareexpectedtokeeptheirwater(Section5).Ourestimatesofwatererosionareonlyupperlimits.Planetsthatareinsidetheboundarieswehavefoundmay,ormaynot,havelosttheirwater.Theirexactnaturewillbeconstrainedbyfurthermodelinganddeterminedbytheobservations.Atfixedmagnitude,starswithsmallerradii,e.g.Kstars,aremorefavorable,buttherearefewerofthemthanGstarsinamagnitude-limitedsample.IntheCoRoTtargetlist,thefractionofstarswithmagnitude11,12,13and14areexpectedtobe1,2.3,7and15%ofthetotalnumberoftargets,respectively(Bordéetal.,2003).Thesestarsarenotthemajorityofthetargetobjects,buttheyarenotaminutefractionofthemeither.However,starswithmv"12willsaturatethebrightestpixelsofthepowerSpreadFunction(PSF)intheCCD.Presently,thesestarsarenotconsideredforphotometrybecausetheyaretechnicallymoredifficulttomoni-tor,butthissituationmayimprovesometimeafterthebeginningofthemission.Tobeoptimistic,wewillkeepthesestarsinourfurtherestimates,especiallyastheyarethosewiththehighestS/N.Withinthearbitraryhypothesisthateachstar,approximatedasaGstar(themostprobablestellartypeinthemagnitude-limitedCoRoTfield),hasa6ME[10ME]planetat0.05AU,thenumberofdetectionsbyCoRoTwithsufficientaccuracyfordeterminingthenatureofthisplanetwouldbethetotalnumberofstarswhicharebrightenough(mv"13,accord-ingtoFig.6a)timestheprobabilityoftransit(Rst/a=9%).Forthetotalmission(5stellarfields)andMpl=6ME,thisJID:YICARAID:8261/FLA[m5+;v1.73;Prn:12/06/2007;8:22]P.13(1-16)Identificationofhotocean-planets13numberwouldbethenumberofstarswithmv"13,or5×12,000×(1%+2.3%+7%)times0.09,or∼550detections.ForMpl=10ME,thelimitismv"14,andthecorrespondingnumberwouldbe∼1400detections.Inallcases,along(80h)Dopplerfollow-upofthetransitsonatleasta3.6mtelescopeisnecessary.However,thereadershouldnotmisinterpretthesenumbers.WearenotclaimingthatCoRoTisgoingtodetecthundredsof10Earth-massplanets,forthemerereasonthatitwouldbeimpossibletodedicate40hofHARPStimeperstarforsuchanumberofobjects(40h×550objects=22,000h∼3000nights∼8yearsfulltimeassuminga100%efficiency!).Theestimateismainlyinterestingforthepurposeofdetermin-ingatwhatleveltheabsenceofanydetectionwouldlimittheabundanceofsuchplanets.Ifthenumberofplanetswithmassesintherange6–10MEarthanddistancetotheirstarlessthan0.1AUwassimilartothenumberofhotJupiters,i.e.1%,thenumberofexpecteddetectionswouldbebetween5and10.Conversely,anon-detectionofsuchplanets,ifouraccuracyestimatesarecorrect,wouldindicatethattheyarenotpresentatthelevelof∼0.5%.IfamoreaccurateDopplerfollow-upwerepossible,thosediscriminatingdetectionswouldbemorenumerousbecausethemassmeasurementaccuracyispresentlythebottleneck(Eq.(28)).ForKepler,accordingtoFig.6b,ifthefollow-upwasmadewithperformedwithaDopplervelocimeteranalogoustoHARPSona8mtelescope,forMpl=6MEarth,onewouldex-pectthatplanetsata=0.10AUcanbecharacterizedaroundstarswithmv"14.5.Thetotalnumberofmonitoredstarsis100,000.IntheKeplerfieldthere∼600,000starswithmv"16.5,15%ofwhichhavemv"14.5(D.Koch,personalcom-munication).Withtheconservativeassumptionthatthedistrib-utioninmagnitudeofthemonitoredstarsisthesameasfortheotherstarsofthefield,andtheassumptionofa1%abundanceofhotearthswithmasses6–10MEarth,onewouldexpecttobeabletocharacterize5to10ofthem.ThedifferencewiththeCoRoTcaseseemsnottobehuge,althoughitmustbepointedoutthattheprospectofcharacterizingterrestrialplanetswithKeplerisonasignificantlyfirmerbasisbecausesomeofthesemeasurementsareexpectedtohaveafairS/N,whereasthisisnotsofrequentwithCoRoT.WepointoutthekeyinterestofbuildingaDopplerspectrom-eterwithperformancesimilartothatofHARPS,onan8mclasstelescopeforthestudyofterrestrialexoplanets.Thisistrueforthefollow-upofCoRoTanditisespeciallyobviousforthatofKepler.Thisisundertheoptimistichypothesisthatforfaintstars(mv=11–14)bothtransitamplitudeandstellarreflexveloc-ityarebasicallylimitedbyshotnoiseandnotbythestellarvariabilityorinstrumentnoise.Ifthestellarnoisedominates,usinglargetelescopeswillnotbeuseful;theonlywayto(slowly)improvetheS/Nwillbetoincreasetheintegrationtime.4.Discussion:Mass–radiusdegeneracybetweenOPs,H2-richplanetsandcarbonplanets?Aspointedoutintheintroduction,theoreticalworks(Alibertetal.,2006;Rafikov,2006)predicttheexistenceofshortperiodplanetswithacoremass>6MEarthandawidediversityofH2–Heenvelopes:fromMenv%1MEarth(OPsandsuper-telluricplanets),toafewMEarth(likethehotNeptunesinthesystemHD69830),andfinallyMenv-1MEarth(giantplanets).Inplanetaryformationmodels,thecoreisinitiallymadeoficy–rockymaterialfrombeyondthesnowline,butitaccretesmorerockyplanetesimalswhenmigratingbetweenthesnowlineanditsfinalshortperiodorbit.Thecollisionsoccurringdur-ingtheplanetformationcanleadtoafinalcompositionsignif-icantlydepletedinvolatilescomparedwithatypicalcometarycomposition.Also,theplanetcanofcourseconsistofrefrac-torymaterialsonly,ifmigrationstartedclosertothesnowlineandifenoughrockyplanetesimalsareavailable.Therefore,agivencouple[R,M]maycorrespondtoeitheranOPofmassMandanegligibleH2–HeatmosphereortoaplanetwithM=Mcore+Menv.Forthesamemass,ahy-drogenenvelopewouldbesignificantlymoreexpandedthananH2Oenvelope.AttheorbitalperiodsrelevantforCoRoT,thehightemperatureduetothestrongirradiationwouldpro-duceanevenmoreimportantswellingoftheH2–Heenvelope.AnH2–Heenvelopemoremassivethanabout0.1MEarthwouldproduceaplanetaryradiusofmorethan3REarth,whichislargerthanan10MEarthOP.Togiveanexample(kindlypro-videdbyI.Baraffe),aplanetat0.1AU,witharockycoreof[Mcore=7.5MEarth,Rcore=1.76REarth]andanH2–Heen-velopeof2.5MEarth,wouldhaveatotalradiusof5.6REarth.Therefore,tomimicanOPwithnoH2–He,themassofH2–HesurroundingarockycorehastobesignificantlylessmassivethanthewaterreservoirofthisOP.ObservingaplanetwiththispreciseamountofH2–He,obtainedbyaccretionofthegaseousenvelopeanditssubsequentpartiallossbyatmosphericescape,appearsqualitativelyratherunlikely.Althoughthediversityofmass–radiusrelationshipsshouldclearlybeaddressedquantitativelyusingamodelofplanetformation(includingmigration),therebyprovidingarealisticdistributionofM(MRocks,MIce,MH2)versusorbitaldistance,weconcludeherethattheidentificationofmassiveOPscannotrigorouslybedonewithasinglecase.Toshowupamongtran-sitingplanetshavingawiderangeofH2–Heenvelopemasses,OPswithnegligibleH2–Heatmosphereshavetoberelativelyabundant.Regardingthispoint,short-period,low-massOPs(M<6MEarth,a<0.1AU)providethebestcase,becausetheaccretionofgasisnotexpectedtobeimportantforthesemasses(Rafikov,2006),whiletheescapeofH2isfast(muchfasterthanH2O)attheseorbitaldistances(Baraffeetal.,2005).Moremassive,H2-deficientOPsatlongerperiodsmaybedifficulttodistinguishwithinapopulationofplanets(OPsortelluric)withH2envelopes.ItisalsoimportanttonotethatallratiosMIce/MRocksbe-tween∼0and1areexpected.Foragivenplanetarymass,planetaryradiishouldbefoundatanyvaluebetweenthatofasilicateplanetandourOPprototypes(i.e.betweenthetwoJID:YICARAID:8261/FLA[m5+;v1.73;Prn:12/06/2007;8:22]P.14(1-16)14F.Selsisetal./Icarus•••(••••)•••–•••curvesofFig.1).ForM>6MEarthwecanexpectatwodis-tributionsofobjects,hopefullyseparatedbyagap,fortheOPprototypesandthehotNeptune/Saturn/Jupiterobjects.KuchnerandSeager(2005)proposedtheexistenceofcar-bonplanets,formedinenvironmentscharacterizedbyC/O>1.Intheabsenceofdetailedmodelingoftheirstructureandoftheirradiusasafunctionoftheirmass,itisdifficulttoknowifthesecarbonplanetscouldbeidentifiedorconfusedwithsilicate-dominatedplanetsorOPs.Thatisanopenquestionforfurtherstudies.WhatwecansayatthispointisthattheC/Oratiocanbemeasured(withinsomesignificantuncertain-ties)instellarphotospheres,soweshouldbeaware,fortherare(any?)mainsequencestarswithhighC/Oratiowhereatran-sitwillbeobserved,thatacarbonplanetmightbethere.Evenforstarswitha“solar”C/Oratio,planets(e.g.Jupiter)canstillbesignificantlyenrichedincarbon.Inaddition,theicymater-ialfoundbeyondthesnowlinedoesnotconsistsonlyofH2Oice,CO2,CH4andCOarealsoimportantconstituents.Theac-cretionofoneEarthmassofH2Oiceimpliestheaccretionofabout0.1Earthmassofthesecarbon-basedvolatile,whichcanhaveasignificanteffectonthestructureandtheevolutionoftheplanet.Weshouldthusremainawarethatalargediversityofel-ementalcompositioncanexistamongplanetsandthatcannotbecharacterizedonlybythew=water/silicateratio.Never-theless,thedetectionbyCoRoTorKepler,andaroundK,GorFstars,ofshort-periodlow-massplanets(Torb"1month,Mpl"6MEarth),withradiisignificantlylargerthansilicate-dominatedplanetswouldunveilapopulationofvolatile-richplanets.ThebestcandidateforthisvolatilebeingH2O.5.ConclusionWehaverevisitedthetheoreticalestimateofthephysicaldensityofhotocean-planets(OPs)andconcludedthatthere-sultsestablishedforplanetsinthehabitablezone(HZ)arestillvalidmuchclosertotheirstar,becausetheirexpectedswellingshouldremainlimitedwhencomparedtotheplanetaryradius.Curiouslyenough,averyhotOPhasathickwateratmosphereindirectcontactwitha(highpressure)icemantle.ModelsoftheinternalstructureofOPsyieldameandensitysmallerthanthatofrockyplanets.Forthesamemassthereisadifferenceinradiusof∼20%,thatcanreach30%forthehottestOPs,duetotheswellingoftheirwaterenvelopeInthispaperwehaveconsideredwhetheranidentificationofOPsbasedonthemeasurementoftheirdensitiesisactuallypossibleinthenear-tomid-future.Ifstrongenough,migrationcouldbringthesespeculativeplanetstowithintheclosevicinityoftheirparentstar,wheretheirmassesandradiicanbebettermeasuredbyRadialVeloc-ityandbytransitphotometry.CoRoTwilldetectplanetswithperiodslessthan75days,whereasKeplerwillhaveaccesstoperiodslessthan1.33years.Inbothcases,close-inplanetswillbetheeasiestonestodetectandtocharacterize.Wehaveestimatedtheerosionrateofahotocean-planetbythermalescapedrivenbyexosphereheating(withXandUVradiation)andbynon-thermalescapedrivenbyejectedparticlesfromthestar.Wefoundthatthewaterreservoirisonlyweaklyaffected,sothatanOPshouldkeepitspeculiarcompositionevenfairlyclosetoitsstar.Forexample,infrontofasolar-typestar,a6Earth-massOPshouldretainmostofitswatercontentformorethan5Gyrsifitisbeyond0.04AUfromitsstar.Then,wehaveaddressedtheaccuracyneededtodiscrimi-natebetweenOPsandrockyplanets,assumingthattheplan-etarymodelsareadequate.ThesourcesofuncertaintyontheplanetarydensityarethoseonthemassdeterminationbyRa-dialVelocitymeasurements,thestellarradiusdetermination,andthephotometricmeasurementduringthetransits.Asex-pected,theaccuracyoftheKeplerphotometryishigherthanthatofCoRoT.Howeverwiththepresentlyavailableinstru-ments,itistheuncertaintyonRadialVelocitymeasurementsthatisthelimitingfactorforexpecteddetectionsbybothtran-sitmissions.Asaresult,thedeterminationofthenatureoftheseplanetsseemspossibleonlyintheadequatedomainsofthe(mv,a)planethatareshowninFig.6forbothmissions.Theycorre-spondtothebrighteststars(mv<14)andtotheplanetsclosesttotheirstar,especiallyforCoRoT,butbeyondthelimitforthesurvivaloftheirwaterreservoir.Ifeachstarhada6–10Earthmassplanetat∼0.10AU,anarbitraryhypothesis,thenumberofdetectionswithCoRoTwouldbeseveralhundred.Conversely,theabsenceofdetectionofsuchplanetswouldindicatethattheyarenotpresentatthelevelof∼1%.AclearconclusionofourstudyisthatfullbenefitofthehighphotometricprecisionofCoRot,andparticularlyKepler,canbeobtainedonlyifanewgenerationofRadialVelocityinstru-mentsisbuiltthatcanmakeaccuratemeasurementsonfaintstars.Inthatcase,theidentificationofOPscouldbedoneonasignificantlylargersampleofstars.Inthemeantime,asignif-icantfractionofhighperformanceDopplervelocimeterssuchasHARPSshouldbedevotedtothefollow-upofthemostin-terestingcandidatesfoundbyCoRoT,ifwehopetobeabletodiscriminatebetweenthedifferentpossiblecompositionsofter-restrialplanets.AcknowledgmentWearegratefultoOdileDutuit,RolandThissen,Alexan-droMorbidelli,andDavidKochforvaluablediscussionsonthewatermoleculeabsorptionintheEUV,thepossibleabundanceoftheseputativeplanetsandpreciseinformationonKepler,re-spectively.WearealsogratefultoananonymousrefereeandtoMarcKuchnerforvaluablecommentsonourinitialmanu-script.ThisworkwassupportedbyCNRSandCNESinpartbytheProgrammeNationaldePlanétologieandtheGroupedeRecherche“Exobio,”andundercontract1256791withtheJetPropulsionLaboratory(JPL)fundedbyNASAthroughtheMichelsonFellowshipProgram.Wealsoacknowledgetheben-efitfromtheISSITeam“EvolutionofHabitablePlanets.”ReferencesAlibert,Y.,Baraffe,I.,Benz,W.,Chabrier,G.,Mordasini,C.,Lovis,C.,Mayor,M.,Pepe,F.,Bouchy,F.,Queloz,D.,Udry,S.,2006.Formationandstruc-tureofthethreeNeptune-massplanetssystemaroundHD69830.Astron.Astrophys.455,L25–L28.JID:YICARAID:8261/FLA[m5+;v1.73;Prn:12/06/2007;8:22]P.15(1-16)Identificationofhotocean-planets15Alonso,R.,Brown,T.M.,Torres,G.,Latham,D.W.,Sozzetti,A.,Mandushev,G.,Belmonte,J.A.,Charbonneau,D.,Deeg,H.J.,Dunham,E.W.,O’Dono-van,F.T.,Stefanik,R.P.,2004.TrES-1:ThetransitingplanetofabrightK0VStar.Astrophys.J.613,L153–L156.Baraffe,I.,Selsis,F.,Chabrier,G.,Barman,T.S.,Allard,F.,Hauschildt,P.H.,Lammer,H.,2004.Theeffectofevaporationontheevolutionofclose-ingiantplanets.Astron.Astrophys.419,L13–L16.Baraffe,I.,Chabrier,G.,Barman,T.S.,Selsis,F.,Allard,F.,Hauschildt,P.H.,2005.HotJupitersandhotNeptunes:Acommonorigin?Astron.Astro-phys.436,L47–L51.Barman,T.,Hauschildt,P.,Allard,F.,2005.Phase-dependentpropertiesofex-trasolarplanetatmospheres.Astrophys.J.632-2,1132–1139.Bordé,P.,Rouan,D.,Léger,A.,2003.ExoplanetdetectioncapabilityoftheCOROTspacemission.Astron.Astrophys.405,1137–1144.Bouchy,F.,Pepe,F.,Queloz,D.,2001.Fundamentalphotonnoiselimittoradialvelocitymeasurements.Astron.Astrophys.374,733–739.Charbonneau,D.,Brown,T.,Burrows,A.,Laughlin,G.,2006.Whenextra-solarplanetstransittheirparentstars.In:ProtostarsandPlanetsV,astro-ph/0603376.Chassefiére,E.,1996.Hydrodynamicescapeofhydrogenfromahotwater-richatmosphere:ThecaseofVenus.J.Geophys.Res.101,26039–26056.Cox,A.N.,2000.AllenAstrophysicalQuantities,fourthed.Springer-Verlag,NewYork,ISBN0-387-98746-0.p.382.DiBenedetto,G.P.,2005.Predictingaccuratestellarangulardiametersbythenear-infraredsurfacebrightnesstechnique.Mon.Not.R.Astron.Soc.357,174–190.Ehrenreich,D.,Tinetti,G.,LecavelierdesEtangs,A.,Vidal-Madjar,A.,Selsis,F.,2006.ThetransmissionspectrumofEarth-sizetransitingplanets.Astron.Astrophys.448,379–393.Fortney,J.J.,2005.Theeffectofcondensatesonthecharacterizationoftran-sitingplanetatmosphereswithtransmissionspectroscopy.Mon.Not.R.Astron.Soc.364,649–653.Fridlund,M.,Baglin,A.,Conroy,L.,Lochard,J.,2006.TheCoRoTBook.ESA-SP1306,Sect.VII,inpress(http://corotsol.obspm.fr/web-instrum/payload.param/).Griessmeier,J.-M.,Stadelmann,A.,Penz,T.,Lammer,H.,Selsis,F.,Ribas,I.,Guinan,E.F.,Motschmann,U.,Biernat,H.-K.,Weiss,W.W.,2004.Theeffectoftidallockingonthemagnetosphericandatmosphericevolutionof“hotJupiters.”Astron.Astrophys.425,753–762.Griessmeier,J.-M.,Stadelmann,A.,Motschmann,U.,Belisheva,N.K.,Lam-mer,H.,Biernat,H.K.,2005.CosmicrayimpactonextrasolarEarth-likeplanetsinclose-inhabitablezones.Astrobiology5,587–603.Kasting,J.F.,1988.Runawayandmoistgreenhouseatmospheresandtheevo-lutionofEarthandVenus.Icarus74,472–494.Kervella,P.,Thevenin,F.,DiFolco,E.,Segransan,D.,2004.Theangularsizesofdwarfstarsandsubgiants:Surfacebrightnessrelationscalibratedbyin-terferometry.Astron.Astrophys.426,297–307.Khodachenko,M.L.,Lammer,H.,Lichtenegger,H.I.M.,Langmayr,D.,Erkaev,N.V.,Griessmeier,J.-M.,Leitner,M.,Penz,T.,Biernat,H.,K.,Motschmann,U.,Rucker,H.O.,2007.Masslossof“hotJupiters”ImplicationsforCoRoTdiscoveries.PartI:Theimportanceofmagnetosphericprotectionofaplanetagainstionlosscausedbycoronalmassejections.Planet.SpaceSci.55(5),631–642.Koch,D.,and15colleagues,2006.TheKeplerMission:Astrophysicsandeclipsingbinaries.CloseBinariesinthe21thCenturySyros,Greece,27–30June2005.Astrophys.SpaceSci.J.,inpress.Kuchner,M.J.,2003.Volatile-richEarth-massplanetsintheHabitableZone.Astrophys.J.596,L105–L108.Kuchner,M.J.,Seager,S.,2005.Extrasolarcarbonplanets.Astrophys.J.,sub-mittedforpublication,astro-ph/0504214.Lammer,H.,Selsis,F.,Ribas,I.,Guinan,E.F.,Bauer,S.J.,Weiss,W.W.,2003.AtmosphericlossofexoplanetsresultingfromstellarX-rayandextreme-ultravioletheating.Astrophys.J.598,L121–L124.Lammer,H.,Lichtenegger,H.I.M.,Kulikov,Yu.N.,Griessmeier,J.-M.,Terada,N.,Erkaev,N.V.,Biernat,H.K.,Khodachenko,M.L.,Ribas,I.,Penz,T.,Selsis,F.,2006.CMEactivityoflowmassMstarsasanimportantfactorforthehabitabilityofterrestrialexoplanets.PartII.CMEinducedionpickupofEarth-likeexoplanetsinclose-inhabitablezones.Astrobiology,inpress.Léger,A.,Selsis,F.,Sotin,C.,Guillot,T.,Despois,D.,Lammer,H.,Ollivier,M.,Brachet,F.,2003.Anewfamilyofplanets?Ocean-planets.In:Fridlund,M.,Henning,T.(Eds.),ProceedingsoftheConference“TowardsOtherEarthsDarwin/TPFandtheSearchforExtrasolarTerrestrialPlanets.”In:ESA-SP539,pp.253–259.Léger,A.,Selsis,F.,Sotin,C.,Guillot,T.,Despois,D.,Mawet,D.,Ollivier,M.,Labèque,A.,Valette,C.,Brachet,F.,Chazelas,B.,Lammer,H.,2004.Anewfamilyofplanets,“ocean-planets”.Icarus169,499–504.Lovis,C.,and13colleagues,2006.AnextrasolarplanetarysystemwiththreeNeptune-massplanets.Nature441,305–309.Mayor,M.,and31colleagues,2003.SettingnewstandardswithHARPS.TheMessenger114,20–24.Moutou,C.,Pont,F.,Barge,P.,and13colleagues,2005.Comparativeblindtestoffiveplanetarytransitdetectionalgorithmsonrealisticsyntheticlightcurves.Astron.Astrophys.437,355–368.O’Donovan,F.T.,Barman,T.S.,Allard,F.,Hauschildt,P.H.,Lammer,H.,and17colleagues,2006.TrES-2:ThefirsttransitingplanetintheKeplerField.Astrophys.J.651,L61–L64.Papaloizou,J.C.B,Nelson,R.P.,Kley,W.,Masset,F.S.Artmowicz,P.,2006.Disk–planetinteractionsduringplanetformation.In:ProtostarsandPlanetsV.Univ.ofArizonaPress,Tucson.Penz,T.,Langmayr,D.,Erkaev,N.V.,Kulikov,Yu.N.,Lammer,H.,Biernat,H.K.,Selsis,F.,Cecchi-Pestellini,C.,Micela,G.,Barge,P.,Deleuil,M.,Léger,A.,2006.Masslossfrom“hotJupiters”—ImplicationsforCoRoTdiscoveries.PartII.Longtimethermalatmosphericevaporationmodeling.P&SS,submittedforpublication.Pepe,F.,Mayor,M.,Queloz,andcolleagues,2005.Onthetrackofverylow-massplanetswithHARPS.TheMessenger120,22–25.Press,W.,Teukolsky,S.,Vetterling,W.,Flannery,B.,2002.NumericalRecipesinC.CambridgeUniv.Press,Cambridge.Protanov,A.,2002.AnalyseStatistiquedesDonnéesExpérimentales.EDPSci-ence,France,ISBN2-86883-590-2,pp.51–57.Rafikov,R.R.,2006.Atmospheresofprotoplanetarycores:Criticalmassfornucleatedinstability.Astrophys.J.648,666–682.Raymond,S.N.,Mandell,A.M.,Sigurdsson,S.,2006a.ExoticEarths:Forminghabitableworldswithgiantplanetmigration.Science313,1413–1416.Raymond,S.N.,Quinn,T.,Lunine,J.I.,2006b.High-resolutionsimulationsofthefinalassemblyofEarth-likeplanets.I.Terrestrialaccretionanddynam-ics.Icarus183,265–282.Raymond,S.N.,Quinn,T.,Lunine,J.I.,2007.High-resolutionsimulationsofthefinalassemblyofEarth-likeplanets.II.Waterdeliveryandplanetaryhabitability.Astrobiology7(1),66–84.Ribas,I.,Guinan,E.F.,Güdel,M.,Audard,M.,2005.Evolutionofthesolaractivityovertimeandeffectsonplanetaryatmospheres.I.High-energyir-radiances(1–1700A).Astrophys.J.622,680–694.Rouan,D.,Baglin,A.,Barge,P.,Copet,E.,Deleuil,M.,Leger,A.,Schneider,J.,Toublanc,D.,Vuillemin,A.,1999.SearchingforexosolarplanetswiththeCOROTspacemission.Phys.Chem.EarthC24,567–571.Santos,N.C.,Pont,F.,Melo,C.,Israelian,G.,Bouchy,F.,Mayor,M.,Moutou,C.,Queloz,D.,Udry,S.,Guillot,T.,2006.Highresolutionspectroscopyofstarswithtransitingplanets:ThecasesofOGLE-TR-10,56,111,113andTrES-1.Astron.Astrophys.450,825–831.Sato,B.,Fischer,D.A.,Henry,G.W.,and18colleagues,2005.TheN2KCon-sortium.II.AtransitinghotSaturnaroundHD149026withalargedensecore.Astrophys.J.633,465–473.Sotin,C.,Grasset,O.,Moquet,A.,2007.Mass/radiusforextrasolarEarth-likeplanetsandocean-planets.Icarus,inpress.Terada,N.,Machida,S.,Shinagawa,H.,2002.GlobalhybridsimulationoftheKelvin–HelmholtzinstabilityattheVenusionopause.J.Geophys.Res.107,1471–1490.Terquem,C.,Papaloizou,J.C.B.,2006.Migrationandtheformationofsystemsofhotsuper-EarthsandNeptunes.AstrophysJ.,inpress,ArXivAstro-physicse-prints:astro-ph/0609779.Tian,F.,Toon,O.B.,Pavlov,A.A.,DeSterck,H.,2005.Transonichydrody-namicescapeofhydrogenfromextrasolarplanetaryatmospheres.Astro-phys.J.621,1049–1060.Pleasecitethisarticleinpressas:F.Selsisetal.,Couldweidentifyhotocean-planetswithCoRoT,KeplerandDopplervelocimetry?,Icarus(2007),doi:10.1016/j.icarus.2007.04.010JID:YICARAID:8261/FLA[m5+;v1.73;Prn:12/06/2007;8:22]P.16(1-16)16F.Selsisetal./Icarus•••(••••)•••–•••Valencia,D.,O’Connell,R.J.,Sasselov,D.,2006.Internalstructureofmassiveterrestrialplanets.Icarus181,545–554.Vidal-Madjar,A.,LecavelierdesEtangs,A.,Désert,J.-M.,Ballester,G.E.,Ferlet,R.,Hébrard,G.,Mayor,M.,2003.Anextendedupperat-mospherearoundtheextrasolarplanetHD209458b.Nature422,143–146.Vidal-Madjar,A.,Désert,J.-M.,LecavelierdesEtangs,A.,Hébrard,G.,Ballester,G.E.,Ehrenreich,D.,Ferlet,R.,McConnell,J.C.,Mayor,M.,Parkinson,C.D.,2004.Detectionofoxygenandcarboninthehydrody-namicallyescapingatmosphereoftheextrasolarplanetHD209458b.As-trophys.J.604,L69–L72.Watson,A.J.,Donahue,T.M.,Walker,J.C.G.,1981.Thedynamicsofarapidlyescapingatmosphere—ApplicationstotheevolutionofEarthandVenus.Icarus48,150–166.Wood,B.E.,Müller,H.-R.,Zank,G.P.,Linsky,J.L.,2002.Measuredmass-lossratesofsolar-likestarsasafunctionofageandactivity.Astrophys.J.574,412–425.Wood,B.E.,Müller,H.-R.,Zank,G.P.,Linsky,J.L.,Redfield,S.,2005.Newmass-lossmeasurementsfromastrosphericLyαabsorption.Astrophys.J.628,L143–L146.Yelle,R.V.,2004.Aeronomyofextrasolargiantplanetsatsmallorbitaldis-tances.Icarus170,167–179. |
0805.2108 | 5 | 0805 | 2009-04-12T18:47:59 | Transitional solar dynamics, cosmic rays and global warming | [
"astro-ph",
"nlin.CD",
"physics.geo-ph"
] | Solar activity is studied using a cluster analysis of the time-fluctuations of the sunspot number. It is shown that in an Historic period the high activity components of the solar cycles exhibit strong clustering, whereas in a Modern period (last seven solar cycles: 1933-2007) they exhibit a white-noise (non-)clustering behavior. Using this observation it is shown that in the Historic period, emergence of the sunspots in the solar photosphere was strongly dominated by turbulent photospheric convection. In the Modern period, this domination was broken by a new more active dynamics of the inner layers of the convection zone. Then, it is shown that the dramatic change of the sun dynamics at the transitional period (between the Historic and Modern periods, solar cycle 1933-1944yy) had a clear detectable impact on Earth climate. A scenario of a chain of transitions in the solar convective zone is suggested in order to explain the observations, and a forecast for the global warming is suggested on the basis of this scenario. A relation between the recent transitions and solar long-period chaotic dynamics has been found. Contribution of the galactic turbulence (due to galactic cosmic rays) has been discussed. These results are also considered in a content of chaotic climate dynamics at millennial timescales. | astro-ph | astro-ph |
Transitional solar dynamics, cosmic rays and global warming
ICAR, P.O. Box 31155, Jerusalem 91000, Israel
A. Bershadskii
Solar activity is studied using a cluster analysis of the time-fluctuations of the sunspot number.
It is shown that in an Historic period the high activity components of the solar cycles exhibit
strong clustering, whereas in a Modern period (last seven solar cycles: 1933-2007) they exhibit
a white-noise (non-)clustering behavior. Using this observation it is shown that in the Historic
period, emergence of the sunspots in the solar photosphere was strongly dominated by turbulent
photospheric convection. In the Modern period, this domination was broken by a new more active
dynamics of the inner layers of the convection zone. Then, it is shown that the dramatic change of
the sun dynamics at the transitional period (between the Historic and Modern periods, solar cycle
1933-1944yy) had a clear detectable impact on Earth climate. A scenario of a chain of transitions
in the solar convective zone is suggested in order to explain the observations, and a forecast for the
global warming is suggested on the basis of this scenario. A relation between the recent transitions
and solar long-period chaotic dynamics has been found. Contribution of the galactic turbulence
(due to galactic cosmic rays) has been discussed. These results are also considered in a content of
chaotic climate dynamics at millennial timescales.
PACS numbers: 92.70.Qr, 92.70.Mn, 96.60.qd, 98.70.Sa
INTRODUCTION
The sunspot number is the main direct and reliable
source of information about the solar dynamics for his-
toric period. This information is crucial, for instance, for
analysis of a possible connection between the sun activity
and the global warming. In a recent papers [1],[2] results
)
C
0
(
y
l
a
m
o
n
a
e
r
u
t
a
r
e
p
m
e
T
0.4
0.2
0
-0.2
-0.4
-0.6
C
T (land and marine)
B
A
I
II
III
300
250
200
150
100
50
N
S
S
1800
1830
1860
1890
1920
1950
1980
0
2010
Year
FIG. 1: Sunspot number (SSN, monthly) vs time [14]. The
dashed straight lines separate between periods of different in-
tensity of the solar activity. The solid curve shows the global
temperature anomaly (combined land and marine, 7-years
running average) [20].
of a reconstruction of the sunspot number were presented
for the past 11,400 years. The reconstruction shows that
[1]: "..the level of solar activity during the past 70 years
is exceptional, and the previous period of equally high
activity occurred more than 8,000 years ago" (section III
in Fig. 1) .
Figure 1 gives first indication ('by eye'), that the
prominent maxima of the global temperature data (solid
curve in the Fig. 1) correspond to transitions between pe-
riods of different intensity of the Sun activity (character-
ized by the monthly sunspot number-SSN). This observa-
tion can be considered as an indication of a strong impact
of the solar activity transitions on the global Earth cli-
mate. Therefore, understanding of the physical processes
in the Sun, which cause these activity transitions, seems
to be crucial for any serious forecast for global Earth cli-
mate.
SUNSPOTS
When magnetic field lines are twisted and poke
through the solar photosphere the sunspots appear as
the visible counterparts of magnetic flux tubes in the
convective zone of the sun. Since a strong magnetic field
is considered as a primary phenomenon that controls gen-
eration of the sunspots the crucial question is: Where has
the magnetic field itself been generated? The location of
the solar dynamos is the subject of vigorous discussions
in recent years. A general consensus had been developed
to consider the shear layer at the bottom of the convec-
tion zone as the main source of the solar magnetic field
[3] (see, for a recent review [4]). In recent years, however,
the existence of a prominent radial shear layer near the
top of the convection zone has become rather obvious and
the problem again became actual. The presence of large-
scale meandering flow fields (like jet streams), banded
zonal flows and evolving meridional circulations together
with intensive multiscale turbulence shows that the near
surface layer is a very complex system, which can sig-
nificantly affect the processes of the magnetic field and
the sunspots generation. There could be two sources for
the poloidal magnetic field: one near the bottom of the
convection zone (or just below it [3]), another resulting
from an active-region tilt near the surface of the convec-
tion zone. For the recently renewed Babcock-Leighton
[5],[6] solar dynamo scenario, for instance, a combina-
tion of the sources was assumed for predicting future so-
lar activity levels [7], [8].
In this scenario the surface
generated poloidal magnetic field is carried to the bot-
tom of the convection zone by turbulent diffusion or by
the meridional circulation. The toroidal magnetic field is
produced from this poloidal field by differential rotation
in the bottom shear layer. Destabilization and emer-
gence of the toroidal fields (in the form of curved tubes)
due to magnetic buoyancy can be considered as a source
of pairs of sunspots of opposite polarity. The turbulent
convection in the convection zone and, especially, in the
near-surface layer captures the magnetic flux tubes and
either disperses or pulls them trough the surface to be-
come sunspots.
The magnetic field plays a passive role in the photo-
sphere and does not participate significantly in the tur-
bulent photospheric energy transfer. On the other hand,
the very complex and turbulent near-surface layer (in-
cluding photosphere) can significantly affect the process
of emergence of sunspots. The similarity of light elements
properties in the spot umbra and granulation is one of the
indication of such phenomena.
The present paper reports a direct relation between
the fluctuations in sunspot number and the temperature
of photospheric turbulent convection in the Historic pe-
riod. This relation allows for certain conclusions about
the generation mechanisms of the magnetic fields and the
sunspots. In the Modern period (section III in Fig. 1,
cf.
[1],[2]) the relative role of the surface layer (photo-
sphere) in the process of emergence of sunspots decreased
in comparison with the Historic period (section II in Fig.
1), implying a drastic increase of the relative role of the
inner layers of the convection zone in the Modern period.
TIME-CLUSTERING OF FLUCTUATIONS
[10] (see also Ref.
In order to extract new information from sunspot num-
ber data we apply the fluctuation clustering analysis sug-
gested in the Ref.
[11]). For a time
depending signal we count the number of 'zero'-crossings
of the signal (the points on the time axis where the signal
is equal to zero) in a time interval τ and consider their
running density nτ . Let us denote fluctuations of the
running density as δnτ = nτ − hnτ i, where the brackets
2
/
2
1
>
2
nt
<d
1
0.1
0.01
1
-0.37
level SSN=85
10
100
1000
[d]
FIG. 2: The standard deviation for δnτ vs τ (Historic period)
in log-log scales. The straight line (the best fit) indicates the
scaling law Eq. (1).
2
/
1
>
2
nt
<d
1
0.1
0.01
1
-0.50
level SSN=125
10
100
1000
[d]
FIG. 3: The standard deviation for δnτ vs τ (Modern period)
in log-log scales. The straight line (the best fit) indicates the
scaling law Eq. (1).
mean the average over long times. We are interested in
scaling variation of the standard deviation of the running
density fluctuations hδn2
τ i1/2 with τ
hδn2
τ i1/2 ∼ τ −α
(1)
For white noise signal
it can be derived analytically
[12],[13] that α = 1/2 (see also [10]). The same con-
sideration can be applied not only to the 'zero'-crossing
points but also to any level-crossing points of the signal.
One can see that for the Modern period (section III in
Fig. 1) the solar activity is significantly different from
that for the Historic period (section II in Fig. 1). There-
fore, in order to calculate the cluster exponent (if exists)
for this signal one should make this calculation separately
for the Modern and for the Historic periods. We are in-
terested in the active parts of the solar cycles. Therefore,
t
t
3
0.1
0.01
2
/
1
>
2
n
<
-0.37
thermal convection (Ra=1.4* 1011)
0.001
0.01
0.1
1
10
100
1000
t [s]
1
0.1
2
/
1
>
2
n
δ
<
0.01
1
-0.37
Br - 1998-2008yy
10
100
1000
τ [h]
FIG. 4: The standard deviation for δnτ vs τ for the tempera-
ture fluctuations in the Rayleigh-Bernard convection labora-
tory experiment with Ra = 1.4 × 1011 [16]. The straight line
indicates the scaling law Eq. (1).
FIG. 5: The standard deviation for δnτ vs. τ for radial com-
ponent Br of the interplanetary magnetic field, as measured
by the ACE magnetometers for the last solar cycle (hourly
average [26]).
for the Historic period let us start from the level SSN=85.
The set of the level-crossing points has a few large voids
corresponding to the weak activity periods. In the tele-
graph signal Ref.
[10], corresponding to the data set,
the large voids became so clear detectable that there is
no problem to cut them off. Then, the remaining data
have been merged providing a statistically stationary set
(about 104 data points). The robustness of the procedure
has been successfully checked.
Fig. 2 shows (in the log-log scales) dependence of the
standard deviation of the running density fluctuations
hδn2
τ i1/2 on τ for this data set. The straight line is drawn
in this figure to indicate the scaling (1). The slope of
this straight line provides us with the cluster-exponent
α = 0.37 ± 0.02. This value turned out to be insensitive
to a reasonable variation of the SSN level. Results of
analogous calculations (including the large voids 'cut off'
procedure) performed for the Modern period are shown
in Fig. 3 for the SSN level SSN=125. The calculations
performed for the Modern period provide us with the
cluster-exponent α = 0.5 ± 0.02 (and again this value
turned out to be insensitive to a reasonable variation of
the SSN level).
The exponent α ≃ 0.5 (for the Modern period) indi-
cates a random (white noise like) situation. While the
exponent α ≃ 0.37 (for the Historic period) indicates
strong clustering. The question is: Where is this strong
clustering coming from? It is shown in the paper [10]
that turbulence produces signals with strong clustering.
Moreover, the cluster exponents for these signals depend
on the turbulence intensity and they are nonsensitive to
the types of the boundary conditions. Fortunately, we
have direct estimates of the value of the main parame-
ter characterizing intensity of the turbulent convection
in photosphere: Rayleigh number Ra ∼ 1011 (see, for in-
stance [15]). In Fig. 4 we show calculation of the cluster
exponent for the temperature fluctuations in the classic
Rayleigh-Bernard convection laboratory experiment for
Ra ∼ 1011 (for a description of the experiment details
see [16]). The calculated value of the cluster exponent
α = 0.37 ± 0.01 coincides with the value of the cluster-
exponent obtained above for the sunspot number fluctu-
ations for the Historic period. The value of the Rayleigh
number Ra in the photosphere for the Historic period
has the same order as for the Modern period: Ra ∼ 1011
(see next Section). Therefore, the photospheric temper-
ature fluctuations can produce the strong clustering of
the sunspot number fluctuations for the Historic period.
This seems to be natural for the case when the photo-
spheric convection determines the sunspot emergence in
the photosphere. However, in the case when the effect of
the photospheric convection on the SSN fluctuations is
comparable with the effects of the inner convection zone
layers on the SSN fluctuations the clustering should be
randomized by the mixing of the sources, and the cluster
exponent α ≃ 0.5 (similar to the white noise signal). The
last case apparently takes place for the modern period.
Since the Rayleigh number Ra of the photospheric con-
vection preserves its order Ra ∼ 1011 with transition
from the Historic period to the Modern one (see next
Section), we can assume that just significant changes of
the dynamics of the inner layers of the convection zone
(most probably - of the bottom layer) were the main rea-
sons for the transition from the Historic to the Modern
period.
d
t
4
Cumulated aa-index
1000000
800000
600000
400000
200000
a
a
S
1
0.1
2
/
1
>
2
τ
n
δ
<
-0.37
0
1860 1880 1900 1920 1940 1960 1980 2000
t [years]
1875-1932yy
1933-2007yy
0.01
1
10
100
1000
τ [d]
FIG. 6: Cumulative aa-index vs. time. Daily aa-index was
taken from [30]. The arrow indicates beginning of the transi-
tional solar cycle.
FIG. 7: The standard deviation for δnτ vs. τ (in log-log
scales) for the daily aa-index [30]: circles correspond to the
Historic period and crosses correspond to the Modern period.
MAGNETIC FIELDS IN THE SOLAR WIND
AND ON EARTH
Although for the Modern period the turbulent con-
vection in the photosphere has no decisive impact on
the sunspots emergence, the large-scale properties of the
magnetic field coming through the sunspots into the pho-
tosphere and then to the interplanetary space (so-called
solar wind) can be strongly affected by the photospheric
motion. In order to be detected the characteristic scal-
ing scales of this impact should be larger than the scal-
ing scales of the interplanetary turbulence (cf. [24],[25]).
In particular, one can expect that the cluster-exponent
of the large-scale interplanetary magnetic field (if ex-
ists) should be close to α ≃ 0.37.
In Fig. 5 we show
cluster-exponent of the large-scale fluctuations of the ra-
dial component Br of the interplanetary magnetic field.
For computing this exponent, we have used the hourly
averaged data obtained from Advanced Composition Ex-
plorer (ACE) satellite magnetometers for the last so-
lar cycle [26]. As it was expected the cluster-exponent
α ≃ 0.37 ± 0.02. Analogous result was obtained for other
components of the interplanetary magnetic field as well.
At the Earth itself the solar wind induced activity is
measured by geomagnetic indexes such as aa-index (in
units of 1 nT). This, the most widely used long-term ge-
omagnetic index [27], presents long-term geomagnetic ac-
tivity and it is produced using two observatories at nearly
antipodal positions on the Earth's surface. The index is
computed from the weighted average of the amplitude
of the field variations at the two sites. It was recently
discovered that variations of this index are strongly cor-
related with the global temperature anomalies (see, for
instance, [28],[29]).
In this paper, however, we will be
mostly interested in the clustering properties of the aa-
index and their relation to the modulation produced by
the photospheric convective motion. The point is that
the data for the aa-index are available for both the Mod-
ern and the Historic periods. This allows us to check the
suggestion that the Rayleigh number of the photospheric
convection has the same order for the both mentioned
periods. The transition between the two periods can be
seen in Figure 6, where we show the cumulated aa-index
a(t)
Saa(t) = Z t
0
a(t′)dt′
(2)
The arrow in this figure indicates beginning of the tran-
sitional solar cycle.
Figure 7 shows cluster-exponents for both the Modern
and the Historic periods calculated for the low intensity
levels of the aa-index (for the Historic period the level
used in the calculations is aa-index=10nT, whereas for
the Modern period the level is aa-index=25nT, the aa-
index was taken from [30]). The low levels of intensity
were chosen in order to avoid effect of the extreme phe-
nomena (magnetic storms and etc.). The straight line in
Fig. 7 is drawn to show the expected value of the cluster-
exponent α ≃ 0.37 for the both periods. It means that
indeed for both the Historic and the Modern periods the
Rayleigh number Ra ∼ 1011 (cf. Fig. 4) in the solar
photosphere.
IMPACT OF THE SOLAR DYNAMICS ON
EARTH CLIMATE
The transitions of the solar convection zone dynamics
can affect the earth climate through (at least) two
channels. First channel is a direct change in the heat
and light output of the Sun, especially during the
)
l
C
0
(
y
a
m
o
n
a
e
r
u
t
a
r
e
p
m
e
T
T (land and marine)
CO2 (Law Dome)
B
A
0.1
0
-0.1
-0.2
-0.3
-0.4
-0.5
-0.6
N
S
S
340
330
320
310
300
290
280
)
m
p
p
(
2
O
C
1850 1870 1890 1910 1930 1950 1970
Year
FIG. 8: Global temperature anomaly (solid curve, combined
land and marine) [20] and atmospheric CO2 (circles) [21] vs
time. Relevant daily SSN are also shown.
)
m
p
p
(
2
O
C
308
306
304
302
300
298
296
294
1900-1930yy (Law Dome)
1977- 2008yy (Mauna Loa)
0
0
5
5
10
10
15
15
20
20
25
25
30
30
t [y]
390
380
370
360
350
340
330
35
35
FIG. 9: Comparison of the CO2 growth for the periods 1900-
1930yy [21] and 1977-2008yy [22]. The solid straight lines
(best fit) indicate a linear growth.
5
)
C
0
(
y
l
a
m
o
n
a
e
r
u
t
a
r
e
p
m
e
T
1975
0
1980
1985
1990
1995
2000
2005
2010
0.5
1912-1942yy
combined land and marine
-0.1
-0.2
-0.3
-0.4
0.4
0.3
0.2
0.1
0
1977- 2007yy
-0.5
1910
1915
1920
1925
1930
1935
1940
-0.1
1945
Year
FIG. 10: Comparison of the temperature anomaly (combined
land and marine) [20] for the periods 1912-1942yy and 1977-
2007yy.
8 shows global temperature anomaly (the solid curve)
[20] and atmospheric CO2 mixing ratios (the circles)
[21] vs.
time. Relevant SSNs are also shown. The
dashed straight line (corresponding to the peak B)
separates between Historic and Modern periods. One
can see, that the transitional solar cycle (1933-1944yy)
is characterized by dramatic changes both in the global
temperature (a huge peak) and in the atmospheric
CO2 (complete suppression of the growth).
After
the transitional period one can observe (during the
three solar cycles: 1944-1977yy) certain growth in the
temperature anomaly and an unusually fast growth in
the atmospheric CO2. These three solar cycles seems
to be an aftershock adaptation of the global climate to
the new conditions. Then, starting from 1977 year, the
atmospheric CO2 growth returns to its linear trend as
before the transition B, but now the rate of the growth
is about four times larger than before the transition (see
Fig. 9). On the other hand, it can be seen from Fig. 1
that the temperature anomaly before the maximum C
returns to about the same growth pattern as it was just
before the transition B. A quantitative comparison of
these patterns, shown in Figure 10, indicates only about
20% difference in the growth rate (cf. also Refs.
[31],
[32]). The thirty year periods used for this comparison
could find a support in the Appendix B.
transitional solar cycle 1933-1944yy. Second channel
is related to the strong increase of the magnetic field
output in the interplanetary space through the sunspots.
The interplanetary magnetic field interacts with the
cosmic rays. Therefore, the change in the magnetic
field intensity can affect the Earth climate through the
change of the cosmic rays intensity and composition
(see,
[17],[18],[19]). Let us look
more closely to the transition B in Fig. 1. The Fig.
for instance, Refs.
If one consider the impact mechanism, related to the
energetic cosmic (charged) particles and the correspond-
ing screen effect of the magnetic fields, one could expect
that this impact mechanism should be also sensitive to
the 11-year solar cycle variability (though considerably
less than to the transitional effects). To find fingerprints
of this variability in the global temperature anomaly data
is not a trivial task due to comparatively (to the 11-
year cycle) short period of observations and due to the
11-year cycle
0.025
0.02
0.015
0.01
0.005
E
global temperature anomaly
sunspot number
0.1
0.2
0.3
0.4
0.5
0
0
f [y-1]
0.03
0.025
0.02
E
0.015
0.01
0.005
0
0
6
11- year cycle
AR spectrum
0.1
0.2
0.3
0.4
0.5
f [y-1]
FIG. 11: The solid curve corresponds to energy spectrum
of the fluctuating (noise-like) component of the decomposed
global temperature anomaly time series (combined land and
marine [20]). The dashed curve corresponds to energy spec-
trum of fluctuating component of the SSN data.
FIG. 12: Energy spectrum of the AR (fluctuating) component
of the decomposed global temperature anomaly time series
(combined land and marine [20]).
statistically non-stationary character of these data (with
considerable time trends). To extract these fingerprints
we will use a time series decomposition method devel-
oped in [33],[34],[35]. A time series is decomposed into a
trend (like that shown in Fig. 1 as the solid curve) and
a fluctuating (noise-like) components (see Appendix A).
The method applies state space modeling and Kalman
filter. Parameters are estimated by maximum likelihood
method. Obtained by this way the fluctuating (noise-
like) component of the (filtered) signal presents a sta-
tistically stationary time series and, therefore, it can be
subject of a spectral analysis. Result of this analysis
for the global temperature anomaly (combined land and
marine [20]) is shown in Figure 11. Although we have
deal with a comparatively short data series one can clear
see the main spectral peak corresponding to the (about)
11-year solar cycle impact (cf. Refs.
If
one also introduces an autoregressive (AR) component
into the time series decomposition (see Appendix A) to
describe the fluctuating part of the global temperature
anomaly, then one can compute energy spectrum of the
AR component. Such spectrum is shown in figure 12.
One can clear see the spectral peak corresponding to the
solar 11-year cycle impact.
[18],[37],[38]).
It is also interesting to look at the reduction in arc-
tic ice extent. Figure 13 shows a trend component of
the anomalies of ice August extent in the Chukchi sea
(the data are taken from [36]) obtained by the time se-
ries decomposition. The dashed straight line indicates
beginning of the transitional solar cycle. Insert to Fig.
13 shows spectrum of corresponding fluctuating (noise-
like) component of the (filtered) signal (cf. Fig. 11).
The dashed straight line indicates beginning of the tran-
sitional solar cycle. This arctic sea is located sufficiently
t
n
e
t
x
e
e
c
i
f
o
)
2
m
k
0
0
0
1
*
(
s
e
i
l
a
m
o
n
A
80
60
40
20
0
-20
-40
-60
1900
E
3000
2000
1000
0
11- year cycle
0
0.1
0.2
f
0.3
0.4
1933 y
1920
1940
1960
1980
2000
Year
FIG. 13: Trend of anomalies of ice August extent in the
Chukchi sea [36]. The insert shows spectrum of fluctuating
(noise-like) component of the time series decomposition.
far from the North Atlantic to diminish its influence,
whereas influence of the Pacific ocean on the Chukchi
sea local climate is not very significant.
A FORECAST
At present time we can already observe a peak C in the
global temperature anomaly (see Fig. 1). The last obser-
vation can be considered as an indication of the current
transition to a new section IV in the solar activity. The
fact that in the 'last' period (1977-2007yy) the temper-
ature anomaly growth returned to the same pattern as
just before the transition peak B (see Fig. 10) provides
an additional support to this suggestion.
)
l
C
0
(
y
a
m
o
n
a
e
r
u
t
a
r
e
p
m
e
T
0.6
0.4
0.2
0
-0.2
-0.4
-0.6
observed
ARMA predicted
I
II
III
IV
1860
1890
1920
1950
1980
2010
Year
250
200
150
100
50
0
N
S
S
7
11 year cycle
-5/3
0
-2
-4
-6
E
n
l
-8
-6
-5
-4
-3
ln f [y-1]
-2
-1
FIG. 14: The upper solid curve corresponds to the observed
yearly global temperature anomaly data (7-year smoothed)
[20]. The lower solid curve corresponds to the observed SSN
data (year-smoothed) [14]. The dashed curves are an ARMA-
prediction: for SSN and for global temperature anomaly.
FIG. 16: Spectrum of the galactic cosmic ray count rate fluc-
tuations in the ln-ln scales (the reconstructed data for pe-
riod 1611-2007yy have been taken from [44]). The straight
dashed line is drawn to indicate the Kolmogorov scaling law
E(f ) ∼ f (−5/3).
E
n
l
2
0
-2
-4
-6
-8
-10
-12
-6
11 - year cycle
-5/3
-5
-4
-3
ln f [y-1]
-2
-1
FIG. 15: ARMA spectrum of the global temperature anomaly
fluctuations in the ln-ln scales. The straight dashed line is
drawn to indicate the Kolmogorov scaling law E(f ) ∼ f (−5/3).
To make a forecast for the new section IV let us as-
sume that there was a storage of certain part of the en-
ergy coming from the radiative zone during the period of
the section II (Fig. 1). Then, in the period of the section
III this stored energy was released and transported to
the Sun surface. This release was not in the form of the
thermal convection, but in the form of strongly concen-
trated toroidal magnetic field structures, which results in
the intense emergence of the sunspots in the active parts
of the Modern period (section III in Fig. 1). It is inter-
esting that the storage and release periods have about
the same length: ≃ 60 years (see also Appendix B). The
storage of energy in a growing magnetic field at the bot-
tom of the convection zone or in the overshoot (by the
convective downflows) area just under the bottom can be
considered as the most plausible for this scenario. Con-
vective overshooting provides means to store magnetic
energy below the convection zone by resisting the buoy-
ancy effect for large time scales. Level of subadiabaticity
in the overshoot layer, needed in order to store the mag-
netic flux for sufficiently long times, can be significant.
Expected intensity of 105 G results an energy density
that is in order of magnitude larger than equipartition
value. Though, suppression of convective motion by the
magnetic flux tubes themselves can be considered as an
additional factor allowing to achieve the sufficient level
of subadiabaticity (see, for instance [3],[4],[23]).
Since the section III in the solar activity was a
it is naturally to assume that the next
release one,
section (IV) in the solar activity will be a 'storage'
one, similar to the 'storage' section II. As for the
corresponding Earth temperature anomaly, the section
IV is also expected to be similar to the section II. But
the temperature anomaly from the beginning of the
section IV will be shifted upward on about 0.6oC in
comparison with the section II. It then will be not
surprising if the total length of the section IV also will
be about 60 years (cf. Appendix B) and the section
will be finished with a maximum D higher than 0.6oC.
A very schematic presentation for a first part of the
section IV is shown in figure 14 (see Appendix A for an
Auto-Regressive-Moving-Average (ARMA) model used
for the prediction, cf. also Ref.
[39]). Figure 15 shows
corresponding ARMA model spectrum of the global
temperature anomaly fluctuations (in ln-ln scales). A
improving the paper. A software provided by K. Yosh-
ioka was used at the model computations.
11 year cycle
8
E
n
l
4
2
0
-2
-4
-6
-8
-10
-6
-5/3
0.4
0.2
0
-0.2
-0.4
-0.6
observed
ARM A predicted
1900
1940
1980
2020
-5
-4
-3
-2
-1
ln f
APPENDIX A
The above applied time series decomposition method
was developed in the Refs. [33],[34],[35]. In this method a
nonstationary mean time series is decomposed into trend,
autoregressive and noise components. A random process
Xt in discrete time t is defined by the following expres-
sion:
Xt = τt + εt + AR(m)
t
(A1)
where τt is a trend component, εt is an observational
noise component, and
AR(m)
t =
m
Xi=1
aiXt−i + η(1)
t
(A2)
is a globally stationary autoregressive (AR) component
of order m (ai are the coefficients of a recursive filter,
and η(1)
is a Gaussian white noise).
t
A k-th order stochastically perturbed difference equa-
tion describes evolution of the trend component
∆kτt = η(2)
t
(A3)
where η(2)
t
is a Gaussian white noise.
In the Ref. [35] algorithms, using state space modeling
and Kalman filter, are suggested for estimation of the
model parameters by the maximum likelihood method.
Using
so-called
autoregressive moving
average
(ARMA) model one could try to make a prediction. For
this model a random process Xt in discrete time t is
defined by the following expression:
Xt = εt + AR(m)
t +
n
Xi=1
biεt−i
(A4)
where εt
is Gaussian white noise with zero mean
[33],[34],[35]. Again, state space modeling and Kalman
filter, are used for estimation of the model parameters
by the maximum likelihood method. Figure 14 shows
an optimistic prediction produced with this model for
a first part of the section IV. Though, continuation of
the analogy to the section II makes the scenario much
less optimistic. A quasi-Newton method and dynamic
restrictions were used for optimization of the model pa-
rameters.
We can also play with the model by cutting the ob-
served time series. This procedure reduces relative con-
tribution of the low-frequency fluctuations, presumably
of Galactic origin (cf. Fig. 17 and Fig. 15). Then, the
insert to the Fig. 17 shows less steep decline of the global
FIG. 17: Spectrum of the global temperature anomaly fluc-
tuations in the ln-ln scales for a cutted observed base ARMA
model. The straight dashed line is drawn to indicate the Kol-
mogorov scaling law E(f ) ∼ f (−5/3).
dashed straight line in this figure indicates a scaling
with '-5/3' exponent: E(f ) ∼ f −5/3. Although, the
scaling interval is short, the value of the exponent is
rather intriguing. This exponent is well known in the
theory of fluid (plasma) turbulence and corresponds to
so-called Kolmogorov's cascade process (see, for instance
[40]). This process is very universal for turbulent fluids
and plasmas [41],[42]. For turbulent processes in Earth
and in Heliosphere the Kolmogorov-like spectra with
such large time scales cannot exist. Therefore, one
should think about a Galactic origin of Kolmogorov
turbulence (or turbulence-like processes [43]) with such
large time scales. This is not surprising if we recall
possible role of the galactic cosmic rays for Earth climate
(see, for instance,
In order to support
this point we show in figure 16 spectrum of galactic
cosmic ray intensity at the Earth's orbit (reconstruction
for period 1611-2007yy [44], cf. also Ref.
[24]). One
can compare Fig. 16 with Figs. 15 corresponding to
the global temperature anomaly fluctuations (see also
Appendix C). Presence of the turbulent-like component
in the low-frequency fluctuations makes any long-range
prediction a very difficult task (see also Appendix B on
a chaotic element in solar dynamics). A chaotic element
in global climate dynamics itself (Appendix C) presents
another problem for the forecast. Such internal forcing
agents as volcanoes and anthropogenic greenhouse gases,
for instance, have a potential to change drastically the
future dynamics of the global climate.
[17],[18],[19]).
The author is grateful to J.J. Niemela, to K.R. Sreeni-
vasan, and to SIDC-team, World Data Center for the
Sunspot Index, Royal Observatory of Belgium for shar-
ing their data and discussions. The author is also grateful
to C.H. Gibson and to R.A. Treumann for their help in
temperature anomaly at this (cutted observed base) ver-
sion of the ARMA model prediction. This game has also
another end. One can assume that using a more long
time series (observed base), than that we have used in
computing the results shown in Fig. 14, one could ob-
tain even more optimistic prediction for the first part of
the section IV.
APPENDIX B
The long-range reconstructions of the sunspot number
fluctuations (see, for instance, Refs.
[1],[2]) allow us to
look on the solar transitional dynamics from a more gen-
eral point of view. In figure 18 we show a spectrum of
such reconstruction for the last 11,000 years (the data,
used for computation of the spectrum,
is available at
[44]). The spectrum was computed using the autore-
gressive model (with maximum entropy method [35]). A
semi-logarithmical representation was used in the figure
to show an exponential law
E(f ) ∼ e−f /fe
(B1)
The straight line is drawn in Fig. 18 to indicate the expo-
nential law Eq. (B1). Slope of the straight line provides
us with the characteristic time scale Te = 1/fe ≃ 176±7y.
The exponential decay of the spectrum excludes the pos-
sibility of random behavior and indicates the chaotic be-
havior of the time series [45],[46]. It is well known that
low-order dynamic (deterministic) systems have as a rule
exponential decay of E(f ) (see, for instance, [46],[47]).
As for infinite dimensional dynamic systems with chaotic
attractors it is interesting to compare Fig. 18 with fig-
ure 3 of the Ref. [48]. It should be noted that the 176y
period is the third doubling of the period 22y. The 22y
period corresponds to the Sun's magnetic poles polarity
switching (see also Appendix C).
The exponential spectrum can be also produced by a
series of Lorentzian pulses with the average width of the
individual pulses equals to τ (though, the distribution of
widths of the pulses should be fairly narrow to result in
the exponential spectrum).
In Fig. 18 a local maximum corresponding to the fre-
quency fe and its first harmonics have been indicated
by arrows.
It should be noted that the harmonic 2fe
corresponds to the well known period T ≃ 88y (see, for
instance, Ref.
[49]), and the harmonic 3fe corresponds
to the period T ≃ 60y (cf. section "A FORECAST").
APPENDIX C
It is also useful to consider the above results in content
of multicentennial and millennial timescales. Figure 19
shows a reconstruction of Northern Hemisphere tempera-
tures for the past 2,000 years (the data for this figure were
9
8
6
4
2
0
-2
0
E~exp(-f/fe)
fe
2fe
3fe
0.1
0.2
0.3
0.4
f [(10y)-1]
)
f
(
E
n
l
FIG. 18: Spectrum of the sunspots number fluctuations in
the ln-linear scales (the reconstructed data for the last 11,000
years have been taken from [44]). The straight line is drawn
to indicate the exponential law Eq. (B1).
1-1980yy
0.5
0
-0.5
-1
)
C
0
(
l
y
a
m
o
n
a
e
r
u
t
a
r
e
p
m
e
T
-1.5
0
500
1000
Year
1500
2000
FIG. 19: A reconstruction of Northern Hemisphere temper-
ature anomaly for the past 2,000 years (the data was taken
from Ref. [50]).
taken from Ref.
[50]). This multi-proxy reconstruction
was performed by the authors of Ref. [51] using combina-
tion of low-resolution proxies (lake and ocean sediments)
with comparatively high-resolution tree-ring data. Fig-
ure 20 shows a power spectrum of the data set calculated
using the maximum entropy method (as for the spectra
shown in Figs. 16 and 18) in the frames of the autore-
gressive model, because it provides an optimal spectral
resolution even for small data sets (see also [47]). The
spectrum exhibits a rather wide peak indicating a peri-
odic component with a period around 22 y, and a broad-
band part with exponential decay. A semilogarithmical
plot was used in Fig. 20 (cf Fig. 18) in order to show
the exponential decay more clearly (at this plot the ex-
2
0
-2
-4
-6
)
f
(
E
n
l
22y
E~exp-f/fe
Te=1/fe=11y
0
0.05
0.1
0.15
f [y-1]
0.2
0.25
0.3
10
)
f
(
E
n
l
1
0
-1
-2
-3
-4
-5/3
22y
-6
-5
-4
ln f [y-1]
-3
-2
FIG. 20: Spectrum of the data, shown in Fig. 19, in semilog-
arithmical scales. The straight line indicates the exponential
decay Eq. (B1).
FIG. 22: Spectrum of the data, shown in Fig. 19, in ln-ln
scales. The dashed straight line indicates the Kolmogorov-
like spectrum: E(f ) ∼ f −5/3. The high-frequency part has
been cutted in order to show the low-frequency part.
)
f
(
E
n
l
10
5
0
-5
-10
-15
-20
0
0.1
0.2
0.3
0.4
0.5
f
FIG. 21: Spectrum of the chaotic fluctuations of the x- com-
ponent for the Rossler system (a = 0.15, b = 0.20, c = 10.0.)
ponential decay corresponds to a straight line). Both
stochastic and deterministic processes can result in the
broad-band part of the spectrum, but the decay in the
spectral power is different for the two cases. The ex-
ponential decay indicates that the broad-band spectrum
for these data arises from a deterministic rather than a
stochastic process. Indeed, for a wide class of determin-
istic systems a broad-band spectrum with exponential
decay is a generic feature of their chaotic solutions (see,
for instance, Refs. [45],[47],[48] and Appendix B).
In order to illustrate this and another significant fea-
ture of the chaotic power spectra we show in figure 21 a
power spectrum for the Rossler system [52]
dx
dt
= −(y + z);
dy
dt
= x + ay;
dz
dt
= b + xz − cz (C1)
chaotic solution, where a, b and c are parameters.
In
this figure one can see a typical picture: a narrow-band
peak (corresponding to the fundamental frequency of the
system) in a low-frequency part and a broad-band expo-
nential decay in a high-frequency part of the spectrum.
Nature of the exponential decay of the power spectra
of the chaotic systems is still an unsolved mathemat-
ical problem. A progress in solution of this problem
has been achieved by the use of the analytical contin-
uation of the equations in the complex domain (see, for
instance, [53],[54]). In this approach the exponential de-
cay of chaotic spectrum is related to a singularity in the
plane of complex time, which lies nearest to the real axis.
Distance between this singularity and the real axis deter-
mines the rate of the exponential decay. If parameters of
the dynamical system periodically fluctuate around their
mean values, then at certain (critical) intensity of these
fluctuations an additional singularity (nearest to the real
time axis) can appear. Distance between this singularity
and the real axis is determined by period of the peri-
odic fluctuation of the system's parameters. Therefore,
exponential decay rate of the broad-band part of the sys-
tem spectrum equals the period of the parametric forcing.
The chaotic spectrum provides two different characteris-
tic time-scales for the system: a period corresponding to
fundamental frequency of the system, Tf un, and a period
corresponding to the exponential decay rate, Te = 1/fe
(cf Eq. (B1)). The fundamental period Tf un can be esti-
mated using position of the low-frequency peak, while the
exponential decay rate period Te = 1/fe can be estimated
using the slope of the straight line of the broad-band part
of the spectrum in the semilogarithmical representation
(Figs. 20 and 21). From Fig. 20 we obtain Tf un ≃ 22±2y
and Te ≃ 11 ± 1y (the estimated errors are statistical
ones). Thus, the solar activity (SSN) period of 11 years
is still a dominating factor in the chaotic temperature
11
fluctuations at the millennial time scales , although it is
hidden for linear interpretation of the power spectrum
(cf Ref.
[55]). In the nonlinear interpretation the addi-
tional period Tf un ≃ 22y might correspond to the funda-
mental frequency of the underlying nonlinear dynamical
system. It is surprising that this period is close to the
22y period of the Sun's magnetic poles polarity switching
(cf Appendix B). It should be noted that the authors of
Ref.
[56] found a persistent 22y cyclicity in sunspot ac-
tivity, presumably related to interaction between the 22y
period of magnetic poles polarity switching and a relic
solar (dipole) magnetic field. Therefore, one cannot rule
out a possibility that the broad peak, in a vicinity of fre-
quency corresponding to the 22y period, is a quasi-linear
response of the global temperature to the weak periodic
modulation by the 22y cyclicity in sunspot activity. I.e.
strong enough periodic forcing results in the non-linear
(chaotic) response whereas a weak periodic forcing results
in a quasi-periodic response.
Finally, figure 22 shows the same power spectrum as in
Fig. 20 but in ln-ln scales, where the dashed straight line
indicates the Kolmogorov-like behavior: E(f ) ∼ f −5/3,
in the lower-frequency part of the spectrum (see section
'A Forecast' and Fig. 16 for relation between this type
of spectrum and galactic turbulence).
[12] G. Molchan, private communication.
[13] M.R. Leadbetter and J.D. Gryer, The variance of the
number of zeros of a stationary normal process, Bull.
Amer. Math. Soc. 71, 561-563 (1965).
[14] The
data
are
available
at
http://sidc.oma.be/sunspot-data/
[15] R.J. Bray, R.E. Loughead, and C.J Durrant, The so-
lar granulation (Cambridge Univ. Press, Cambridge, 2ed,
1984).
[16] J.J. Niemela, L. Skrbek, K.R. Sreenivasan, R. J. Don-
nelly, Turbulent convection at very high Rayleigh num-
bers, Nature, 404, 837-840 (2000).
[17] I. G. Usoskin and G. A. Kovaltsov, Cosmic Ray Induced
Ionization in the Atmosphere: Full Modeling and Practi-
cal Applications, J. Geophys. Res., 111, D21206, (2006).
[18] N.J. Shaviv, On climate response to changes in the cos-
mic ray flux and radiative budget, J. Geophys. Res. 110,
A08105 (2005).
[19] J. Kirkby, Cosmic Rays and Climate, Surveys in Geo-
physics, 28, 333-375 (2007) (see also arXiv:0804.1938).
[20] The data are available at http://www.cru.uea.ac.uk/cru/
data/temperature/ (N.A. Rayner, et al., Improved anal-
yses of changes and uncertainties in marine temperature
measured in situ since the mid-nineteenth century: the
HadSST2 dataset, J. Climate, 19, 446-469 (2006)).
[21] The data are available at http://cdiac.ornl.gov/trends/
co2/lawdome-data.html (D.M. Etheridge, et al., Histori-
cal CO2 records from the Law Dome DE08, DE08-2, and
DSS ice cores).
[22] The data
are
available
at http://cdiac.ornl.gov/
ftp/trends/co2/maunaloa.co2 (C. D. Keeling, and T.P.
Whorf. Atmospheric CO2 records from sites in the SIO
air sampling network).
[1] S. K. Solanki, I. G. Usoskin, B. Kromer, M. Schussler
and J. Beer, Unusual activity of the Sun during recent
decades compared to the previous 11,000 years, Nature,
431, 1084-1087 (2004).
[2] I.G. Usoskin, S.K. Solanki, M. Schussler,
et al.,
Millennium-Scale Sunspot Number Reconstruction: Ev-
idence
the
1940s, Phys. Rev. Lett. 91, 211101 (2003) (see also
arXiv:astro-ph/0310823).
an Unusually Active Sun since
for
[3] E.A. Spiegel, and N.O. Weiss, Magnetic activity and vari-
ations in solar luminosity, Nature, 287, 616-617 (1980)
[4] A. Brandenburg, The Case for a Distributed Solar Dy-
namo Shaped by Near-Surface Shear, ApJ, 625 539-547
(2005) (see also arXiv:astro-ph/0502275).
[5] H.W. Babcock, The topology of the suns magnetic field
and the 22-year cycle, ApJ. 133 572-589 (1961).
[6] R.B. Leighton, A Magneto-Kinematic Model of the Solar
Cycle, ApJ, 156, 1-26 (1969).
[7] M. Dikpati, G.de Toma, and P.A. Gilman, Predicting the
strength of solar cycle 24 using a flux-transport dynamo-
based tool, Geophys. Res. Lett., 33 L05102 (2006).
[8] A.R. Choudhuri, P. Chatterjee, and J. Jiang, Predicting
Solar Cycle 24 With a Solar Dynamo Model, Phys. Rev.
Lett. 98, 131101 (2007).
[9] R.J. Bray, R.E. Loughhead, Sunspots, (Dover Publica-
[23] S.M. Tobias and D.W. Hughes, The Influence of Veloc-
ity Shear on Magnetic Buoyancy Instability in the Solar
Tachocline, ApJ. 603, 785802 (2004).
[24] A. Bershadskii, Multiscaling of Galactic Cosmic Ray
Flux, Phys. Rev. Lett., 90 041101 (2003) (see also
arXiv:astro-ph/0305453).
[25] M.L. Goldstein, Major Unsolved Problems in Space
Plasma Physics, Astrophys. Space Sci. 227, 349369
(2001).
[26] The data are available at http://www.srl.caltech.edu/
ACE/ASC/
[27] P.N. Mayaud, Derivation, Meaning, and Use of Geo-
magnetic Indices, AGU 1129 Geophys. Monograph 22,
(Washington D.C., 1980).
[28] E. W. Cliver, V. Boriakoff, J. Feynman, Solar Variabil-
ity and Climate Change: Geomagnetic aa Index and
Global Surface Temperature, J. Geophys. Res. Lett. 25,
10351038 (1998).
[29] T. Landscheidt, in "The solar cycle and terrestrial cli-
mate"', Vazquez, M. and Schmiedere, E, ed., European
Space Agency, Special Publication, 463, pp 497-500
(2000).
[30] The data are available at http://www.ukssdc.ac.uk/
data/wdcc1/wdc menu.html (World Data Centre for
Solar-Terrestrial Physics, Chilton).
tions, New York, 1979).
[31] C. Fabara
[10] K.R. Sreenivasan and A. Bershadskii, Clustering Proper-
ties in Turbulent Signals, J. Stat. Phys., 125, 1141-1153
(2006).
[11] A. Bershadskii, Transitional dynamics of the solar con-
vection zone, Europhys. Lett., 85, 49002 (2009).
and B. Hoeneisen, Global Warm-
calculations,
back-of-the-envelope
ing:
arXiv:physics/0503119 (2005).
some
[32] G. Gerlich and R. D. Tscheuschner, Falsification Of The
Atmospheric CO2 Greenhouse Effects Within The Frame
Of Physics, arXiv:0707.1161 (2007).
12
[33] G. Kitagawa, and W. Gersch, Smoothness Priors Analy-
(2002)).
sis of Time Series. (Springer, 1996).
[34] P.J. Brockwell, and R.A. Davis, Introduction to Time
Series and Forecasting. (Springer, 1996).
[35] G. Kitagawa, Time Series Analysis Programming,
Iwanami Shoten. (Tokyo, 1993, in Japanese).
[36] The data are available at http://www.frontier.iarc.
uaf.edu:8080/ igor/research/ice/icedata.php (see also I.
Polyakov, et al., Long-Term Ice Variability in Arctic
Marginal Seas, Journal of Climate 16, 2078-2085 (2003)).
[37] N. Scaffeta and B.J. West, Is climate sensitive to solar
variability? , Physics Today, 51, 50-51 (2008).
[38] N. Scaffeta and B.J. West, Estimated solar contribu-
tion to the global surface warming using the ACRIM
TSI satellite composite, Geophys. Res. Lett., 33, L05708
(2006) (see also arXiv:physics/0509248).
[39] The data are available at http://www.sec.noaa.gov/ So-
larCycle/SC24.
[45] D.E. Sigeti, Survival of deterministic dynamics in the
presence of noise and the exponential decay of power
spectrum at high frequencies. Phys. Rev. E, 52, 2443-
2457 (1995).
[46] J. Sun, Y. Zhao, T. Nakamura, and M. Small, From phase
space to frequency domain: A time-frequency analysis for
chaotic time series, Phys. Rev. E 76, 016220 (2007).
[47] N. Ohtomo, K. Tokiwano, Y. Tanaka, A. Sumi, S.
Terachi, and H. Konno, Exponential Characteristics of
Power Spectral Densities Caused by Chaotic Phenomena,
J. Phys. Soc. Jpn. 64 1104-1113 (1995).
[48] J. D. Farmer, Chaotic attractors of an infinite dimen-
sional dynamic system, Physica D, 4, 366-393 (1982).
[49] J. Feynman, and S.B. Gabriel, Period and phase of the
88-year solar cycle and the Maunder Minimum: Evidence
for a chaotic Sun, Solar Physics, 127, 393-403 (1990).
[50] The data are available at http://www.ncdc.noaa.gov/
[40] A.C. Monin and A.M. Yaglom, Statistical Fluid Mechan-
paleo/metadata/noaarecon-6267.html
ics (MIT Press, Cambridge, 1975), Vol. 2.
[41] C.H Gibson, Kolmogorov Similarity Hypotheses for
Scalar Fields: Sampling Intermittent Turbulent Mixing
in the Ocean and Galaxy, Proc. Roy. Soc. Lond. 434, 149
(1991).
[42] J. Cho, A. Lazarian, and E.T. Vishniac, Simula-
tions of MHD Turbulence in a Strongly Magnetized
Medium, Astrophys. J. 564, 291-301 (2002) (see also
arXiv:astro-ph/0205286).
[43] C.H. Gibson, R.N. Keeler, V.G. Bondur,
et al.,
Submerged turbulence detection with optical
satel-
lites, Proc. of SPIE, 6680, 6680-33 (2007) (see also
arXiv:0709.0074v2 [astro-ph]).
[44] The data are available at http://www1.ncdc.noaa.gov/
pub/data/paleo/climate forcing/solar variability/usoskin-
cosmic-ray.txt (see also I.G. Usoskin, K. Mursula, S.K.
Solanki, M. Schuessler, and G.A. Kovaltsov, A physical
reconstruction of cosmic ray intensity since 1610. J.
Geophys. Res., J. Geophys. Res., 107(A11), 1374-1380
[51] A. Moberg, D. M. Sonechkin, K. Holmgren, N. M. Dat-
senko and W. Karlen, Highly variable Northern Hemi-
sphere temperatures reconstructed from low- and high
resolution proxy data Nature, 433, 613-617 (2005).
[52] O.E. Rossler, An equation for continuous chaos, Phys.
Lett. A, 35, 397-398 (1976).
[53] F. Fucito, F. Marchesoni, E. Marianari, G. Parisi,
L.Peliti, S. Ruffo, and V. Vulpiani, Approach to equi-
librium in a chain of nonlinear oscillators, J. Physique,
43, 707-713 (1982).
[54] U. Frisch and R. Morf, Intermittency in non-linear dy-
namics and singularities at complex times, Phys. Rev.
23, 2673 (1981).
[55] A. Bershadskii, Chaotic dynamics of atmospheric CO2 at
millennial timescales, arXiv:0903.2795 (2009).
[56] K. Mursula, I. G. Usoskin, and G. A. Kovaltsov, Per-
sistent 22-year cycle in sunspot activity: Evidence for a
relic solar magnetic field, Solar Phys., 198, 51-56, 2001.
|
astro-ph/9804297 | 1 | 9804 | 1998-04-28T11:46:11 | Towards the absolute planes: a new calibration of the Bolometric Corrections and Temperature scales for Population II Giants | [
"astro-ph"
] | We present new determinations of bolometric corrections and effective temperature scales as a function of infrared and optical colors, using a large database of photometric observations of about 6500 Population II giants in Galactic Globular Clusters (GGCs), covering a wide range in metallicity (-2.0<[Fe/H]<0.0). New relations for BC_K vs (V-K), (J-K) and BC_V vs (B-V), (V-I), (V-J), and new calibrations for T_eff, using both an empirical relation and model atmospheres, are provided. Moreover, an empirical relation to derive the R parameter of the Infrared Flux Method as a function of the stellar temperature is also presented. | astro-ph | astro-ph |
Mon. Not. R. Astron. Soc. 000, 000 -- 000 (1997)
Printed 31 May 2021
(MN LATEX style file v1.4)
Towards the absolute planes: a new calibration of the
Bolometric Corrections and Temperature scales
for Population II Giants ⋆
P. Montegriffo,1 F.R. Ferraro,1 L. Origlia,1 and F. Fusi Pecci2†
1Osservatorio Astronomico di Bologna, Via Zamboni 33, 40126 Bologna, ITALY
2Stazione Astronomica, 09012 Capoterra, Cagliari, ITALY
31 May 2021
ABSTRACT
We present new determinations of bolometric corrections and effective temperature
scales as a function of infrared and optical colors, using a large database of photometric
observations of about 6500 Population II giants in Galactic Globular Clusters (GGCs),
covering a wide range in metallicity ( -- 2.0<[Fe/H]<0.0).
New relations for BCK vs (V -- K), (J -- K) and BCV vs (B -- V), (V -- I), (V -- J), and new
calibrations for Tef f , using both an empirical relation and model atmospheres, are
provided.
Moreover, an empirical relation to derive the R parameter of the Infrared Flux Method
as a function of the stellar temperature is also presented.
Key words: Clusters: Globular -- Stars: Evolution -- Stars: Fundamental Parameters
-- Stars: Hertzsprung -- Russell diagram -- Photometry: IR -- Array
1
INTRODUCTION
A global test of stellar evolutionary models requires a direct
comparison between theoretical tracks and observations for
stars spanning a wide range in stellar parameters, such as
temperature, luminosity and metallicity. In order to achieve
these goals at least two fundamental ingredients are needed:
i) a complete and homogeneous database of photomet-
ric observations;
ii) a suitable set of transformations between observ-
ables and absolute quantities.
GGCs are the best empirical laboratory to obtain com-
plete and homogeneous spectrophotometric information on
Pop. II stars over a wide range of metallicities.
Rewieving the published works on the transformations
to the absolute plane (see Sect.3), that is bolometric correc-
tions (BCs) and temperature scales as a function of differ-
ent colors, it is easy to see that very often these calibrations
are not based on a complete and homogeneous set of data,
spanning a wide range of stellar parameters. Moreover, ad
hoc correction factors are usually adopted to take into ac-
count for example possible systematic differences between
different photometric systems and/or different assumptions
⋆ Based on data taken at the ESO -- MPI 2.2m Telescope equipped
with the near IR camera IRAC2 - ESO, La Silla (Chile).
† on leave from Osservatorio Astronomico di Bologna
c(cid:13) 1997 RAS
for the reference solar quantities, for the adopted model at-
mospheres, and for different laws to extrapolate the data
towards the UV/IR ranges.
Such a scenario indicates that any calibration in the
absolute plane can hardly be fully self -- consistent as it always
depends on the adopted transformations and, more crucial,
the residuals among different scales are very rarely linear
with the involved parameters.
In order to improve the available determinations of
the bolometric corrections and temperature scales, we use
here our IR photometric database on GGC stars combined
with available optical data from the literature to calibrate
new, independent, (hopefully) self -- consistent transforma-
tions, particularly useful to study the red stellar sequences
in Pop II stars.
In Sect.2 we present the complete database used in our
analysis which includes about 6500 RGB and HB stars in
a sample of 10 GGCs observed in both optical and near
IR bands. In Sect.3 we derive the transformations from ob-
served magnitudes and colors to absolute quantities, such
as bolometric corrections and effective temperatures. All
the results are listed in Table 3. In Sect.4 we compare the
inferred scales with existent ones and we give a fully em-
pirical calibration of the R parameter of the Infrared Flux
Method (IRFM) as a function of the effective temperature.
Schematic conclusions are eventually presented in Sect.5.
2
P. Montegriffo et al.
2 THE DATABASE
The IR database used in this study includes photometric ob-
servations in J and K bands of about 17000 stars belonging
to 10 GGCs spanning the whole range in metallicity (from
[Fe/H]= -- 2.15 to [Fe/H]≈0.0). The data were obtained at
ESO, La Silla (Chile), during two different runs (on June
1992 and June 1993), using the ESO -- MPI 2.2m telescope
and the near -- IR camera IRAC -- 2 (Moorwood et al. 1992)
equipped with a NICMOS -- 3 256x256 array detector. The
complete description of the observations and data reduction
can be found in Ferraro et al. (1994a,b) and Montegriffo et
al. (1995) and in a series of forthcoming papers, where each
individual cluster is discussed in more detail.
Our IR database was cross-correlated with other optical
and IR catalogs in the literature in order to provide complete
UBVRIJK photometry of as many stars as possible. The
final sample used in the following analysis includes about
6500 RGB and HB stars.
The reddening correction in each photometric band was
performed starting from the E(B -- V) color excess (as re-
ported by Armandroff (1989) with the exception of M68
and M69 for which we used the values by Walker 1994 and
Ferraro et al. 1994a, respectively) and using the extinction
law proposed by Rieke & Lebovsky (1985). In Table 1 the
main features of the database are listed: the GGC names, the
cluster metallicity as quoted by Zinn (1985), the available
photometric bands, and the adopted E(B -- V) color excess.
In the following the observed colors are always corrected for
reddening.
2.1 Notes on individual clusters
M15: UBVR photometry from Stetson (1994).
M30: UBV photometry from Bergbusch (1996).
M68: BVI photometry from Walker (1994).
M55: BV photometry from Piotto (1996) and VI from Or-
tolani & Desidera (1996): V was averaged.
M4: BV photometry from Lee (1977).
M107: BV photometry from Ferraro et al. (1991).
M69: BV photometry from Ferraro et al. (1994a).
47Tuc: UBV photometry from Auriere & Lauzeral (1996)
and VI from Ortolani & Desidera (1996). The VI photom-
etry was compared with that one obtained by Da Costa &
Armandroff (1990) for the few stars in common. There is
an excellent agreement in V, while a systematic shift in the
zero point by about 0.12 mag in I has been found: we ap-
plied this correction to the I photometry from Ortolani &
Desidera (1996).
NGC6553: VI photometry from Guarnieri et al. (1997) us-
ing HST data.
NGC6528: VI photometry from Ortolani et al. (1995) using
HST data.
3 ABSOLUTE QUANTITIES
Various calibrations of the bolometric correction and tem-
perature scale have been proposed by different authors since
many years.
Table 1. Adopted IR and optical database.
Cluster
NGC7078 M15
NGC7099 M30
NGC4590 M68
NGC6809 M55
NGC6121
M4
NGC6171 M107
NGC6637 M69
NGC104
47Tuc
NGC6553
NGC6528
[Fe/H]a
-- 2.15
-- 2.13
-- 2.09
-- 1.82
-- 1.19
-- 0.89
-- 0.75
-- 0.70
-- 0.29
-- 0.07
Photometryb
E(B -- V)c
UBVRJK
UBVJK
BVIJK
BVIJK
BVJK
BVJK
BVJK
UBVIJK
VIJK
VIJK
0.10
0.04
0.07
0.06
0.40
0.31
0.17
0.04
0.78
0.56
a From Zinn (1985).
b J,K from our IR -- array survey, U,B,V,R,I from the literature
(refers to Sect.2.1 for the bibliographic sources).
c E(B -- V) from Armandroff (1989), but M68 from Walker
(1994) and M69 from Ferraro et al. (1994a), respectively.
Johnson (1966) used a sample of 15 giants observed in
the UBVRIJKLMN to get bolometric corrections in the V
band, integrating the spectral energy distribution and apply-
ing small corrections (<0.1 mag) in the wavelength ranges
not covered by his survey (such as for example the H band)
to take into account possible molecular blending in cool
stars. The zero point is defined as BC⊙
V =0.00.
Carney & Aaronson (1979) reconstructed the energy
distribution of a sample of dwarfs and subdwarfs by means
of polynomial fits to the optical photometry, while the UV
range was extrapolated using the Kurucz (1979) models and
the IR region using blackbodies. Their zero point is BC⊙
V = --
0.12.
Frogel, Persson & Cohen (1981, hereafter FPC81) per-
formed trapezoidal integrations of the energy distribution of
a sample of stars observed in the UBVRIJHKL, and extrap-
olating the UV and IR fluxes. They adopted as zero point
the value BC⊙
V = -- 0.08.
More recently, Tinney et al. (1993) reconstructed the
spectral energy distribution of cool stars combining pho-
tometric data in the VIJHKLL' bands with low resolution
spectra.
Alonso et al. (1995) determined the bolometric correc-
tions in the K band of a sample of F,G,K dwarfs. They
performed trapezoidal integrations using UBVRIJHK pho-
tometric data and included a correction factor C = f (Tef f ,
log g, [Fe/H]) to the UV and IR fluxes as a function of the
stellar parameters, using the Kurucz (1993) models.
Concerning the determination of the stellar effective
temperature, the only pure experimental way to derive it
is to know the star intrinsic luminosity and apparent angu-
lar diameter. Ridgway et al. (1979) for example measured
the angular diameter of a sample of nearby Pop I giants by
means of lunar occultations, while Di Benedetto & Rabbia
(1987, hereafter DBR87) and Di Benedetto (1993, hereafter
DB93) used interferometric techniques.
Once the apparent stellar diameters have been mea-
sured, the effective stellar temperatures can be estimated
assuming a proper scale of bolometric corrections. DBR87
have shown how an error of 0.05 mag in the determination
of the BCK transfers into an error of ≈60 K in the tem-
perature estimates for Tef f around 5000 K and of ≈35 K
for Tef f ≈3000 K. These uncertainties are of the same or-
c(cid:13) 1997 RAS, MNRAS 000, 000 -- 000
Bolometric Corrections and Temperature scales for Population II giants
3
Table 2. Absolute flux calibrations of a zero magnitude star in UBVRIJHKL
bands for different photometric systems. Wavelengths are in µm and fluxes in
units of 10−8 erg s−1 cm−2 µm−1.
Filter
U
B
V
R
I
J
H
K
L
Johnsona
λef f
0.36
0.44
0.55
0.70
0.90
1.25
1.62
2.20
3.50
fλ0
4345
7194
3917
1762
829.9
289.7
107.9
38.02
7.834
Bessell b
λef f
0.366
0.438
0.545
0.641
0.798
1.221
1.632
2.187
3.451
fλ0
4175
6320
3631
2177
1126
314.7
113.8
39.61
7.080
ESOc
λef f
0.363
0.438
0.548
0.641
0.795
1.244
1.634
2.190
3.770
fλ0
4080
6490
3650
2190
1180
312
120
41.7
5.41
a Johnson system revised by Buzzoni (1996).
b Bessell system adopted in BCP97 (Castelli 1997).
c UBVRI from Bessell (1990) and from Megessier (1995).
der of magnitude as the empirical errors, hence an accurate
estimate of BCK is an important issue and severely affects
the overall accuracy of the temperature estimated using this
technique.
Another method which allows one to derive stellar tem-
peratures in more distant objects is represented by the so --
called IRFM proposed by Blackwell, Shallis & Selby (1979)
and Blackwell & Lynas -- Gray (1994).
This method uses as a fundamental parameter the quan-
tity R = fbol/fλ, where fbol is the observed bolometric flux
and fλ is the observed flux in a certain photometric band.
Such a quantity is compared with the corresponding theo-
retical one, derived by means of models of atmospheres, in
order to derive a temperature scale. Tinney et al. (1993) for
their sample of dwarfs (see above) derived Tef f by subse-
quent iterations of the IRFM.
Finally, a fully theoretical approach is the extensive use
of model atmospheres to derive synthetic colors for different
input stellar parameters.
Before describing the adopted procedures to calibrate
our relations, we want to add a further comment on a cru-
cial point: each absolute flux calibration requires a reference
zero -- magnitude star in all the explored photometric bands,
that is in the specific photometric system used. As pointed
out by many authors (see e.g. Bessell & Brett 1988; Buz-
zoni 1989; Bessell 1990, Megessier 1995) the photometric
CCD+filter characterization is often provided with a quite
large uncertainty (typically ≈5%), and this may severely af-
fect the accuracy of the derived quantities, particularly the
bolometric fluxes, based on this instrumental calibration.
Moreover, it is difficult to reconstruct the adopted values
from the various authors since very often they are not tab-
ulated. Since in the calibrations of the present paper we
worked with different photometric systems, in Table 2 we
report all the adopted zero magnitude fluxes.
3.1 The determination of bolometric corrections
3.1.1 The BCK vs (V -- K) and (J -- K)
For all the observed stars in the selected sample of GGCs
we computed the bolometric correction BCK in the K band
c(cid:13) 1997 RAS, MNRAS 000, 000 -- 000
by means of the relation:
BCK = mbol − K
where mbol is the apparent bolometric magnitude.
mbol was computed as follows: first we convert the ap-
parent magnitudes corrected for reddening into fluxes using
the classical formula:
fλ = fλ0 · 10−0.4·mλ
where fλ0 is the absolute flux in a certain photometric band
corresponding to the zero magnitude. We used the UBVJK
filters of Johnson and RI of Cousins in the ESO standard
photometric system. In Table 2 we report the absolute flux
calibrations of a zero magnitude star for all the observed
photometric bands, for different photometric systems. The
values we adopted in our analysis are those listed in columns
#6 and #7.
The energy distribution was reconstructed starting from
a set of Planck functions Bλ(Tc) computed between two ad-
jacent observed bands, where Tc is the corresponding color
temperature. We want to stress that the color temperature
is here simply a "working parameter" and it is not used to
constrain the effective stellar temperature.
The final dense grid was obtained interpolating between
two adjacent bands and extrapolating in the UV and IR
ranges not covered by the observations. The apparent bolo-
metric magnitude was then computed by means of integra-
tion over the energy distribution as follows:
mbol = − 2.5 log10 Z fλ dλ + ZP
The zero point (ZP) of the bolometric scale is given by
ZP = M ⊙
bol + 2.5 log10 (F ⊙
bol)
so its actual value depends upon the assumed M ⊙
bol and
V =4.82 (so BC⊙
F ⊙
bol. We used M⊙
V = -- 0.07),
L⊙
bol = 3.86×1033 erg s−1, as reported by Armandroff (1989),
and the relation: F ⊙
bol/(4πd2), where d=10 pc. We
obtained ZP= -- 11.478.
bol = 4.75, M⊙
bol = L⊙
The accuracy of the inferred bolometric magnitudes pri-
marily depends on the number of observed bands since ex-
trapolations over a wide wavelength range may represent a
severe approximation of the true energy distribution.
4
P. Montegriffo et al.
(a) mbol (tot) -- mbol (UBVRJK), (b) mbol (tot) --
Figure 1.
mbol (UBVIJK), (c) mbol (tot) -- mbol (BVIJK), and (d) mbol
(tot) -- mbol (VIJK) vs the (V -- K) color (see Sect.3.1.1). mbol (tot)
was computed using the complete UBVRIJHKLMN set of filters.
Open circles mark the stars by Morel & Magnenat (1978), while
the continuous lines are our best fits.
Figure 2. Differences between the inferred BCK using our proce-
dure and that of the models, as a function of the (V -- K) color. This
comparison is performed using as reference database the set of
synthetic grids of stellar atmospheres by BCP97. The dashed hor-
izontal line is a systematic zero point shift of ∼0.01 between the
two calibrations (our -- models) (see Sect.3.1.1). At (V -- K)≤ 1 the
Balmer discontinuity strongly affects the accuracy of our BCK .
In order to check the possible corrections to apply to the
inferred values when a small number of bands was observed,
we performed the following experiment: we used the catalog
by Morel & Magnenat (1978) which contains 212 stars with
a complete set of UBVRIJHKL photometry in the Johnson
system and we applied our method to derive the apparent
bolometric magnitudes with varying the number of available
photometric bands, using the Johnson absolute flux calibra-
tion (columns #2 and #3 in Table 2).
In Fig.1 we plot as an example a few simulations of
the differences between mbol computed using all the Morel
& Magnenat (1978) bands and only those available in our
GGC database vs the (V -- K) color. Similar simulations were
performed using (J -- K). The continuous line in the plots is
our numerical best fit and represents the final corrections
we applied to the bolometric magnitudes of the stars in our
database as a function of their (V -- K) and (J -- K) colors and
the available photometric bands.
On average, the discrepancies between the bolometric
corrections computed using the entire set of photometric
data and those computed using only a few bands are always
within a few hundredths of a magnitude as one can see in
Fig.1. Larger discrepancies, up to 0.1 mag, were found in
the case of hot stars which lack U band observations: in this
case, the UV continuum computed with a Planck function
using the (B -- V) color temperature is overestimated.
As a further check on the accuracy of the proposed
method to derive the bolometric corrections we also recon-
structed the energy distribution using direct trapezoidal in-
tegrations as suggested by many authors. The results of such
a computation indicated that the corrections to apply if only
a few colors are available should be even larger than in the
case of integrations by means of Planck functions.
We did not apply any further correction due to the
Balmer discontinuity in the UV and possible molecular
sources of opacity in the IR. Nevertheless, we performed
some tests to evaluate the entity of such an effect using
UBVRIJHKL synthetic color indices computed by Bessell,
Castelli & Plez (1997, hereafter BCP97) and derived from
different grids of model atmospheres, which include both
the ATLAS9 models computed without any overshooting
for the convection (Castelli et al. 1997) and NMARCS mod-
els computed by Plez, Brett, & Nordlund (1992) and Plez
(1995). The synthetic indices derived from Plez (1995) mod-
els are computed for Tef f ranging from 3600 K to 4750
K, gravity log g ranging from from -- 0.5 to 3.5, and metal-
licities [Fe/H]= -- 0.6, -- 0.3,0.0,0.3, 0.6. Synthetic indices from
ATLAS9 models and from Plez et al. (1992) models are tab-
ulated only for the solar metallicity.
We computed the BCK of these synthetic stars follow-
ing the same procedure as for true stars in our database
using the Bessell absolute flux calibration (columns #4 and
#5 in Table 2). We compared the inferred values with those
of the models themselves. The results of this comparison are
plotted in Fig.2 as a function of the (V -- K) color. For (V --
K)>1, the agreement between the two scales is satisfactory
apart from an average shift in the zero point ∆BCK =0.01,
possibly ascribed to the different assumption on M⊙
bol (they
adopt M⊙
bol = 4.74).
For bluer colors the agreement between the two scales
gets exponentially worse and worse since the continuum
c(cid:13) 1997 RAS, MNRAS 000, 000 -- 000
Bolometric Corrections and Temperature scales for Population II giants
5
Figure 3. BCK as a function of (V -- K) and (J -- K) colors for
metal poor (left panels) and metal rich (right panels) stars. Con-
tinuous lines are our best fit.
Figure 4. BCV vs (B -- V), (V -- I) and (V -- J) colors for metal poor
(left panels) and metal rich (right panels) stars. Continuous lines
are our best fits.
shape is strongly affected by the Balmer discontinuity and
a Planck function is unable to properly reproduce it.
In order to investigate a possible dependence of the
bolometric correction vs color relations on the metallicity,
we divided the observed clusters into two groups: one at low
metallicity ([Fe/H]< -- 1.0), using about 2000 stars in M15,
M30, M68, M55 and M4, and the other at higher metallic-
ity ([Fe/H]> -- 1.0), using about 4500 stars in M107, M69, 47
Tuc, NGC6553 and NGC6528). We found a small (a few hun-
dredths of a magnitude) difference when fitting our metal
poor and metal rich stars in the empirical BCK vs (V -- K)
and (J -- K) planes, more evident at (V -- K)<2 and (J -- K)<0.7.
This difference, that we want to stress is purely observa-
tional, could be explained in terms of a small metallicity
dependence of the (V -- K) and (J -- K) colors as suggested by
the model atmospheres whose continuum opacities, particu-
larly at optical wavelengths, are slightly affected by changing
the metal content. A similar behaviour was found by Alonso
et al. (1995) using an independent, semi -- empirical approach
to estimate the bolometric flux (see Sect.1).
Nevertheless, since the difference we found is similar in
size to the intrinsic photometric error, we cannot completely
exclude the possibility that the observed effect is due to a
systematic bias in the data .
In Fig.3, we plot the derived BCK as a function of the
(V -- K) and (J -- K) colors for metal poor and metal rich stars
separately. Our best fits are shown as continuous lines. The
relation with the lowest spread is, as expected, that based on
the (V -- K) color which has the largest wavelength baseline.
In Table 3 we report the corresponding BCK values as
a function of the (V -- K) and (J -- K) colors from our best fits,
both for metal poor and metal rich stars. The BCK is fixed
(in other words we fixed the stellar temperature, see next
section) and the colors slightly change (≤ 0.1 mag) in the
two metallicity regimes.
c(cid:13) 1997 RAS, MNRAS 000, 000 -- 000
3.1.2 Other bolometric corrections
Using our database we also inferred an estimate of BCV as a
function of the (B -- V), (V -- I) and (V -- J) colors, respectively,
for metal poor and metal rich stars, as in the case of the
BCK . In Fig.4, we plot the results and in Table 3 we listed
the values for the corresponding BCK .
The possible small dependence of the (V -- K) color on
metallicity should imply a similar dependence of the BCV ;
hence, BCK could be a purer thermometer than BCV .
Moreover, in the BCV vs (B -- V) plane the difference
between metal poor and metal rich stars is more pronounced
(up to about 0.2 mag), since metallicity effects are more
important at shorter wavelengths.
3.2 The determination of Tef f
Our photometric database does not allow us to calibrate di-
rectly an empirical temperature scale since we do not know
the angular diameter of the observed stars. However DB93
provided a sample of field giants with measured angular di-
ameters by means of Michelson interferometry and lunar
occultations. We used his database, our BC scale and zero
point to calibrate a new empirical relation which links the
effective temperature to the BCK .
We also provided fully theoretical temperature scales as
a function of suitable colors and BC, using the grids of stellar
atmospheres computed by BCP97 at solar metallicity.
3.2.1 Empirical Tef f
We used the DB93's giants listed in his Table 1 and 3 of
DB93, for which accurate estimates of the angular diameter
exist, and by means of a second order polynomial fit we
obtained V fluxes as a function of the (V -- K) color, using our
6
P. Montegriffo et al.
V -- K
Poor
-0.413
-0.338
-0.259
-0.177
-0.091
-0.001
0.092
0.190
0.292
0.399
0.510
0.626
0.745
0.869
0.996
1.123
1.249
1.380
1.506
1.641
1.778
1.925
2.083
2.239
2.415
2.595
2.791
3.006
3.265
3.551
3.860
4.216
4.626
--
--
--
--
--
-0.40
-0.30
-0.20
-0.10
0.00
0.10
0.20
0.30
0.40
0.50
0.60
0.70
0.80
0.90
1.00
1.10
1.20
1.30
1.40
1.50
1.60
1.70
1.80
1.90
2.00
2.10
2.20
2.30
2.40
2.50
2.60
2.70
2.80
2.90
3.00
3.10
3.20
3.30
-0.137
-0.119
-0.099
-0.075
-0.052
-0.030
-0.003
0.026
0.055
0.082
0.113
0.145
0.178
0.210
0.244
0.279
0.316
0.350
0.385
0.423
0.461
0.498
0.532
0.572
0.612
0.653
0.690
0.738
0.790
0.846
--
--
--
--
--
--
--
--
0.013
0.038
0.059
0.077
0.091
0.101
0.108
0.110
0.108
0.101
0.090
0.074
0.055
0.031
0.004
-0.023
-0.049
-0.080
-0.106
-0.141
-0.178
-0.225
-0.283
-0.339
-0.415
-0.495
-0.591
-0.706
-0.865
-1.051
-1.260
-1.516
-1.826
--
--
--
--
--
-0.118
-0.089
-0.062
-0.040
-0.009
0.033
0.068
0.078
0.078
0.100
0.133
0.168
0.191
0.218
0.259
0.307
0.353
0.406
0.451
0.510
0.572
0.651
0.747
0.838
0.961
1.088
1.239
1.416
--
--
--
--
--
--
--
--
--
--
-0.241
-0.204
-0.169
-0.132
-0.098
-0.068
-0.032
-0.004
0.203
0.253
0.293
0.341
0.390
0.444
0.498
0.548
0.592
0.642
0.681
0.731
0.782
0.843
0.915
0.981
1.066
1.152
1.250
1.362
1.512
--
--
--
--
--
--
--
--
--
-0.420
-0.360
-0.300
-0.237
-0.176
-0.120
-0.064
0.157
0.234
0.319
0.365
0.436
0.522
0.615
0.721
0.828
0.918
1.018
1.095
1.193
1.289
1.402
1.531
1.647
1.792
1.934
2.090
2.263
2.481
2.713
2.950
--
--
--
--
--
--
--
Table 3. BCK as a function of (V -- K), (J -- K), BCV , (B -- V), (V -- I), (V -- J) colors for metal poor and metal rich stars, respectively, and
both empirical and theoretical Tef f .
BCK
BCV
B -- V
J -- K
BCV
V -- I
V -- J
J -- K
V -- K
Rich
--
--
--
--
--
--
--
--
--
0.425
0.522
0.621
0.723
0.829
0.938
1.050
1.168
1.289
1.417
1.549
1.688
1.834
1.987
2.147
2.317
2.498
2.691
2.900
3.127
3.390
3.703
4.117
4.696
5.378
6.263
7.485
9.444
19.746
B -- V
V -- I
V -- J
--
--
--
--
--
--
--
--
--
0.082
0.107
0.132
0.159
0.187
0.216
0.245
0.274
0.305
0.336
0.369
0.405
0.444
0.486
0.533
0.582
0.634
0.689
0.747
0.807
0.870
0.937
1.006
1.077
1.152
1.229
1.310
1.393
1.478
--
--
--
--
--
--
--
--
--
0.075
0.078
0.079
0.077
0.071
0.062
0.050
0.032
0.011
-0.017
-0.049
-0.088
-0.134
-0.187
-0.247
-0.317
-0.398
-0.491
-0.600
-0.727
-0.890
-1.103
-1.417
-1.896
-2.478
-3.263
-4.385
-6.244
-16.446
--
--
--
--
--
--
--
--
--
--
--
0.373
0.378
0.393
0.413
0.440
0.476
0.516
0.564
0.614
0.670
0.731
0.795
0.863
0.932
1.005
1.087
1.173
1.263
1.362
1.468
1.612
1.659
1.690
1.732
1.793
--
--
--
--
--
--
--
--
--
--
--
--
--
--
--
--
--
--
--
--
--
-0.067
0.111
0.321
0.561
0.761
0.900
1.027
1.133
1.226
1.319
1.426
1.559
1.762
2.111
2.564
3.033
3.620
4.443
8.119
--
--
--
--
--
--
--
--
--
--
--
0.563
0.575
0.611
0.662
0.724
0.808
0.897
1.003
1.111
1.228
1.352
1.480
1.610
1.746
1.889
2.038
2.195
2.359
2.544
2.767
3.102
3.637
4.268
5.099
6.258
8.142
18.471
Tef f
Tef f
Emp Theor
10462
10139
9824
9518
9219
8928
8645
8370
8102
7842
7589
7343
7105
6871
6642
6421
6208
6003
5806
5618
5438
5264
5097
4933
4774
4618
4468
4321
4179
4040
3904
3771
3639
3516
3385
3243
2929
2499
--
--
--
--
--
--
--
--
--
7602
7386
7173
6963
6756
6553
6353
6156
5963
5774
5589
5407
5229
5056
4887
4723
4563
4406
4254
4106
3962
3819
3667
3523
3409
3317
3247
3199
3174
zero points. We did not use DB93's linear relations since they
have a discontinuity around (V -- K)=3.7 which propagates
into the relation which gives Tef f as a function of (V -- K)
and BCK .
Following the procedure described in DB93 we derived
the relation:
log Tef f = 3.9619 − 0.0466(V − K) + 0.0038(V − K)2−
0.1BCK
Since DB93's database is at solar metallicity the pre-
vious relation can be regarded as reliable only for metal
rich stars. Hence we computed the effective temperature us-
ing the latter relation only for the metal rich stars in our
database and the results are reported in Fig.5 as a function
of BCK , while the mean relation is listed in Table 3. The
average dispersion around the fiducial line is ∼30 K. Outside
the 1< BCK <3.1 range the relation was extrapolated.
3.2.2 Theoretical Tef f
The calibration of the effective temperature was performed
using the grid of stellar atmospheres by BCP97 at solar
metallicity. In order to evaluate the possible shift with vary-
ing metallicity we also used a grid of color indices for so-
lar and 1/100 solar metallicity computed by Castelli (1997).
These grids are based on the ATLAS9 model atmospheres in
which the modification of the "approximate overshooting" to
the mixing-length treatment of the convection was dropped.
The indices slightly differ from the BCP97 indices owing to
different passbands and zero points. For a given tempera-
ture and metallicity we interpolated the grids in gravity by
means of the VandenBerg (1996) isochrones. For each star
in our database we derived a mean temperature using both
the (V -- K), (V -- J), (J -- K) colors and the BCK .
The relations were derived in the Bessell (1990) pho-
tometric system, as adopted by BCP97 to compute their
stellar atmosphere grids. On the other hand, our observed
database refers to the ESO photometric system. We used
the relations quoted by Bessell & Brett (1988) to transform
the (V -- K), (V -- J) and (J -- K) colors of the models from the
Bessell system to the ESO one. Moreover, since between the
BCK computed with our procedure and those of the models
a systematic shift in the zero point of 0.01 mag was found
(see Sect.3.1.1 and Fig.2), we added this shift to the BCK of
c(cid:13) 1997 RAS, MNRAS 000, 000 -- 000
Bolometric Corrections and Temperature scales for Population II giants
7
Figure 5. Empirical Tef f as a function of BCK (upper panel) for
metal rich stars. The continuous line is our best fit. The dispersion
(on average 30 K) around the fiducial line is also plotted (lower
panel).
Figure 6. Theoretical Tef f as a function of (V -- K), (V -- J), (J -- K)
and BCK .
the models. This avoids to have a systematic drift between
Tef f derived from the colors and the BCK .
In Fig.6 we plot the behaviour of the effective tempera-
ture as a function of both the (V -- K), (V -- J) and (J -- K) colors
in the ESO photometric system and the BCK . Our best fits
(continuous lines) are plotted as well. The spread visible in
the Tef f vs color distributions at the lowest temperature is
due to a gravity effect. It demonstrates how small variations
of such a parameter may strongly affect the accuracy of the
derived temperatures below 3500 K. This effect is less pro-
nounced in the Tef f vs BCK distribution. The latter is also
the most linear and sensitive relation, particularly at low
temperatures, confirming that BCK is probably the purest
stellar thermometer for cool stars.
In Fig.7 the theoretical Tef f derived from averaging the
temperatures obtained by the color and BCK as a function
of the latter are plotted together with the dispersion (on
average 180 K) around the fiducial line and in Table 3 we
listed the values.
3.2.3 Comparison between the two scales
Comparing the empirical and theoretical temperature scales
discussed in the previous sections, we infer the following
behaviours:
Figure 7. Theoretical Tef f derived from averaging the temper-
atures obtained by the color and BCK as a function of the latter
(upper panel). The continuous line is our best fit. The dispersion
(on average 180 K) around the fiducial line is also plotted (lower
panel).
• in the range 4000 -- 7500 K, a general good agreement
has been found, the theoretical scale being systematically
warmer by about 50 K;
• between 3200 and 4000 K, this difference increases up
The behaviour in the coolest temperature range can
be explained as a gravity effect, as already mentioned in
Sect.3.2.2, in the sense that the interpolation in gravity may
be quite uncertain.
to about 100 K;
• below 3200 K, the two scales diverge, the theoretical
one becoming progressively cooler, up to 800 K.
c(cid:13) 1997 RAS, MNRAS 000, 000 -- 000
8
P. Montegriffo et al.
Figure 8. (J -- K) vs (V -- K) color -- color diagram for the mean
loci of the metal rich (continuous line) and metal poor (dashed
line) observed stars and different model atmospheres. ATLAS9
models by Castelli et al. (1997) are without overshooting. The
mean sequences of a sample of field giants and dwarfs by Bessell
& Brett (1988) and the reddening vector are plotted as well.
Figure 9. K vs (J -- K) diagram of the isochrones by Bergbusch &
VandenBerg (1992) transformed into the observational plane at
two different metallicities: dotted lines are the relations inferred
using BG89 transformations, continuous lines represent those cor-
rected as suggested by Bell (1992) to reproduce the observed RGB
distribution, and stars are those obtained using our BCK and
Tef f transformations.
4 DISCUSSION
4.1 A few more tests
In order to check the consistency between our observed
database and the adopted model atmospheres, both used
in the calibrations of the temperature scale (see Sect.3.2.2),
we constructed a (J -- K) vs (V -- K) color -- color diagram in
the ESO photometric standard system for all the stars in our
database (see Fig.8). The mean loci of the observed metal
poor and metal rich stars, different models, and the mean
sequences of a sample of field giants and dwarfs as reported
in Table 2 and 3 of Bessell & Brett (1988) are plotted as
well. The observed and synthetic distributions well overlap
one to each other and also overlap the field giant sequence,
indicating that the two photometric samples are fully con-
sistent.
Another interesting test we performed is to apply our
BCK and Tef f relations to the widely used isochrones com-
puted by Bergbusch & VandenBerg (1992), the only ones
presently published in the IR -- planes. Bell (1992, and refer-
ence therein) has shown how these theoretical isochrones
transformed into the observational plane using the Bell
& Gustafsson (1989, hereafter BG89) model atmospheres
hardly reproduce the sequence of cool stars observed by
Glass (1974a,b), being the former too red by ≈0.1 mag along
the RGB.
The results of our test are plotted in Fig.9: the se-
quences obtained using our new calibrations well overlap
those computed by using BG89 and corrected as suggested
by Bell (1992) along the RGB (continuous lines), while
they overlap the uncorrected ones in the turn -- off region
(dashed lines). Hence, it is clear that the discrepancy found
by Bell (1992) along the RGB was not due to the theoret-
ical isochrones but should rather be a consequence of the
adopted transformations into the IR observational plane.
4.2 Comparison with published BC scales
Our BCs, as reported in Table 3, in both the metal poor
and metal rich regimes, were then compared with published
scales.
In the BCK vs (V -- K) plane, we considered the relations
presented by Johnson (1966), Lee (1970), Ridgway et al.
(1980), DBR87 and BG89, as tabulated by DB93 and scaled
by 0.01 mag to take into account the different BC⊙ adopted.
Moreover, we considered those obtained by FPC81, Bessell
& Wood (1984), Blackwell & Petford (1991) and Alonso et
al. (1995). And, finally, the BCK of the model (BCP97), as
computed to derive Tef f (see Sect.3.2.2), were also included.
The residuals (our -- others) vs the (V -- K) color are plot-
ted in Fig.10 for metal rich (lower panel) and metal poor
(upper panel) stars. As can be seen, all the scales agree one
to each other within 0.1 mag. Interesting enough is the op-
posite behaviour of ∆BCK for metal rich and metal poor
stars. While our BCK are systematically larger than the
other scales in metal rich stars, they are systematically lower
in metal poor ones for a given (V -- K). This may reflect the
fact that all the other scales average the information of the
two metallicity regimes.
Concerning BCK as a function of (J -- K), we compared
our scale with those presented by Frogel, Persson & Cohen
(1980), Bessell & Wood (1984), Bessell (1991) and with the
BCK of the model (BCP97).
c(cid:13) 1997 RAS, MNRAS 000, 000 -- 000
Bolometric Corrections and Temperature scales for Population II giants
9
Figure 12. ∆Tef f vs log R and empirical Tef f between our
empirical scales and the theoretical ones by Blackwell & Lynas --
Gray (1994).
Table 4. R -- parameter of the IRFM as a function of Tef f
and BCK .
Tef f
3250
3500
3750
4000
4250
4500
4750
5000
5250
5500
BCK
3.10
2.82
2.65
2.47
2.30
2.14
1.98
1.83
1.69
1.55
log R†
0.5504
0.6613
0.7303
0.7993
0.8676
0.9327
0.9953
1.0555
1.1134
1.1692
Tef f
5750
6000
6250
6500
6750
7000
7250
7500
7750
BCK i
1.41
1.28
1.15
1.03
0.90
0.78
0.66
0.55
0.43
log R†
1.2235
1.2764
1.3278
1.3781
1.4275
1.4757
1.5232
1.5698
1.6158
† logarithmic value of R = Fbol
F2.2µm
in units of µm.
the adopted colors are plotted in Fig.11 for both metal poor
(left panels) and metal rich (right panels) stars. The average
spread among different scales is larger in these planes than
that inferred in the BCK vs (V -- K) plane.
4.3 The IRFM
Using our BCK and empirical Tef f we can also provide an
independent, empirical calibration of the R parameter. This
parameter was computed using the same ZP of our bolomet-
ric scale and the K band flux calibration of Table 2 (columns
#6 and #7). The results are listed in Table 4.
This empirical scale has been compared with the re-
lation obtained interpolating the theoretical Blackwell &
Lynas -- Gray (1994) grid, which tabulated log R as a func-
tion of Tef f and log g. The interpolation in gravity was
performed using the VandenBerg (1996) isochrones
In Fig.12 we plot the residual Tef f between the two
relations as a function of our empirical log R and Tef f . For
log R>1 there is a good agreement (∆Tef f < ± -- 50 K)
among the two relations. For log R<1 the empirical Tef f
seem to be systematically hotter than those of the Blackwell
& Lynas -- Gray (1994) models.
Figure 10. Comparison between our BCK vs (V -- K) scale as
reported in Table 3 and those obtained by other authors for metal
rich (lower panel) and metal poor (upper panel) stars (see keys
in the plot and Sect.4.2).
Figure 11. Comparison between our BCK vs (J -- K), BCI vs
(V -- I) and BCV vs (B -- V) scales and those by other authors for
metal poor (left panels) and metal rich (right panels) stars. Keys
are as in Fig.9, otherwise indicated.
The relation in the BCI vs (V -- I) plane was compared
with those listed by Bessell & Wood (1984), Da Costa &
Armandroff (1990) and Bessell (1991).
Finally, the BCV vs (B -- V) relation was compared with
that obtained by Blackwell & Petford (1991).
5 CONCLUSIONS
The main results obtained from our analysis can be summa-
rized as follow:
The residuals drawn from each individual comparison vs
• By exploiting the use of a large photometric database
c(cid:13) 1997 RAS, MNRAS 000, 000 -- 000
10
P. Montegriffo et al.
of Pop II stars in GGCs we derived new relations for: BCK
vs (V -- K) and (J -- K), BCV vs (B -- V), (V -- I) and (V -- J), to
infer empirical bolometric corrections from observed colors.
• By making use of both an empirical relation and model
atmospheres we calibrated two different scales to infer reli-
able stellar effective temperatures in the range 3000 -- 7500
K.
• We also calibrated an empirical relation between the R
parameter of the IRFM and the stellar temperature.
All these relations are summarized in Table 3, where for a
given BCK the corresponding colors and BCV in two differ-
ent metallicity regimes and Tef f can be read. These rela-
tions should be the most suitable to calibrate the red stellar
sequences in the color -- magnitude diagrams, like the RGB
and AGB in old GGCs.
ACKNOWLEDGMENTS
We wish to thank all the people who kindly made available
to us photometric data, model atmospheres, and isochrones.
We warmly thank Mike Bessell for his very careful read-
ing of the paper, Fiorella Castelli, Giampaolo Di Benedetto,
Maria Lucia Malagnini and Bob Kurucz for helpful discus-
sions. A special thank to the ADC Service to provide the
Morel & Magnenat (1978) Catalogue. Paolo Montegriffo was
supported by a 1996-grant of the Fondazione del Monte,
Rolo Banca 1473. For their financial support, we also thank
the Ministero della Universit`a e della Ricerca Tecnologica
(MURST) and the Agenzia Spaziale Italiana (ASI).
REFERENCES
Armandroff T.E., 1989, AJ 97, 375
Auriere M., Lauzeral C., 1996, Private Communication
Alonso A., Arribas S., Martinez -- Roger C., 1995, A&A 297, 197
Bell R.A., Gustafsson B., 1989, MNRAS 236, 653 (BG89)
Bell R.A., 1992, MNRAS 257, 423
Bergbusch P.A., 1996, AJ 112, 1061
Bergbusch P.A., VandenBerg D.A., 1992, ApJS 81, 163
Bessell M.S., 1990, PASP 102, 1181
Bessell M.S., 1991, AJ 101, 662
Bessell M.S., Brett J.M., 1988, PASP 100, 1134
Bessell M.S., Castelli F., Plez B., 1997, in preparation (BCP97)
Bessell M.S., Wood P.R., 1984, PASP 96, 247
Blackwell D.E., Shallis M.J., Selby M.J., 1979, MNRAS 188, 847
Blackwell D.E., Petford A.D., 1991, A&A 250, 459
Blackwell D.E., Lynas -- Gray A.E., 1994, A&A 282, 899
Buzzoni A., 1989, AJS 71, 817
Buzzoni A., 1996, Private Communication
Carney B.W., Aaronson M., 1979, AJ 84, 867
Castelli F., Gratton R.G., Kurucz R.L., 1997, A&A 318, 841
Castelli F., 1997, Private Communication
Da Costa G.S, Armandroff T.E., 1990, AJ 100, 162
Di Benedetto G.P., 1993, A&A 270, 315
Di Benedetto G.P., Rabbia Y., 1987, A&A 188, 114 (DBR87)
Ferraro F.R., Clementini G., Fusi Pecci F., Buonanno R., 1991,
MNRAS 252, 357
Ferraro F.R., Fusi Pecci F., Guarnieri M.D., Moneti A., Origlia
L., Testa V., 1994a, MNRAS 266, 829
Ferraro F.R., Fusi Pecci F., Guarnieri M.D., Moneti A., Origlia
L., Testa V., Ortolani S., 1994b, in McLean I.S., ed., Infrared
Astronomy with Arrays: The next Generation. Kluwer, Dor-
drecht, p.235
Frogel J.A., Persson S.E., Cohen J.G., 1980, ApJ 239, 495
Frogel J.A., Persson S.E., Cohen J.G., 1981, ApJ 246, 842
(FPC81)
Glass I.S., 1974a, MNRAS South Afr. 33, 54
Glass I.S., 1974b, MNRAS South Afr. 33, 71
Guarnieri M.D., Ortolani S., Renzini A., Montegriffo P., Moneti
A., Barbuy B., Bica E., 1997, A&A (in press)
Johnson H.L., 1966, ARAA 4, 193
Kurucz R.L., 1979, ApJS 40, 1
Kurucz R.L., 1993, ATLAS9 Stellar Atmosphere Programs and 2
km s−1 grid, Kurucz CD -- ROM No 18
Lee S.W., 1977, A&AS 27, 367
Megessier C., 1995, A&A 296, 771
Montegriffo P., Ferraro F.R., Fusi Pecci F., Origlia L., 1995, MN-
RAS 276, 739
Moorwood A.F.M et al., 1992, The Messenger 69, 61
Morel M., Magnenat P. 1978, A&AS 34, 477
Ortolani S., Desidera S., 1996, Private Communication
Ortolani S. et al., 1995, Nature 377, 701
Piotto G., 1996, Private Communication
Plez B., Brett J.M., Nordlund A., 1992, A&A 256, 551
Plez B., 1995, Private Communication
Ridgway S.T., Wells D.C., Joyce R.R., Allen R.G., 1979, AJ 84,
247
Rieke G.H., Lebovsky M.J., 1985, ApJ 288, 618
Stetson P.B., 1994 PASP 106, 250
Tinney C.G., Mould J.R., Reid I.N., 1993, AJ 105, 1045
VandenBerg D.A., 1996, Private Communication
Walker A.R., 1994, AJ 108, 555
Zinn R.J., 1985, ApJ 293, 424
c(cid:13) 1997 RAS, MNRAS 000, 000 -- 000
|
astro-ph/0205313 | 2 | 0205 | 2003-02-21T00:00:07 | Breaking the "Redshift Deadlock" -- II: The redshift distribution for the submillimetre population of galaxies | [
"astro-ph"
] | In this paper we apply our Monte-Carlo photometric-redshift technique, introduced in paper I (Hughes et al. 2002), to the multi-wavelength data available for 77 galaxies selected at 850um and 1.25mm. We calculate a probability distribution for the redshift of each galaxy, which includes a detailed treatment of the observational errors and uncertainties in the evolutionary model. The cumulative redshift distribution of the submillimetre galaxy population that we present in this paper, based on 50 galaxies found in wide-area SCUBA surveys, is asymmetric, and broader than those published elsewhere, with a significant high-z tail for some of the evolutionary models considered. Approximately 40 to 90 per cent of the sub-mm population is expected to have redshifts in the interval 2 < z < 4. Whilst this result is completely consistent with earlier estimates for the sub-mm galaxy population, we also show that the colours of many (< 50 per cent) individual sub-mm sources, detected only at 850um with non-detections at other wavelengths, are consistent with those of starburst galaxies that lie at extreme redshifts, z > 4. Spectroscopic confirmation of the redshifts, through the detection of rest-frame FIR--mm wavelength molecular transition-lines, will ultimately calibrate the accuracy of this technique. We use the redshift probability distribution of HDF850.1 to illustrate the ability of the method to guide the choice of possible frequency tunings on the broad-band spectroscopic receivers that equip the large aperture single-dish mm and cm-wavelength telescopes. | astro-ph | astro-ph | Mon. Not. R. Astron. Soc. 000, 1 -- ?? (2002)
Printed 31 October 2018
(MN LATEX style file v1.4)
Breaking the "Redshift Deadlock" -- II: The redshift
distribution for the submillimetre population of galaxies
Itziar Aretxaga1, David H. Hughes1, Edward L. Chapin1, Enrique Gaztanaga2,1,
James S. Dunlop3, Rob J. Ivison4
1Instituto Nacional de Astrof´ısica, ´Optica y Electr´onica (INAOE), Aptdo. Postal 51 y 216, 72000 Puebla, Mexico
2IEEC/CSIC, Edifici Nexus-201, c/ Gran Capit´an 2-4, 08034 Barcelona, Spain
3Institute for Astronomy, University of Edinburgh, Blackford Hill, Edinburgh, EH9 3HJ, UK
4Astronomy Technology Center, Royal Observatory, Blackford Hill, Edinburgh, EH9 3HJ, UK
31 October 2018
ABSTRACT
Ground-based sub-mm and mm-wavelength blank-field surveys have identified more
than 100 sources, the majority of which are believed to be dusty optically-obscured
starburst galaxies. Colours derived from various combinations of FIR, submillimetre,
millimetre, and radio fluxes provide the only currently available means to determine
the redshift distribution of this new galaxy population.
In this paper we apply our Monte-Carlo photometric-redshift technique, intro-
duced in paper I (Hughes et al. 2002), to the multi-wavelength data available for 77
galaxies selected at 850µm and 1.25 mm. We calculate a probability distribution for
the redshift of each galaxy, which includes a detailed treatment of the observational
errors and uncertainties in the evolutionary model. The cumulative redshift distribu-
tion of the submillimetre galaxy population that we present in this paper, based on 50
galaxies found in wide-area SCUBA surveys, is asymmetric, and broader than those
published elsewhere, with a significant high-z tail for some of the evolutionary models
considered. Approximately 40 to 90 per cent of the sub-mm population is expected
to have redshifts in the interval 2 ≤ z ≤ 4. Whilst this result is completely consis-
tent with earlier estimates for the sub-mm galaxy population, we also show that the
colours of many (<∼ 50 per cent ) individual sub-mm sources, detected only at 850µm
with non-detections at other wavelengths, are consistent with those of starburst galax-
ies that lie at extreme redshifts, z > 4. Spectroscopic confirmation of the redshifts,
through the detection of rest-frame FIR -- mm wavelength molecular transition-lines,
will ultimately calibrate the accuracy of this technique. We use the redshift probabil-
ity distribution of HDF850.1 to illustrate the ability of the method to guide the choice
of possible frequency tunings on the broad-band spectroscopic receivers that equip the
large aperture single-dish mm and cm-wavelength telescopes.
Key words: submillimetre, millimetre, cosmology, galaxy evolution, star-formation
3
0
0
2
b
e
F
1
2
2
v
3
1
3
5
0
2
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
1
INTRODUCTION
The lack of robust redshift information is one of the out-
standing problems in understanding the nature of the pre-
sumed high-redshift galaxies identified in blank-field sub-
mm and mm wavelength surveys.
In paper I of this series (Hughes et al. 2002) we de-
scribed Monte Carlo simulations of 250 -- 500µm surveys from
the balloon-borne telescope BLAST (Devlin et al. 2001)
and the SPIRE instrument on the Herschel satellite, as
well as longer-wavelength ground-based 850µm surveys with
SCUBA. These simulations included a detailed treatment of
the observational and evolutionary model-dependent errors.
The sub-mm colours of galaxies derived from these mock
surveys demonstrated that it will be possible to derive pho-
tometric redshifts, with a conservative r.m.s. accuracy of
∆z ∼ ±0.5, for thousands of optically-obscured galaxies.
Thus we can look forward to breaking the redshift deadlock
which, at the moment, prevents a reliable estimate of the
evolutionary history of a population that contributes a sig-
nificant fraction (>∼ 50 per cent for S850µm > 2 mJy) to the
sub-mm background (Hughes et al. 1998, Smail et al. 2002).
In this second paper we apply the same Monte Carlo
photometric-redshift technique to the existing FIR -- radio
c(cid:13) 2002 RAS
2
Itziar Aretxaga et al.
multi-wavelength data for 77 sources first identified in blank-
field 850µm (SCUBA) and 1.2 mm (MAMBO) surveys. We
calculate the redshift probability distributions of the indi-
vidual sub-mm and mm galaxies, taking into account obser-
vational and model-dependent uncertainties, and thus pro-
vide a measurement of the cumulative redshift distribution
for the blank-field sub-mm galaxy population. This redshift
distribution is completely insensitive to ambiguities in the
identification of the optical counterparts.
The structure of the paper is as follows: section 2 gives a
brief description of the Monte Carlo simulations, the method
to derive photometric redshifts from the radio -- sub-mm -- FIR
colours of galaxies drawn from the mock catalogues, and the
caveats and advantages of this technique over other redshift
indicators; section 3 presents individual redshift probabil-
ity distributions for blank-field sub-mm galaxies. In addi-
tion, we include a brief discussion of the photometric red-
shifts for radio and FIR-selected galaxies subsequently de-
tected at 850µm. Furthermore, we determine the accuracy
of our method through the comparison of the photometric
redshifts for those 8 sub-mm sources that have published
rest-frame optical spectroscopic redshifts. Finally, in sec-
tion 3 we present the cumulative redshift distribution of
the blank-field sub-mm population. Section 4 discusses the
likelihood of a radio detection of previously detected blank-
field sub-mm sources, based only on the known dispersion
of the 850µm/1.4 GHz colour in starburst galaxies, radio-
quiet AGN and ULIRGs, which are believed to be represen-
tative of the high-z sub-mm galaxy population. Section 5
describes how the determination of redshift probability dis-
tributions for individual sub-mm galaxies is motivating the
efforts to provide spectroscopic confirmation of the redshifts
at mm and cm wavelengths. Ultimately, these spectroscopic
redshifts will accurately calibrate the photometric redshifts
derived from rest-frame FIR -- radio-wavelength data. Ap-
pendix A contains the complete catalogue of spectral en-
ergy distributions (SEDs) and redshift distributions derived
in this paper.
2 PHOTOMETRIC REDSHIFT ESTIMATION
TECHNIQUE
The Monte Carlo simulations described in this paper pro-
duce the redshift probability distribution for an individual
galaxy. The advantage of this technique, compared to the
popular maximum-likelihood methods, is that it provides
more information than a simple estimate of the first and
second moments of the redshift distributions. A detailed de-
scription of the technique can be found in paper I; we only
offer a summary here. Hereafter we will refer to a galaxy de-
tected in a SCUBA or MAMBO survey as a sub-mm galaxy,
and a sub-mm galaxy generated in the Monte Carlo simula-
tions as a mock galaxy.
We choose an evolutionary model for the 60µm lumi-
nosity function that fits the observed 850µm number-counts.
Under this model, and assuming a sub-mm survey of ∼10
deg2, we generate a catalogue of 60µm luminosities and red-
shifts. Randomly selected template SEDs are drawn from a
library of local starbursts, ULIRGs and AGN, to provide
FIR -- radio fluxes, and hence colours, for this mock cata-
logue. The fluxes of the mock galaxies include both photo-
metric and calibration errors, consistent with the quality of
the observational data for the sub-mm galaxy detected in a
particular survey. We reject from the catalogue those mock
galaxies that do not respect the detection thresholds and
upper-limits of the particular sub-mm galaxy under analy-
sis. In this paper the redshift probability distribution of a
sub-mm galaxy is calculated as the normalized distribution
of the redshifts of the mock galaxies in the reduced cata-
logue, weighted by the likelihood of identifying the colours
and fluxes of each mock galaxy with those of the sub-mm
galaxy in question. Thus, we also take into account as much
prior information on the sub-mm galaxy population (and
their uncertainties) as possible. For example, we consider
uncertainties in the sub-mm galaxy number counts, their
favoured evolutionary models, the luminosity function of
dusty galaxies, lensing by a foreground magnifying cluster,
etc. In the case that the sub-mm galaxy is identified as a
lensed source, then the mock catalogue is also amplified by
the appropriate factor, allowing intrinsically fainter objects
to be introduced into the reduced catalogue, using the same
observational measurement errors.
We adopt a flat, Λ-dominated cosmological model
(H0 = 67 km s−1 Mpc−1, ΩM = 0.3, ΩΛ = 0.7). To quan-
tify the sensitivity of the individual redshift distributions
on the assumed evolutionary history of the sub-mm galaxy
population, we consider six different models that are able
to reproduce the observed 850µm number-counts within the
uncertainties. We will refer to them by the following codes:
le1: luminosity evolution of the 60µm luminosity function
Φ[L, z = 0] (Saunders et al. 1990) as (1 + z)3.2 for 0 < z <
2.2. Constant evolution is imposed between z = 2.2 − 10. No
galaxies beyond z = 10, Φ[L, z > 10] = 0.
le1L13: same evolutionary description as in le1, but im-
posing a luminosity cut-off above 1013L⊙.
le2: luminosity evolution of Φ[L, 0] as (1 + z)3 for 0 < z <
2.3. Constant evolution is imposed between z = 2.3 − 6.
Φ[L, z > 6] = 0.
le2L13: same evolutionary description as in le2, but im-
posing a luminosity cut-off above 1013L⊙.
lde1 luminosity evolution of Φ[L, 0] as (1 + z)1.9 and den-
sity evolution as (1 + z)2.5 for 0 < z < 3.5; luminosity and
density evolution as (1 + z)−3.0 for 3.5 ≤ z ≤ 7
lde2 luminosity evolution of Φ[L, 0] as (1 + z)3.5 and den-
sity evolution as (1 + z)3 for 0 < z < 2.0; no further lu-
minosity evolution, and density evolution as (1 + z)−2 for
2 ≤ z ≤ 6
The models used in this analysis provide a wide range of
the many possible evolutionary scenarios that reproduce the
sub-mm to mm number counts. These models will allow us
to test the robustness of the photometric redshifts that we
will derive in the next section.
An extended library of SEDs (compared to paper I)
is used in the Monte Carlo simulations. The library con-
tains 20 starbursts, ULIRGs and AGN that have been either
extensively mapped at FIR -- radio wavelengths, or galax-
ies that are sufficiently distant that single-beam photomet-
ric observations measure the total flux: M 82, NGC 3227,
NGC 2992, NGC 4151, NGC 7469, NGC 7771, NGC 1614,
I Zw 1, Mkn 231, Arp 220, Mkn 273, NGC 6240, UGC 5101,
IRAS 10214+4724,
IRAS 08572+3915,
IRAS 14348−1447
IRAS 15250+3609,
IRAS 05189−2524,
IRAS 12112+0305,
c(cid:13) 2002 RAS, MNRAS 000, 1 -- ??
and Cloverleaf. This selection has the important conse-
quence of reducing any dispersion in the SEDs due to aper-
ture effects. A continuum SED that includes synchrotron
and free-free emission in the radio to mm regime, and ther-
mal emission from dust grains in the mm-to FIR regime, is
fitted to the observational data (Chapin et al. 2003). The
luminosities and temperatures of the dust that dominates
the rest-frame FIR emission are evenly distributed in the
ranges 9.0 <∼ log LFIR/L⊙ <∼ 12.3 and 25 < T /K < 65.
Half of the galaxies in the SED library used in the
derivation of photometric redshifts in this paper are in com-
mon with those considered by Yun & Carilli (2002), here-
after YC02. The temperature range covered by our sample
of galaxies (25 − 65K, median 41K) is broader than that of
YC02 due to the increased fraction of cooler sources, and
is, for the same reason, colder than those of well-studied
ULIRGs (46 − 77K, median 60K, Klaas et al. 2001). Despite
the similar temperatures of our sample to late-type galaxies
selected at 60µm (25 − 55K, median 35K), drawn from the
SLUGS catalogue of Dunne et al. (2000), the shapes of the
SEDs in our library differ significantly from the local galax-
ies in the SLUGS. This illustrates an important point: it is
the shapes of the SEDs, and not the derived dust temper-
atures, that influence the estimated photometric redshifts.
Given the limited FIR -- sub-mm data, there is a degeneracy
between the dust temperatures, grain emissivity index (β),
source-size (Ω) and optical depth (τν). This makes it possible
to fit a single observed SED with a broad range of tempera-
ture, by tuning the choice of β, Ω and τν (e.g. Hughes et al
1993).
It remains unclear whether the lower-luminosity, local
galaxies in the SLUGS are closely related to the blank-field
SCUBA population which have rest-frame FIR luminosities
> 1012L⊙, assuming the sub-mm galaxies lie at z > 1. The
claim that cold SEDs match the number-counts in hydro-
dynamical cosmological simulations (e.g. Fardal et al. 2002)
should not be a basis for favouring cold SEDs, since a differ-
ent evolutionary model with warmer SEDs can fit the counts
equally well.
The strongest caveat in any photometric redshift analy-
sis is the validity of the assumption that the SED templates
of local starbursts, ULIRGs and AGN are good analogs of
the high-redshift sub-mm galaxy population. It is not until
future rest-frame FIR -- sub-mm observations from sensitive
balloon-borne or satellite experiments (paper I) accurately
measure the variety of SEDs present in the sub-mm galaxy
population, and spectroscopic redshifts for these are avail-
able, that it will possible to improve the error estimates
presented in this paper.
The technique adopted in this paper differs philosoph-
ically from those developed in other photometric-redshift
studies. Our redshift estimates do not rely on the match of
the observed data to a single template SED, and our errors
are not drawn from the departure of other local SEDs from
the adopted standard template, which has become the com-
mon practice (Carilli & Yun 1999, 2000, hereafter referred to
as CY99, CY00; Gear et al. 2000; YC02). Instead, we allow
the test galaxies to be matched against the whole variety
of template SEDs that can be detected at any given red-
shift. This point should be emphasized, since although we
assign our template SEDs at random, not all SED shapes
(scaled to the same FIR luminosity) are detected with equal
c(cid:13) 2002 RAS, MNRAS 000, 1 -- ??
Breaking the "Redshift Deadlock" -- II
3
probability at any given redshift and wavelength. The red-
shift distributions derived in this paper are often asymmet-
ric, and broader than those inferred from the single SED
template analysis. Despite the efforts of some authors to in-
clude realistic error-bars in their single-template estimates,
we consider these to have been underestimated: error bars
derived from 68% of the galaxies in a template catalogue, is
not equivalent to the 68% dispersion in the fluxes that the
whole catalogue can create, since some of the outliers con-
tribute very significantly to the departures from the mean.
Our redshift estimation method also has the advan-
tage of including a common-sense evaluation of whether a
source at a given redshift is simultaneously consistent with
the observed colours and fluxes. The probability of gener-
ating a certain luminosity at a certain redshift (i.e. flux)
is introduced by the luminosity function and the evolution-
ary model adopted. As shown below, when sub-mm galaxies
with detections at more than two wavelengths are consid-
ered, the results are quite insensitive to the differences in
the evolutionary models adopted to generate the mock cata-
logue. The consideration of reasonable evolutionary models,
however, allow plausible colour-based redshift solutions to
be excluded.
3 MONTE CARLO BASED PHOTOMETRIC
REDSHIFTS
The available data-set for the sub-mm population of galax-
ies is usually restricted to a low-S/N detection at 850µm
and an upper-limit at 450µm (from SCUBA surveys), or a
1.2 mm detection (from MAMBO surveys or an IRAM PdB
follow-up observation), with, perhaps, radio observations at
1.4, 5 or 8.6 GHz. Although a number of the SCUBA surveys
have been conducted in fields previously observed by the In-
frared Space Observatory (ISO) with the ISOPHOT camera
at 170 and 90µm, e.g. HDF (Hughes et al. 1998; Borys et
al. 2002), ELAIS N2 and the Lockman Hole (Scott et al.
2002; Fox et al. 2002), the ISO measurements are generally
too insensitive to provide any additional constraint on the
photometric redshifts.
The method described in section 2 produces mock-
catalogues that replicate the sensitivities and calibration ac-
curacies of the original surveys in which each sub-mm galaxy
was detected. The colours and fluxes of the mock galaxies
provide diagnostic diagrams to illustrate the derivation of
the most probable redshift: colour-colour-redshift (C −C −z)
diagrams, when more than 3-band detections are available;
colour-flux-redshift (C − f − z) diagrams, when just two
bands are available; and flux-redshift (f −z) diagrams; when
just one band detection is available. From these diagrams
we derive the corresponding redshift probability distribu-
tion (P − z) for each sub-mm galaxy. The distributions for
a selection of sub-mm galaxies considered in this paper are
shown in Appendix A.
3.1 Specific examples: LH850.1 and BCR11
LH850.1 is the brightest 850µm source identified in the wide-
area UK 8 mJy SCUBA survey of the Lockman Hole (Scott
et al. 2002), and has one of the most complete FIR -- radio
SEDs (Lutz et al. 2001). BCR11, named after source 11 in
4
Itziar Aretxaga et al.
Figure 1. Colour-colour-redshift (C − C − z) plot for LH850.1
(Lutz et al. 2001, Scott et al. 2002). The flux ratios of the mock
galaxies, generated under the evolutionary model le2 (section 2),
are represented as diamonds, and their redshifts are coded in
colour according to the scale shown to the right of the panel
(colours can be seen in the electronic version of the paper). The
cross represents the measured colours of LH850.1, and the dashed
box shows the 1σ uncertainty in each colour.
the sub-mm catalogue of Barger, Cowie & Richards (2000),
is a low S/N, but nonetheless representative, 850µm source
detected in a follow-up SCUBA photometry survey of the
micro-Jansky 1.4 GHz radio sources in the Hubble Deep
Field (Richards 2000).
Even though BCR11 was not first identified in a blank-
field sub-mm survey, and hence is not included in the com-
bined redshift distribution of the sub-mm selected galaxy
population, it provides a useful source to discuss. The sub-
mm photometry observations of BCR11 at the position of a
known radio source (Richards 2000) guarantees that we have
the correct association of a radio and sub-mm source, and
hence the correct SED to analyse. BCR11 allows us to confi-
dently derive the photometric redshift and accuracy that we
can expect from only the 850µm/1.4 GHz colour, which is
commonly used to measure redshifts of blank-field sub-mm
galaxies with follow-up radio observations (CY99, CY00).
Figs. 1,2 and 3, respectively, show the C − C − z and
P −z distributions for the multi-wavelength data of LH850.1,
and the comparison of the observed SED and template SEDs
(redshifted to the mode of the P − z distribution). Figs. 4,5
and 6 show the same distributions for BCR11. In both ex-
Figure 2. Discrete and cumulative redshift probability distribu-
tions of LH850.1. The six estimates, plotted with lines of differ-
ent style, correspond to the six evolutionary models introduced
in section 2.
Figure 3. The observed SED of LH850.1 normalised to the flux
density at 850µm is shown as squares and arrows. The arrows indi-
cate 3σ upper limits. The squares denote detection at a level ≥ 3σ,
with 1σ error bars. The template SEDs (lines) are redshifted to
z = 2.6, the mode of the redshift probability distribution of most
models considered (see table 1). The template SEDs at this red-
shift compatible within 3σ error bars with the SED of LH850.1
are displayed as darker lines (blue in the electronic version).
c(cid:13) 2002 RAS, MNRAS 000, 1 -- ??
Breaking the "Redshift Deadlock" -- II
5
Figure 4. Colour-flux-redshift (C −f −z) plot for BCR11 (Barger
et al. 2000). The description of the symbols is as in Fig. 1
Figure 5. Redshift probability distributions for BCR11. The de-
scription of the plot is as in Fig. 2
amples (Figs. 2 and 5) the peak of the redshift probability
distribution depends little on the details of the evolution-
ary model used to compute the mock catalogues. In both
cases, the models that invoke pure luminosity evolution (le1,
le1L13, le2, le2L13) have more extended wings, since they
provide a significant population of sources at redshifts be-
yond 3 that can reproduce the colours of the observed galax-
ies. On the other hand, the combined luminosity and den-
sity evolution models (led1, led2) have a very strong de-
cline in the density of sources at redshifts beyond 2 and 3.5,
and thus, the corresponding redshift distributions are much
more concentrated to values between 2 and 3. Model led1
is the most extreme of these two evolutionary models, since
the very strong negative luminosity and density evolution
does not allow any source beyond redshift z = 3.5 to be de-
tected at any reasonable depth. The concentration effect is
more dramatic on BCR11, where the colour constraints are
the weakest, and the wings of the redshift distribution are
strongly dependent on the prior information on the galaxy
population that has been used. The sharpness of this dis-
tribution is, thus, due to the prior, and not to the colour
constraints. In contrast, the sharpness of the distribution of
LH850.1 is intrinsic to the well determined colours of the
source, and is almost independent of the prior.
The comparison of LH850.1 and BCR11 demonstrates
the improved redshift accuracy that, in general, one can
expect for those sources with the greatest combination of
c(cid:13) 2002 RAS, MNRAS 000, 1 -- ??
Figure 6. Comparison of the observed SED of BCR11 with those
of template galaxies, redshifted to z = 2.7 (table 2). The descrip-
tion of the plot is as in Fig. 3. Note that the selection of SEDs to
represent BCR11 is different from that in LH850.1.
6
Itziar Aretxaga et al.
deep multi-wavelength data. In the case of BCR11, it also
illustrates the difficulty of constraining individual redshifts
with observations at 1.4 GHz and 850µm only. The redshift
probability distributions of both sources have very similar
modes (z = 2.6−2.8), yet for the less constraining evolution-
ary models (le1, le2, le1L13, le2L13) there remains 40 -- 50%
and 25 -- 35% probabilities that BCR11 has a redshift > 3
and > 4 respectively. The corresponding probabilities for
LH850.1 are 25-30% and 5% respectively. For models with a
strong decline in the number of sources with redshift (led1,
led2), the high-z tails are not produced. The sharpness of
the derived redshift distributions is due to the lack of bright
sources at high redshift, and not by the colours of the galaxy.
Fig. 7 shows the manner in which the redshift distribu-
tion derived for LH850.1 (using model le2) degrades as dif-
ferent observational constraints are successively taken into
account. When the 450, 850µm, 1.2mm and 1.4GHz detec-
tions are included in the analysis, the addition of upper lim-
its to the SED makes no difference to the derived redshift
distribution. It is also apparent that the distribution does
not change significantly whilst the detection at 450µm is
used in the analysis, in combination with at least one ad-
ditional band in the mm -- sub-mm and radio regime. How-
ever, as soon as the 450µm data is excluded, a high redshift
tail appears. This is due to the large intrinsic scatter in the
1.4GHz/850µm ratio amongst the template SEDs which, in
the absence of the 450µm data, allows galaxies over a wide
range of redshifts to satisfy the colour constraints. The elim-
ination of the radio data from the analysis allows more low-
redshift sources to be possible counterparts of the detected
source, and hence this flattens the resulting redshift distri-
bution.
The improvement in constraint on the photometric red-
shift is extremely important if it is to be used to deter-
mine the tuning of any broad-band mm -- cm spectrograph,
in an attempt to detect redshifted molecular CO-lines. This
is discussed further in section 5. Higher frequency radio ob-
servations at ∼ 5 − 15 GHz and deep 450µm observations,
including sensitive upper-limits, are essential ground-based
data that can provide the additional diagnostic power to
constrain redshifts (e.g. LH850.1).
3.2 Individual redshift distributions
Before we calculate the cumulative redshift-distribution for
the 850µm sources detected in the wide-area (> 200 sq. ar-
cmin) blank-field SCUBA surveys, which include the Lock-
man Hole and ELAIS N2 regions (Scott et al. 2002; Fox et
al. 2002), and the CUDSS fields (Lilly et al. 1999; Eales et
al. 2000; Webb et al. 2002), it is instructive to discuss some
of the typical redshift distributions that can be derived from
data of different quality (i.e. sensitivity and wavelength cov-
erage) for individual sources.
Therefore, we compare the redshift distributions of
wide-area blank-field sub-mm sources with those detected in
various lensing-cluster surveys (Smail et al. 1997; Smail et
al. 1999; Bertoldi et al. 2000; Chapman et al. 2002a). We also
consider radio, FIR and UV-selected objects with follow-up
sub-mm data (Barger et al. 2000; Scott et al. 2000; Chap-
man et al. 2000), and two bright sub-mm sources with ex-
tensive multi-wavelength observations, HDF850.1 (Hughes
et al. 1998; Downes et al. 1999; Dunlop et al. 2002), and the
extremely-red object HR10 (Dey et al. 1999).
Tables 1, 2, and 3 summarize the photometric redshifts
of all the above sources: the modes of their redshift distri-
butions, 68% and 90% confidence intervals derived from the
six models studied in this paper, and other relevant infor-
mation. Appendix A contains the C − C − z, C − f − z, P − z
diagrams and SEDs of the galaxies listed in Tables 1 and 2,
and also a selection of those listed in Table 3 for model le2,
which is one of the less constraining models considered and
serves as an illustration of the points made here.
When we have the most complete multi-wavelength
data-set (rest-frame radio -- sub-mm -- FIR,
including upper
limits) then photometric redshifts can be constrained to a
68% confidence band of width ∆z ∼ ±0.5 (Table 1 and
Fig.A1). Those sub-mm galaxies for which just two detec-
tions, and perhaps some shallow upper-limits are available,
do much worse, with accuracies at a 68% confidence level
of width ∆z >∼ ±1 (Table 2 and Figs. A2). A merit of the
Monte Carlo photometric-redshift method is that it can de-
termine redshift probability distributions even in the case
that the galaxies have just been detected in one band (Ta-
ble 3 and Fig. A3), the result being very flat redshift distri-
butions, that are highly dependent on the assumed evolu-
tionary model. This is especially important because half of
the sub-mm sources in the blank-field surveys belong to this
category, although it is worth noting that several of these
previously reported sub-mm sources have now been found
to be spurious (Ivison et al. 2002).
Often, the well constrained distributions are asymmet-
ric, skewed towards lower redshifts with a high-redshift tail
(e.g. LH850.1, N2850.2, HR10, ... in Figs. 2 and A1). This is
a natural consequence of the ability of the data (particularly
with a deep 450µm detection or upper limit, < 30 mJy) to
strongly reject the possibility of sub-mm galaxies lying at
z <∼ 2, whilst the same observational data cannot reject the
possibility that the sub-mm galaxies lie at a much higher
redshift, z ≫ 3. Examples of the power with which the
450µm detections or upper-limits can reject z <∼ 2 can be
seen in the SEDs of LH850.1 and HDF850.1 (Figs. 3 and
A1), respectively. On the other hand, the bright ISOPHOT
detections at 170µm (> 120mJy) make a strong case that the
sources lie at z <∼ 2, for example N1 -- 40, N1 -- 64 (Fig. A1) and
N1 -- 8 (Fig. A2), whilst the ISOPHOT non-detections are
too shallow to offer any additional constraints on the SEDs
of the majority of the sub-mm population (e.g. LH850.1,
HDF850.1 in Figs. 3 and A1).
It is also possible to obtain a bimodal redshift proba-
bility distribution for some galaxies (e.g. SMMJ00266+1708
in Fig. A1, BCR33 in Fig. A2), an effect created by two or
more template SEDs reproducing the same observed colours
at very different redshifts.
The figures in appendix A show the insensitivity of the
850µm/1.4GHz ratio to redshift when the sources are at
z >∼ 2 (a point also made by CY99, CY00). The redshift dis-
tributions in these cases are usually shallow, with long high-
redshift tails (e.g. BCR11, BCR13, MMJ154127+6615, in
Fig. A2), although exceptions also occur (SMMJ16403+4644
in Fig. 2A). The distributions are even broader in the cases
when only a sub-mm detection and a radio upper-limit are
available (Fig. A3). All these redshift distributions are nar-
rower for those evolutionary models with a strong decline
c(cid:13) 2002 RAS, MNRAS 000, 1 -- ??
Breaking the "Redshift Deadlock" -- II
7
Figure 7. Redshift distributions and compatible template SEDs for LH850.1, when different photometry bands are taken into account.
All estimates are based on a Monte Carlo produced with the same evolutionary model (le2), The SED models represented in the right-hand
panels have been redshifted to the mode of distributions represented in the left. Symbol and line codes are as in Fig. 3.
c(cid:13) 2002 RAS, MNRAS 000, 1 -- ??
8
Itziar Aretxaga et al.
beyond z ∼ 2 − 3, as can be seen from the measurements in
Tables 2 and 3, especially for lde1 which essentially contains
no sources at high-z.
A quick browse through Appendix A illustrates the va-
riety of template SEDs that reproduce the colours of an
individual sub-mm galaxy, which lends support to the use
of a template library in the derivation of the photometric
redshifts. A few sub-mm sources, however, still cannot be
reproduced at any redshift with the template SEDs avail-
able (e.g. LH850.12, SMMJ16403+46437). A possible expla-
nation for this problem is the contribution of AGN emission
to the radio-fluxes at a level not included in the SEDs of
the template library, or a mis-identification of a sub-mm
galaxy with a nearby radio-source (see also section 4). A
wider range of SEDs, including modest fractions of galaxies
with an excess of radio-emission, could partially solve this
problem but, on the other hand, it would also introduce an
extra error term in the redshift computation for the rest
of the sub-mm galaxies. Other sub-mm galaxies have only
a small representation of analogs among the mock galax-
ies (e.g. N1-40, SMMJ14009+0252), most probably because
they are rare objects in the Universe, either by their intrinsi-
cally high luminosity or their low redshift. This latter point
is illustrated by the rarity of mock galaxies in the corre-
sponding C − C − z or C − f − z diagrams, assuming any of
the evolutionary scenarios considered in this paper.
3.2.1 Accuracy of the photometric-redshift method
Fig. 8 shows the comparison of photometric and pub-
lished spectroscopic redshifts for 8 galaxies: in order of in-
creasing spectroscopic redshift, N1 -- 40, CUDSS14.18, N1 --
64,
HR10,
SMMJ14011+0252,
SMMJ02399−0136, and W−MM11.
SMMJ02399 -- 0134,
The two galaxies which depart most from the zphot =
zspec line are SMMJ02399-0134 and W−MM11. The first
has a photometric redshift calculated only on the basis of its
1.4 GHz/850µm colour (Fig. A2), and the latter on a 850µm
detection with a 450µm upper limit (Fig. A3). These two
objects illustrate the large confidence intervals that have to
be considered when the redshifts are derived from a single
colour. Specifically, SMMJ02399−0134 illustrates the uncer-
tainties attached to the large dispersion of 1.4GHz/850µm
colours implied from a local sample of galaxies. The agree-
ment between spectroscopic and photometric redshifts for
the remaining six galaxies (with more than one colour avail-
able) is remarkably good. The asymmetry of the error bars
is a reflection of the skewness of their respective P −z distri-
butions (Figs. A1, A2). These sources illustrate the impor-
tance of obtaining the description of the whole redshift dis-
tribution instead of relying on measurements of the median,
which departs significantly from the most probable redshift
in skewed distributions.
A larger sample of spectroscopic redshifts is obviously
essential to assess the statistical significance of the good-
ness of this technique, and the accuracy of the error bars
we derive. For the sources with more than one measured
colour, these errors are mainly the result of the scatter in
the shapes of the template SEDs. A larger sample of spec-
troscopic redshifts would allow us to fine-tune the selection
of SEDs and priors in order to trim the error bars if, for
instance, the modes of the distributions do not populate
Figure 8. Comparison of the photometric-redshift estimates and
the true redshifts of sub-mm sources with published optical/IR
spectroscopy. The diamonds represent the modes of the redshift
distributions for le2. The solid and dashed error bars represent
68% and 90% confidence intervals, respectively. Sources repre-
sented in black (in increasing redshift: N1 -- 40, CUDSS14.18, N1 --
64, HR10, SMMJ14011+0252, and SMMJ02399−0136) have pho-
tometric redshifts derived from a combination of colours, and are
the most precise. Sources represented in grey (in increasing red-
shift: SMMJ02399−0134, and W−MM11) have photometric red-
shifts derived from only one colour (or limit), and have shallow
probability distributions, still compatible with their spectroscopic
redshifts. Note that the asymmetry of the error-bars is a reflec-
tion of the asymmetry of the redshift probability distributions,
which in cases (CUDSS14.18 and HR10) are significantly skewed.
the envelope of error bars implied from the analysis. The
six best constrained redshift distributions (shown in black
in Fig. 7) do show a tendency to cluster more around the
zspec = zphot line than the collective extension of their error
bars. However, fine-tuning the inputs of the Monte-Carlo at
this early stage of confirmation of the photometric redshifts
is premature.
3.2.2 Notes on individual sources
LH850.1: Other redshift estimates are 2.95±1.49
0.98 based
on the 1.4GHz/850µm spectral index technique; and 2.72 ±
0.37 based on a χ2 minimization with one template SED
(YC02) . Our own estimate z = 2.6±0.4
0.5 (le2) is consistent
with these measurements.
LH850.12: No SED can reproduce the observed pho-
tometry of this source. Ivison et al. (2002) find evidence of
variability at 1.4GHz, and report associated X ray emission.
They conclude that LH850.12 might be a radio-loud QSO.
This kind of objects is actually not represented among our
SEDs, and thus the redshift we derive is unreliable. Note
that there are no counterpart mock-galaxies in the vicinity
of the error box of its C − C − z diagram in Fig. A1.
HDF850.1: Other redshift estimates are > 2.6, based
on the 1.4GHz/850µm spectral index technique; and 4.11 ±
c(cid:13) 2002 RAS, MNRAS 000, 1 -- ??
Breaking the "Redshift Deadlock" -- II
9
Table 1. Sources detected in three or more passbands at a ≥ 3σ level. Column 1 gives the name of the source; column 2 the
references for the SED data; columns 3-8 the mode of the photometric-redshift distributions and 68% confidence intervals for
different evolutionary models, and in parenthesis below, the corresponding 90% confidence intervals; and column 9 notes for each
object. The references are coded as follows: (1) Lutz et al. 2001, (2) Scott et al. 2002, (3) Ivison et al. 2002, (4) Downes et al. 2001,
(5) Dunlop et al. 2002; (6) Eales et al. 2000, (7) Smail et al. 1999, (8) Frayer et al. 2000, (9) Ivison et al. 1998, (10) Ivison et al.
2000, (11) Dey et al. 1999, (12) Chapman et al. 2002b. (13) Cowie et al. 2002 The A in the notes denotes the amplification factor
reported in the literature for the lensed sources.
zle1
phot
zle1L13
phot
zle2
phot
zle2L13
phot
zlde1
phot
zlde2
phot
notes
object
LH850.1
LH850.3
LH850.8
LH850.12
N2850.1
N2850.2
N2850.4
HDF850.1
CUDSS14.1
ref
1,2,3
2,3
2,3
2,3
2,3
2,3
2,3
4,5
6
SMMJ00266+1708
7,8
SMMJ02399−0136
8,9
SMMJ09429+4658
7,13
SMMJ14009+0252
10
SMMJ14011+0252
8
HR10
N1-40
N1-64
11
12
12
2.6±0.4
0.6
(2.0 -- 4.1)
2.7±0.2
(2.1 -- 3.0)
3.8±1.2
1.4
(2.0 -- 6.5)
0.6±0.3
0.1
(0.1 -- 1.0)
2.9±0.6
(2.0 -- 4.1)
2.3±0.6
0.3
(1.5 -- 3.4)
2.8±1.7
0.3
(1.3 -- 4.5)
4.2±0.5
0.7
(3.3 -- 5.5)
3.8±0.2
0.7
(2.5 -- 5.2)
2.7±2.2
0.2
(2.1 -- 5.5)
3.5±0.0
1.0
(2.3 -- 5.0)
2.8±0.7
(2.0 -- 4.5)
2.8±1.7
0.3
(2.0 -- 5.6)
4.4±0.7
0.9
(2.7 -- 5.5)
2.8±0.5
0.8
(2.0 -- 4.5)
1.5±1.0
0.0
(1.5 -- 3.1)
1.2 ± 0.2
(0.6 -- 1.5)
1.2±0.3
0.8
(0.5 -- 1.8)
2.6±0.4
0.5
(2.0 -- 3.8)
2.7±0.2
(1.8 -- 3.0)
3.8±1.3
1.2
(1.6 -- 6.0)
0.6±0.3
0.1
(0.1 -- 1.0)
2.8±0.7
0.4
(2.0 -- 3.7)
2.3±0.5
0.3
(1.5 -- 3.2)
2.7±0.7
(2.0 -- 4.5)
4.1±0.6
(3.1 -- 5.0)
3.8±0.8
0.7
(2.5 -- 5.5)
2.7±2.0
0.2
(2.2 -- 5.5)
3.1±0.8
0.6
(2.0 -- 4.5)
2.9±0.6
(2.0 -- 4.4)
2.8±0.6
(1.9 -- 4.0)
4.1±0.8
0.6
(3.0 -- 5.1)
2.8±0.5
0.8
(2.0 -- 4.5)
1.5±0.5
0.0
(1.5 -- 2.5)
0.7±0.6
0.1
(0.5 -- 1.5)
1.2±0.3
(0.5 -- 1.5)
2.6±0.4
0.5
(2.0 -- 3.7)
3.0±0.0
0.5
(2.1 -- 3.0)
3.7±1.5
0.7
(2.3 -- 6.0)
0.2±0.8
0.2
(0.0 -- 1.3)
2.8±0.7
0.5
(2.0 -- 4.1)
2.3±0.6
0.3
(1.5 -- 3.5)
2.9±1.2
0.9
(1.2 -- 4.5)
4.1±0.6
(3.0 -- 5.1)
3.8±0.7
0.8
(2.0 -- 4.9)
2.7±1.6
0.7
(2.0 -- 5.4)
3.9±0.9
(2.5 -- 5.0)
2.8±1.3
0.3
(2.0 -- 5.0)
2.8±1.5
0.3
(2.0 -- 4.9)
4.1±0.8
(3.1 -- 5.0)
2.7±0.6
0.7
(2.0 -- 4.4)
1.5±0.8
0.0
(1.5 -- 2.9)
0.6±0.6
0.1
(0.5 -- 1.5)
1.2±0.8
0.3
(0.5 -- 2.1)
2.6±0.4
0.5
(2.0 -- 3.9)
2.5±0.0
1.3
(1.0 -- 3.0)
3.8±1.2
0.8
(2.0 -- 5.3)
0.2±0.6
0.2
(0.0 -- 1.2)
2.8±0.3
0.8
(2.0 -- 5.3)
2.3±0.5
0.3
(1.5 -- 3.0)
2.8±0.9
0.8
(1.6 -- 4.5)
4.0±0.7
0.5
(3.0 -- 5.0)
3.7±0.8
(2.2 -- 5.0)
2.8±2.2
0.2
(1.0 -- 5.3)
3.1±0.7
0.6
(2.5 -- 4.6)
2.8±0.7
0.5
(2.0 -- 4.3)
2.8±0.7
0.4
(1.9 -- 4.0)
4.1±0.8
(2.5 -- 5.2)
2.7±0.6
0.7
(2.0 -- 4.5)
1.5±0.7
0.0
(1.5 -- 2.6)
0.6±0.2
(0.3 -- 1.0)
1.2±0.3
0.6
(0.5 -- 1.8)
2.7±0.3
(2.0 -- 3.1)
2.9±0.6
1.2
(1.0 -- 3.5)
3.5±0.0
0.6
(2.6 -- 3.5)
0.7±0.2
(0.5 -- 1.0)
2.9±0.3
0.4
(2.4 -- 3.5)
2.3±0.3
(2.0 -- 3.4)
2.8±0.5
0.3
(2.2 -- 3.5)
3.5±0.0
0.6
(2.7 -- 3.5)
2.9±0.4
(2.1 -- 3.5)
2.8±0.5
0.3
(2.2 -- 3.5)
3.5±0.0
0.8
(2.2 -- 3.5)
2.8±0.4
0.3
(2.0 -- 3.5)
3.0±0.5
0.1
(2.3 -- 3.5)
3.5±0.0
0.9
(2.1 -- 3.5)
2.7±0.6
0.2
(2.3 -- 3.5)
1.5±1.0
0.0
(1.5 -- 3.1)
0.7±0.2
(0.5 -- 1.0)
1.2±0.3
0.6
( -- )
2.7±0.3
0.2
(2.3 -- 3.5)
3.0±0.0
0.5
(2.1 -- 3.0)
3.0±0.0
0.9
(2.0 -- 4.3)
1.0±0.5
(0.8 -- 1.5)
2.7±0.3
0.4
(2.0 -- 3.5)
2.6±0.4
0.1
(2.1 -- 3.0)
2.7±0.3
0.7
(1.5 -- 3.4)
4.0±0.5
0.6
(3.2 -- 5.0)
2.8±1.1
0.8
(2.0 -- 4.8)
2.9±0.5
0.4
(2.5 -- 5.3)
3.8±1.7
zspec = 2.8, A = 2.4,
0.8
(2.4 -- 5.0) AGN, 3.4cm too low?
2.8±0.8
0.3
(2.0 -- 4.4)
2.7±0.3
0.6
(2.0 -- 4.3)
4.1±0.4
1.2
(2.0 -- 5.0) AGN contribution?
2.7±0.3
zspec = 2.57, A = 3.
0.4
(2.0 -- 3.5)
1.5±0.7
0.0
(1.5 -- 2.4)
1.0±0.5
0.0
(0.5 -- 1.5)
1.2±0.6
0.2
(0.7 -- 2.0)
zspec = 1.44
zspec = 0.45
A = 1.5
A = 1.5
A = 3
A = 1.6
zspec = 0.91
radio-loud QSO(3), zphot is
unreliable
disregarding the 3.4cm upper limit
0.51 based on a χ2 minimization with one template SED
(YC02). Our own estimate z = 4.1±0.6
0.5 (le2) is in good
agreement with these measurements, and is unaffected by
the inclusion or exclusion of the possible lensing amplifica-
tion experienced by this source (A ≈ 3, Dunlop et al. 2002).
CUDSS14.1: Other published redshift estimates are
2.01±1.10
0.71, based on the 1.4GHz/850µm spectral index tech-
nique; and 2.06 ± 0.31 based on a χ2 minimization with
one template SED (YC02). Both these estimates are in the
low-redshift tail of our redshift distribution, which at a 68%
significance level places this object at z = 3.8±0.7
0.8 (le2).
CUDSS14.18: The inclusion of the 450 and 850µm
fluxes as detections at 2.3 and 2.8σ (instead of upper limits)
gives a similar distribution to the one shown in Figure A2.
Our redshift estimate places this object at z = 0.7±1.3
0.2 with
a 68% confidence level (le2), in good agreement with its true
spectroscopic redshift, zspec = 0.66. Other redshift estimates
c(cid:13) 2002 RAS, MNRAS 000, 1 -- ??
are 1.12±0.53
0.45, based on the 1.4GHz/850µm spectral index
technique; and 1.11 ± 0.21 based on a χ2 minimization with
one template SED (YC02).
SMMJ00266+1708: Other redshift estimates are
3.49±2.03
1.23 based on the 1.4GHz/850µm spectral index tech-
nique; and 3.50 ± 0.45 based on a χ2 minimization with one
template SED (YC02). Our estimate z = 2.7±2.3
0.2 (le2) is
consistent with these measurements. The redshift distribu-
tion is bimodal, due to the range of SEDs used in the redshift
estimation technique.
SMMJ02399−0136: The complete SED reported in
Frayer et al. (2000), Ivison et al. (1998) is inconsistent with
the SEDs of our template library at any redshift. This is
reported to be an AGN (Smail et al. 1999). One possible
explanation for the failure to detect SMMJ02399-0136 at
3.4cm is a high-frequency break (between 5.5cm and 0.9cm)
in the synchrotron spectrum. Such sharp spectral features
10
Itziar Aretxaga et al.
Table 2. Sources detected in two pass-bands at ≥ 3σ level. The columns are identical to those described in
Table 1. The reference codes are as follows: (1) Scott et al. 2002, (2) Ivison et al. 2002 (3) Barger et al. 2000,
(4) Eales et al. 2000, (5) Webb et al. 2002, (6) Smail et al. 1999, (7) Chapman et al. 2002a, (8) Bertoldi et
al. 2000, (9) Dannerbauer et al. 2002, (10) Chapman et al. 2002b.
object
LH850.6
LH850.7
LH850.14
LH850.16
LH850.18
N2850.7
N2850.13
BCR7
BCR11
BCR13
BCR33
BCR49
CUDSS14.3
CUDSS14.9
CUDSS14.18
CUDSS3.1
CUDSS3.8
CUDSS3.9
CUDSS3.10
CUDSS3.15
SMMJ02399−0134
SMMJ10236+0412
SMMJ10237+0410
SMMJ16403+4644
SMMJ16403+46440
SMMJ16403+46437
MMJ154127+6615
MMJ154127+6616
MMJ120546−0741.5
MMJ120539−0745.4
ref
1,2
1,2
1,2
1,2
1,2
1,2
1,2
3
3
3
3
3
4
4
4
5
5
5
5
5
6
7
7
7
7
7
8
8
9
9
N1−008
10
zle1
phot
zle1L13
phot
zle2
phot
zle2L13
phot
zlde1
phot
zlde2
phot
notes
2.7±2.0
0.7
(2.0 -- 6.4)
3.6±1.9
0.8
(1.6 -- 5.5)
2.5±1.9
0.4
(1.5 -- 6.3)
3.6±1.0
1.6
(1.5 -- 6.0)
3.2±1.1
(1.5 -- 5.8)
2.8±1.5
0.8
(1.9 -- 5.5)
2.7±0.8
1.0
(1.5 -- 5.0)
2.7±0.8
1.1
(1.5 -- 4.8)
2.6±0.9
(1.5 -- 4.7)
2.3±1.9
0.3
(1.5 -- 5.5)
2.9±1.0
(1.5 -- 5.2)
2.3±1.1
0.5
(1.6 -- 4.5)
3.1±0.9
(1.7 -- 5.0)
2.6±0.8
0.6
(1.5 -- 4.1)
0.7±0.8
0.4
(0.0 -- 1.8)
2.1±0.4
0.6
(1.0 -- 2.5)
2.1±0.9
0.2
(1.4 -- 3.0)
1.0±0.2
0.5
(0.5 -- 1.5)
2.6±0.8
(1.4 -- 4.5)
2.2±0.7
(1.0 -- 3.7)
2.2±0.7
(1.0 -- 3.6)
2.3±0.8
(1.3 -- 4.0)
2.0±1.1
0.5
(1.0 -- 3.5)
2.2±0.8
(0.5 -- 3.1)
2.8±0.7
(1.5 -- 4.1)
2.2±0.8
0.6
(1.0 -- 3.3)
2.8±1.7
0.5
(2.0 -- 6.0)
2.8±1.7
0.3
(2.0 -- 6.0)
4.7±1.3
1.9
(2.5 -- 7.8)
2.8±1.7
0.6
(2.0 -- 6.3)
0.0±0.5
0.0
(0.0 -- 0.5)
2.7±2.0
0.7
(2.0 -- 6.2)
3.5±0.5
1.8
(1.5 -- 5.1)
2.4±2.0
0.4
(1.5 -- 5.5)
3.5±1.0
1.5
(1.5 -- 6.1)
3.4±1.0
1.4
(1.5 -- 5.6)
2.8±1.2
0.6
(2.0 -- 5.3)
3.0±0.5
1.5
(1.5 -- 5.0)
2.7±0.8
1.0
(1.5 -- 4.7)
2.7±0.8
(1.5 -- 4.7)
2.3±2.0
0.3
(1.5 -- 5.6)
3.0±1.0
(1.5 -- 5.3)
2.4±0.5
(1.5 -- 4.2)
3.1±0.9
(1.5 -- 4.8)
2.6±0.9
0.6
(1.5 -- 4.0)
0.7±0.8
0.4
(0.0 -- 1.8)
1.8±0.7
0.3
(0.7 -- 2.5)
2.0±0.5
(1.4 -- 3.0)
1.0±0.4
0.0
(1.0 -- 2-0)
2.6±0.9
(1.3 -- 4.5)
2.3±0.7
(1.5 -- 4.0)
2.3±0.8
(1.4 -- 4.0)
2.3±0.8
(1.3 -- 4.0)
2.0±1.2
0.5
(1.0 -- 3.7)
2.2±0.8
(0.8 -- 3.5)
2.8±0.7
(1.5 -- 4.1)
2.3±0.7
0.8
(1.1 -- 3.5)
3.6±0.9
1.4
(2.0 -- 5.9)
2.8±2.0
0.3
(2.0 -- 6.0)
3.4±2.7
1.9
(2.5 -- 7.0)
2.8±1.7
0.8
(2.0 -- 6.0)
0.0±0.5
0.0
(0.0 -- 0.5)
2.7±1.8
0.7
(2.0 -- 5.4)
3.6±0.4
1.9
(1.5 -- 5.3)
2.4±1.9
0.4
(2.0 -- 5.4)
3.3±1.1
1.3
(2.0 -- 5.5)
3.0±1.1
1.0
(1.5 -- 4.9)
3.0±1.0
(1.5 -- 4.9)
2.7±0.8
1.0
(1.5 -- 4.6)
2.2±1.1
0.7
(1.5 -- 4.8)
2.7±0.8
0.9
(1.5 -- 4.6)
2.3±1.7
0.3
(1.5 -- 4.7)
3.0±1.0
(1.5 -- 4.6)
2.4±1.1
0.6
(1.0 -- 4.0)
3.2±0.8
1.0
(2.0 -- 5.0)
2.5±0.5
1.0
(1.5 -- 4.5)
0.7±0.8
0.4
(0.0 -- 1.8)
1.7±0.7
0.2
(1.5 -- 2.5)
2.0±0.8
0.5
(1.0 -- 3.0)
1.1±0.8
0.1
(0.5 -- 2.0)
2.5±1.0
(1.5 -- 4.5)
2.2±0.9
0.7
(1.5 -- 4.0)
2.3±0.8
(1.0 -- 3.8)
2.3±0.8
(1.5 -- 4.0)
2.1±1.0
0.6
(1.0 -- 3.6)
2.1±0.4
1.1
(1.0 -- 3.5)
2.7±0.8
(1.5 -- 4.1)
2.6±0.4
1.1
(1.0 -- 3.3)
2.8±1.7
0.3
(2.0 -- 5.0)
2.8±1.8
0.3
(2.5 -- 5.0)
3.8±2.1
0.3
(3.0 -- 6.0)
2.8±1.7
0.3
(2.0 -- 4.9)
0.0±0.4
0.0
(0.0 -- 0.5)
2.7±1.8
0.7
(2.0 -- 5.4)
3.5±0.5
1.9
(1.5 -- 5.3)
2.3±2.0
0.3
(2.0 -- 5.4)
3.6±0.8
1.5
(2.0 -- 5.4)
3.2±1.0
1.2
(1.8 -- 4.9)
3.0±0.9
(1.7 -- 5.0)
2.8±0.7
1.2
(1.5 -- 4.8)
2.7±0.6
1.2
(1.5 -- 4.7)
2.6±0.7
1.1
(1.5 -- 4.7)
2.3±1.7
0.3
(1.5 -- 4.7)
3.0±1.0
(1.5 -- 4.5)
2.5±1.0
0.5
(1.0 -- 3.7)
3.1±0.9
(1.7 -- 5.0)
2.6±0.4
1.1
(1.5 -- 4.0)
0.7±1.1
0.2
(0.0 -- 2.0)
1.8±0.5
0.3
(0.9 -- 2.5)
2.0±0.5
(1.2 -- 3.0)
0.7±0.3
0.5
(0.0 -- 1.4)
2.5±0.9
(1.0 -- 4.2)
2.3±0.7
(1.0 -- 3.7)
2.3±0.8
(1.0 -- 3.7)
2.3±0.8
(1.3 -- 4.0)
2.0±1.1
0.5
(1.0 -- 3.6)
2.2±0.7
(0.5 -- 3.5)
2.8±0.7
(1.5 -- 4.1)
2.6±0.4
1.0
(1.2 -- 3.5)
3.7±0.8
1.2
(2.0 -- 5.0)
2.8±2.2
0.3
(2.1 -- 5.0)
3.8±2.0
0.3
(2.5 -- 6.0)
2.5±1.7
0.5
(2.0 -- 5.3)
0.0±0.4
0.0
(0.0 -- 0.5)
2.7±0.6
0.2
(2.3 -- 3.5)
3.5±0.0
1.0
(1.7 -- 3.5)
2.7±0.3
0.5
(2.0 -- 3.4)
2.8±0.5
0.3
(2.0 -- 3.5)
3.2±0.3
0.7
(2.0 -- 3.5)
3.0±0.3
0.5
(2.5 -- 3.5)
2.8±0.5
0.3
(2.0 -- 3.5)
2.8±0.5
0.3
(1.9 -- 3.5)
2.7±0.6
0.2
(2.1 -- 3.5)
2.8±0.7
0.3
(2.2 -- 3.5)
3.0±0.5
0.3
(2.0 -- 3.5)
2.7±0.3
0.7
(1.7 -- 3.5)
3.5±0.0
0.8
(2.2 -- 3.5)
2.7±0.8
0.2
(2.0 -- 3.4)
0.7±0.7
(0.0 -- 1.9)
1.9±0.4
(1.5 -- 2.6)
2.6±0.4
0.2
(2.0 -- 3.0)
1.0±0.4
0.0
(1.0 -- 1.5)
2.7±0.8
0.2
(1.9 -- 3.5)
2.5±0.5
(1.9 -- 3.5)
2.6±0.7
0.5
(1.7 -- 3.5)
2.7±0.3
0.7
(2.0 -- 3.5)
2.6±0.6
(1.7 -- 3.5)
2.4±0.6
0.8
(1.2 -- 3.5)
2.8±0.6
0.3
(2.1 -- 3.5)
2.6±0.5
(1.8 -- 3.5)
2.7±0.3
(2.0 -- 3.1)
3.0±0.5
0.4
(2.2 -- 3.5)
3.0±0.2
0.5
(2.5 -- 3.4)
2.5±0.3
(2.0 -- 3.3)
0.0±0.4
0.0
(0.0 -- 0.5)
2.8±0.2
0.6
(2.0 -- 4.4)
2.5±0.2
0.5
(1.5 -- 3.8)
3.0±0.0
0.7
(2.0 -- 3.5)
3.5±0.0
0.8
(1.5 -- 3.5)
3.0±0.0
0.7
(2.0 -- 3.3)
2.5±0.5
(2.0 -- 4.3)
2.5±0.3
0.5
(2.0 -- 3.5)
2.4±0.4
(1.5 -- 3.2)
2.4±0.4
(1.5 -- 3.0)
2.9±0.1
0.6
(2.0 -- 3.3)
3.0±0.0
0.7
(2.0 -- 3.0)
2.8±0.2
0.6
(1.6 -- 3.0)
3.0±0.0
0.8
(2.2 -- 4.1)
2.8±0.2
0.6
(1.9 -- 3.5)
1.2±0.5
0.7
(0.2 -- 2.0)
1.8±0.4
0.3
(1.2 -- 2.5)
1.9±0.4
(1.2 -- 2.5)
1.3±0.5
0.3
(0.6 -- 2.0)
2.5±0.0
0.9
(1.0 -- 2.5)
2.5±0.0
0.7
(1.2 -- 2.5)
2.5±0.0
0.8
(1.2 -- 2.5)
2.8±0.2
0.5
(1.7 -- 3.0)
2.4±0.4
(1.6 -- 3.0)
2.5±0.4
(1.6 -- 3.0)
2.9±0.1
0.7
(1.5 -- 3.1)
2.5±0.5
(1.5 -- 3.0)
2.8±0.3
(2.0 -- 4.1)
2.9±1.0
0.4
(2.6 -- 5.3)
3.5±1.0
(2.5 -- 6.0)
2.8±0.2
0.6
(2.0 -- 3.7)
0.0±0.4
0.0
(0.0 -- 0.5)
zspec = 0.66
high 1.4GHz, AGN?
A = 2.5, zspec = 1.06
A = 3.2
A = 3.7
A = 3.0
A = 3.6
A = 3.2, 1.4GHz high, AGN?
A = 1
A = 1
c(cid:13) 2002 RAS, MNRAS 000, 1 -- ??
Breaking the "Redshift Deadlock" -- II
11
Table 3. Sources detected at 850µm or 1.2mm at a ≥ 3.5σ level, and undetected at any other wavelength.
The columns are identical to those described in Table 1. The reference codes are as follows: (1) Scott et
al. 2002, (2) Eales et al. 2000, (3) Webb et al. 2002, (4) Smail et al. 1999, (5) Chapman et al. 2002a, (6)
Dannerbauer et al. 2002. (7) Ivison et al. 2002
object
LH850.2
LH850.4
LH850.5
LH850.13
N2850.3
N2850.5
N2850.6
N2850.8
N2850.9
N2850.10
N2850.11
CUDSS14.2
CUDSS14.4
CUDSS14.5
CUDSS14.6
CUDSS14.8
CUDSS14.10
CUDSS3.2
CUDSS3.3
CUDSS3.4
CUDSS3.5
CUDSS3.6
CUDSS3.7
CUDSS3.11
CUDSS3.12
CUDSS3.13
CUDSS3.16
SMMJ02400−0134
SMMJ04431+0210 (N4)
SMMJ04541−0302
c(cid:13) 2002 RAS, MNRAS 000, 1 -- ??
ref
1,7
1,7
1,7
1,7
1,7
1,7
1,7
1,7
1,7
1,7
1,7
2
2
2
2
2
2
3
3
3
3
3
3
3
3
3
3
4
4
5
zle1
phot
zle1L13
phot
zle2
phot
zle2L13
phot
zlde1
phot
zlde2
phot
notes
5.9±1.3
2.4
(3.5 -- 9.0)
5.5±2.0
1.8
(3.0 -- 8.8)
5.3±2.2
1.3
(3.3 -- 9.0)
4.9±2.6
1.4
(3.0 -- 8.8)
4.8±2.8
0.8
(3.5 -- 9.0)
5.0±2.0
(2.5 -- 8.6)
5.0±2.0
(2.5 -- 8.7)
4.0±2.7
1.5
(2.1 -- 8.5)
5.2±2.0
1.7
(3.0 -- 9.0)
4.7±2.3
1.9
(2.1 -- 8.5)
5.5±1.9
(3.0 -- 8.9)
5.2±2.4
1.7
(2.5 -- 8.7)
5.0±2.6
1.5
(2.5 -- 8.8)
4.8±2.6
1.3
(2.8 -- 9.0)
4.7±2.8
1.5
(2.5 -- 8.8)
4.7±2.6
1.7
(2.5 -- 9.0)
4.7±2.6
1.7
(2.5 -- 9.0)
4.5±1.8
2.2
(2.0 -- 8.5)
3.8±2.5
1.5
(2.0 -- 8.4)
3.2±2.9
1.2
(2.0 -- 8.5)
3.7±2.6
1.7
(1.8 -- 8.5)
3.7±2.6
1.7
(2.0 -- 8.7)
3.7±2.6
1.2
(2.0 -- 8.4)
3.8±3.0
1.3
(1.9 -- 8.5)
3.7±2.6
1.7
(1.9 -- 8.5)
4.0±2.3
2.0
(1.8 -- 8.5)
3.5±2.9
1.5
(1.6 -- 8.5)
5.0±2.6
1.5
(3.0 -- 9.1)
4.5±2.5
1.6
(2.5 -- 8.8)
4.7±2.0
2.2
(2.2 -- 8.5)
5.5±1.5
1.9
(3.3 -- 8.5)
5.3±2.2
1.4
(3.0 -- 8.5)
6.1±1.5
1.7
(3.0 -- 8.5)
5.1±2.1
1.6
(3.0 -- 8.4)
4.6±2.9
0.6
(3.5 -- 8.8)
4.7±2.3
1.4
(2.7 -- 8.5)
4.8±1.9
2.3
(2.5 -- 8.3)
4.7±2.0
2.2
(2.0 -- 8.4)
5.0±2.0
1.8
(2.8 -- 8.5)
4.5±2.5
1.7
(2.1 -- 8.5)
4.62.9
0.9
(3.0 -- 8.7)
5.0±2.5
1.6
(2.5 -- 8.8)
5.1±2.5
1.6
(2.5 -- 8.8)
4.9±2.7
1.4
(2.8 -- 9.0)
4.7±3.0
1.2
(2.5 -- 8.8)
4.7±2.0
2.2
(2.5 -- 9.0)
4.8±2.5
1.8
(2.5 -- 9.0)
3.8±2.7
1.6
(1.9 -- 8.5)
3.7±2.4
1.7
(2.0 -- 8.4)
3.7±2.8
1.7
(2.0 -- 8.3)
3.7±2.6
1.4
(1.8 -- 8.5)
4.0±2.4
2.0
(2.0 -- 8.8)
3.7±2.6
1.2
(2.0 -- 8.1)
4.5±1.7
2.5
(2.0 -- 8.5)
3.2±2.1
1.2
(1.9 -- 8.5)
3.2±2.2
1.5
(1.7 -- 8.5)
3.2±2.2
1.2
(1.6 -- 8.5)
5.2±2.3
1.7
(3.0 -- 9.1)
4.6±2.2
1.6
(2.5 -- 8.7)
4.0±2.6
1.5
(2.3 -- 8.5)
5.8±0.2
1.7
(3.3 -- 6.0)
4.7±1.3
0.5
(3.5 -- 6.0)
4.7±1.3
0.5
(3.4 -- 6.0)
5.0±1.0
(3.0 -- 6.0)
5.8±0.2
1.4
(3.6 -- 6.0)
5.0±0.7
1.5
(2.9 -- 6.0)
4.8±0.8
1.3
(3.0 -- 6.0)
4.5±1.0
1.5
(2.6 -- 6.0)
5.9±0.1
1.8
(3.2 -- 6.0)
4.7±0.8
1.6
(2.6 -- 6.0)
5.0±1.0
(3.2 -- 6.0)
4.7±0.8
1.0
(2.9 -- 6.0)
4.6±1.4
0.8
(2.9 -- 6.0)
4.8±1.2
0.8
(3.0 -- 6.0)
5.2±0.8
1.3
(2.9 -- 6.0)
4.5±1.3
1.0
(2.6 -- 6.0)
3.9±2.0
0.4
(2.6 -- 6.0)
3.5±1.7
1.0
(2.0 -- 5.8)
4.0±1.2
1.5
(2.2 -- 6.0)
4.3±0.9
1.2
(2.3 -- 6.0)
3.8±1.5
1.3
(2.0 -- 5.9)
3.7±1.5
1.2
(2.0 -- 5.8)
4.0±1.5
1.2
(2.3 -- 6.0)
3.8±1.7
1.0
(2.3 -- 6.0)
3.5±1.8
1.0
(2.0 -- 5.8)
3.5±1.8
1.0
(2.0 -- 5.8)
3.7±1.5
1.2
(2.1 -- 6.0)
4.9±1.1
0.9
(3.1 -- 6.0)
4.5±1.4
1.0
(2.7 -- 6.0)
4.7±1.2
(2.6 -- 6.0)
5.1±0.4
1.5
(3.3 -- 6.0)
4.6±1.4
0.5
(3.3 -- 6.0)
6.0±0.0
1.9
(3.4 -- 6.0)
5.0±1.0
(3.3 -- 6.0)
4.8±0.9
(3.6 -- 6.0)
4.5±1.5
0.6
(2.9 -- 6.0)
4.8±1.2
(2.5 -- 6.0)
4.8±1.2
(2.6 -- 6.0)
5.0±1.0
(3.1 -- 6.0)
4.8±1.2
(2.6 -- 6.0)
4.81.2
0.8
(3.2 -- 6.0)
5.0±0.7
1.5
(2.8 -- 6.0)
4.7±1.3
0.9
(2.8 -- 6.0)
5.3±0.7
1.2
(3.0 -- 6.0)
5.2±0.8
1.3
(2.9 -- 6.0)
4.7±1.9
1.2
(2.6 -- 6.0)
4.3±1.2
(2.6 -- 6.0)
3.9±1.4
(2.2 -- 6.0)
4.0±1.3
1.5
(2.0 -- 5.8)
3.9±1.4
(2.2 -- 6.0)
3.9±1.4
(2.0 -- 5.8)
3.9±1.4
(2.0 -- 5.8)
3.7±1.3
(2.0 -- 5.7)
3.7±1.8
0.9
(2.0 -- 5.7)
3.9±1.3
(2.0 -- 5.8)
3.9±1.4
(2.1 -- 6.0)
3.5±1.9
1.0
(2.0 -- 5.9)
5.2±0.8
1.2
(3.1 -- 6.0)
4.2±1.6
0.7
(2.7 -- 6.0)
4.7±0.8
1.6
(2.6 -- 6.0)
3.5±0.0
0.5
(2.7 -- 3.5)
3.5±0.0
0.6
(2.6 -- 3.5)
3.5±0.0
0.6
(2.6 -- 3.5)
3.5±0.0
0.6
(2.6 -- 3.5)
3.5±0.0
0.5
(2.7 -- 3.5)
3.5±0.0
0.7
(2.5 -- 3.5)
3.5±0.0
0.8
(2.5 -- 3.5)
3.0±0.5
0.3
(2.3 -- 3.5)
3.5±0.0
0.6
(2.6 -- 3.5)
3.5±0.0
0.8
(2.3 -- 3.5)
3.5±0.0
0.7
(2.6 -- 3.5)
3.5±0.0
0.7
(2.5 -- 3.5)
3.5±0.0
0.8
(2.4 -- 3.5)
3.5±0.0
0.8
(2.5 -- 3.5)
3.5±0.0
0.8
(2.5 -- 3.5)
3.5±0.0
0.8
(2.3 -- 3.5)
2.9±0.5
(2.3 -- 3.5)
2.9±0.5
(2.2 -- 3.5)
2.9±0.4
(2.1 -- 3.5)
2.9±0.4
(2.2 -- 3.5)
2.9±0.5
(2.1 -- 3.5)
2.9±0.4
(2.2 -- 3.5)
2.8±0.5
0.3
(2.0 -- 3.5)
2.9±0.4
(2.2 -- 3.5)
2.9±0.5
(2.2 -- 3.5)
2.9±0.5
(2.1 -- 3.5)
2.9±0.4
(2.2 -- 3.5)
3.5±0.0
0.7
(2.5 -- 3.5)
3.5±0.0
0.7
(2.5 -- 3.5)
3.0±0.5
0.3
(2.4 -- 3.5)
4.3±1.2
0.8
(3.2 -- 6.0)
3.9±1.7
0.4
(3.0 -- 6.0)
4.5±1.0
(3.0 -- 6.0)
4.6±1.2
(2.7 -- 6.0)
4.3±1.0
(3.2 -- 6.0)
2.5±2.0
0.5
(3.0 -- 6.0)
2.5±1.9
0.5
(2.0 -- 5.6)
2.5±0.3
1.0
(2.5 -- 6.0)
3.7±1.5
0.7
(2.5 -- 6.0)
2.5±0.5
1.0
(1.5 -- 4.8)
4.1±1.2
(2.6 -- 6.0)
2.5±1.2
1.0
(1.5 -- 5.1)
2.5±1.2
1.0
(1.5 -- 5.2)
2.5±1.9
0.5
(2.0 -- 5.7)
2.5±1.4
1.0
(1.5 -- 5.3)
2.5±0.8
1.0
(1.5 -- 5.0)
2.5±0.5
1.0
(1.5 -- 4.8)
2.5±0.0
0.9
(1.0 -- 4.0)
2.5±0.0
0.9
(1.0 -- 4.0)
2.5±0.0
1.0
(1.0 -- 4.0)
2.5±0.0
1.0
(1.0 -- 3.9)
2.5±0.0
1.0
(1.0 -- 3.9)
2.5±0.0
1.0
(1.4 -- 4.5)
2.5±0.0
1.0
(1.0 -- 4.1)
2.5±0.0
1.0
(1.0 -- 3.9)
2.5±0.0
1.0
(1.0 -- 4.0)
2.5±0.0
1.0
(1.0 -- 3.8)
4.0±0.7
2.0
(2.0 -- 5.5)
2.4±1.0
(1.5 -- 4.9)
2.5±0.6
1.0
(1.5 -- 4.7)
A > 1.9 (2.5?)
A > 1.
A = 2.6
12
Itziar Aretxaga et al.
Table 3. (cont.)
object
ref
zle1
phot
zle1L13
phot
zle2
phot
zle2L13
phot
zlde1
phot
zlde2
phot
notes
SMMJ04542−0301
SMMJ04543+0256
SMMJ04543+0257
SMMJ10237+0412
SMMJ14573+2220
SMMJ17223+3207
SMMJ22471−0206
MMJ120517−0743.1
W−MM11
5
5
5
5
5
5
4
6
5
4.7±2.2
(2.5 -- 8.9)
3.9±2.7
1.4
(2.0 -- 8.4)
4.7±2.2
1.7
(2.0 -- 8.2)
4.0±2.9
1.5
(2.0 -- 8.7)
4.0±2.7
1.5
(2.0 -- 8.5)
4.0±2.5
1.5
(2.0 -- 8.4)
4.8±2.4
1.8
(2.5 -- 8.8)
7.0±1.5
2.4
(4.3 -- 10.)
4.2±2.0
2.2
(2.0 -- 8.7)
4.3±2.7
1.5
(2.1 -- 8.5)
3.8±2.7
1.6
(2.0 -- 8.5)
4.5±2.1
1.5
(2.5 -- 8.3)
3.9±3.0
1.4
(2.0 -- 8.7)
4.2±2.6
1.7
(2.0 -- 8.6)
4.2±2.6
1.7
(2.0 -- 8.4)
4.8±2.3
1.8
(2.5 -- 8.8)
7.1±1.4
2.6
(4.3 -- 10.)
3.4±2.9
1.5
(2.0 -- 8.6)
4.2±1.7
0.7
(2.7 -- 6.0)
4.0±1.5
1.1
(2.4 -- 6.0)
4.2±1.3
1.1
(2.6 -- 6.0)
4.5±0.7
2.0
(2.0 -- 6.0)
4.0±1.6
1.0
(2.0 -- 6.0)
3.9±1.6
1.0
(2.4 -- 6.0)
4.7±1.3
0.9
(2.8 -- 6.0)
6.0±0.0
1.6
(3.6 -- 6.0)
3.7±1.6
1.2
(2.0 -- 5.8)
4.5±1.4
1.0
(2.6 -- 6.0)
4.0±1.5
1.1
(2.0 -- 6.0)
4.5±1.0
1.4
(2.6 -- 6.0)
4.7±0.8
1.8
(2.6 -- 6.0)
4.2±1.8
0.8
(2.5 -- 6.0)
4.0±1.1
1.5
(2.5 -- 6.0)
4.9±1.1
(2.8 -- 6.0)
6.0±0.0
1.6
(3.5 -- 6.0)
3.9±1.6
1.2
(2.2 -- 6.0)
3.3±0.2
0.6
(2.4 -- 3.5)
2.9±0.4
(2.0 -- 3.5)
3.5±0.0
0.8
(2.3 -- 3.5)
2.9±0.4
(2.2 -- 3.5)
2.9±0.4
(2.2 -- 3.5)
3.0±0.3
0.5
(2.2 -- 3.5)
3.4±0.1
0.6
(2.5 -- 3.5)
3.5±0.0
0.4
(2.7 -- 3.5)
2.9±0.4
(2.0 -- 3.5)
2.5±0.9
1.0
(1.5 -- 4.9)
2.5±0.0
1.0
(1.5 -- 4.5)
2.5±0.7
1.0
(1.5 -- 4.8)
2.5±0.0
1.0
(1.0 -- 4.1)
2.5±0.1
1.0
(1.5 -- 4.7)
2.5±0.1
1.1
(1.5 -- 4.6)
2.9±1.3
1.0
(1.5 -- 5.2)
4.4±1.4
0.4
(3.5 -- 6.0)
2.5±0.0
0.9
(1.0 -- 4.0)
A = 4.5
A = 3.1
A = 3.4
A = 3.5
A = 2.8
A = 2.8
A > 1.9 (2.5?)
zspec = 2.98
are not taken into account in the fitting of the radio spec-
trum. We note however that excluding the 3.4cm upper-limit
from the modelling of the multi-wavelength data provides a
redshift estimate z = 2.8±1.3
0.3 (le2) that is in good agreement
with the true redshift of SMMJ02399−0136: z = 2.80. Other
redshift estimates are 1.65±0.80
0.53 based on the 1.4GHz/850µm
spectral index technique; and 2.83 ± 0.38 based on a χ2 min-
imization with one template SED (YC02).
SMMJ09429+4658 (H5): Other redshift estimates
are ≥ 3.6 based on the 1.4GHz/850µm spectral index tech-
nique; and 3.86 ± 0.49 based on a χ2 minimization with one
template SED (YC02). Our estimates actually place this
source at a slightly lower redshift z = 2.4±1.5
0.4 (le2). Given
the sub-mm detection, the low flux density at 1.4 GHz ex-
cludes almost all of the template SEDs in our library.
SMMJ14009+0252: Other redshift estimates are
1.27±1.59
0.54 based on the 1.4GHz/850µm spectral index tech-
nique; and 1.30±0.23 based a on a χ2 minimization with one
template SED (YC02). However large discrepancies (∼ 3σ)
exist between the template SED of YC02 and the observed
450µm and 1.4GHz fluxes, suggesting the fit will have a high
reduced-χ2, and that their errors (inferred for good matches)
might be under-estimated. The flattening of the spectrum
at 450µm, relative to the Raleigh-Jeans tail defined by the
850µm and 1.35mm fluxes, supports a higher redshift esti-
mate. Our own fit indicates that this object has a redshift
z ∼ 4.1 ± 0.8 (le2). The discrepancies between different red-
shift estimators might be attributed to the possible AGN
nature of this object (Ivison et al. 2000).
SMMJ14011+0252: Other redshift estimates are
2.53±1.24
0.82 based on the 1.4GHz/850µm spectral index tech-
nique; and 2.73 ± 0.37 based on a χ2 minimization with one
template SED (YC02). Our own estimate is z = 2.7±0.6
0.7
(le2) in good agreement with these estimates, and with the
spectroscopic redshift of the optical counterpart zspec =
2.57.
HR10: Other redshift estimates are 1.49±0.75
0.53 based on
the 1.4GHz/850µm spectral index technique; and 1.75±0.28
based on a χ2 minimization with one template SED (YC02).
Our own estimate z = 1.5±0.9
0.0 (le2) is in good agreement
with their measurements, and with the true spectroscopic
redshift of this source zspec = 1.44.
SMMJ02399−0134: Other redshift estimates are
1.17±0.53
0.45 based on the 1.4GHz/850µm spectral index tech-
nique; and 0.93 ± 0.19 based on a χ2 minimization with one
template SED (YC02). The spectroscopic redshift of this
source (zspec = 1.06) is inconsistent with our best estimate
of the redshift, which within a 68% confidence interval places
it at z = 2.3 ± 0.8 (le2). Redshifts in the range 1.0 ≤ z ≤ 1.5
have a 10% integrated probability.
SMMJ16403+46437: The detection at 1.4GHz
(593µJy) is high when compared with the 850µm level, and
there is no acceptable correspondence with any of the SEDs
in our template library. This, plausibly, might be a misiden-
tification or a radio-quiet AGN, with enhanced radio lumi-
nosity compared to the rest-frame FIR luminosity.
SMMJ16403+46440: Treating the 1.35mm observa-
tion as a detection (2.2σ) provides no additional constraint
and gives a similar redshift distribution to the one derived
from the 450µm/850µm colour and the 1.4GHz and 1.35mm
upper limits, shown in Figure A2.
3.3 Cumulative redshift distribution of the
sub-mm galaxy population
It is straightforward to calculate the cumulative redshift dis-
tribution for the sub-mm galaxy population as the coaddi-
tion of the individual probability distributions. It should be
noted that the calculation carried out in this paper has the
advantage of including the whole redshift probability dis-
tribution of each individual galaxy, and not just the mean
values of the distributions. We have included in this cal-
culation of the cumulative distribution only those sub-mm
sources identified in wide-area blank-field surveys (UK 8mJy
survey and the CUDSS) at a ≥ 3.5σ level: 50 sources. We
do not include sub-mm galaxies identified in surveys car-
ried out towards lensing clusters, or targeted observations
c(cid:13) 2002 RAS, MNRAS 000, 1 -- ??
Breaking the "Redshift Deadlock" -- II
13
Figure 9. Combined redshift distribution for the population of
sub-mm galaxies in the 8mJy SCUBA survey with one or more
measured colours (the 14 galaxies contained in Tables 1 & 2). The
different lines represent different evolutionary models assumed for
the sub-mm galaxy population (section 2). Upper panel: discrete
bin probabilities. Lower panel: cumulative distribution.
of other populations of galaxies (radio, ISO, Lyman-break
galaxies or extremely red objects), or catalogs that have not
been fully published (MAMBO surveys), since they can dis-
tort and bias the characteristic redshift distribution of the
sub-mm galaxy population.
Half of sources in complete flux-limited sub-mm sam-
ples, such as the UK 8mJy and CUDSS surveys, have a
single 850µm detection with one, or more, additional upper-
limits at other wavelengths: 28/50 sources detected (> 3.5σ)
at 850µm belong to this category. Since the redshift distribu-
tions of these sources are the most dependent on the priors
used, the interpretation of the combined redshift distribu-
tions should be based on the range of results given by the
different evolutionary models.
Fig. 9 shows the cumulative distribution and the com-
bined redshift distribution of the 14 galaxies in the UK 8mJy
survey with redshift distributions derived from SEDs which
contain at least 2-band detections (objects in Tables 1 and
2). This figure indicates that 65 -- 90 per cent of the sub-
mm galaxies could have redshifts in the range 2 <∼ z <∼ 4.
The results are consistent with the expectation that sub-
mm galaxies detected in the SCUBA (and also MAMBO)
surveys represent a high-z population (Dunlop 2001, Blain
et al. 2002).
c(cid:13) 2002 RAS, MNRAS 000, 1 -- ??
Figure 10. Combined redshift distribution for the population of
all 50 sub-mm galaxies detected in wide-area surveys at 850µm
(> 3.5σ) and listed in Tables 1,2, and 3. The different lines repre-
sent different evolutionary models assumed for the sub-mm galaxy
population (see text). Beware that the extended tail towards high-
redshifts is due to sources which are just detected at 850µm. The
upper limits at other wavelengths do not usually help to constrain
their redshifts between 2 ≤ z ≤ 10, and hence their probability
distributions are very flat in those regimes. Upper panel: discrete
bin probabilities. Lower panel: cumulative distribution.
The sources in the CUDSS sample with one or more
measured colours (9 sources) appear to have a significantly
different lower redshift distribution, with 40 per cent of the
galaxies at z <∼ 2 and approximately 50 per cent of the
galaxies at 2 <∼ z <∼ 4. The high redshift tail of the CUDSS
sample is similar to that shown by the 8mJy survey with ∼10
per cent probability at z > 4. The two galaxies responsible
for the low-redshift difference are CUDSS3.1 and 3.9, both
with relatively high 450µm flux densities S450µm ≥ 77 mJy,
compared with the 30 -- 40 mJy detection levels for the UK
8mJy survey sources. Taking the data for the CUDSS sample
at face-value suggests that it includes an over-density of low-
z galaxies at z <∼ 2, compared to the galaxies in the UK 8mJy
survey, and vice-versa. This discrepancy, however, could also
be explained by a difference in the flux-calibration of the
450µm data in the two surveys.
Fig. 10 shows the cumulative redshift distributions of
the sub-mm galaxy population for the 50 galaxies in the
8mJy and CUDSS surveys detected at a ≥ 3.5σ level. This
extended sample includes the 27 sub-mm galaxies with a
14
Itziar Aretxaga et al.
Figure 11. Distribution of intrinsic fluxes at 1.4GHz correspond-
ing to a population of galaxies discovered at 850µm, at a level
≥ 8mJy. Depending on the evolutionary model adopted, 80% of
the sources would have a flux in excess of 2 to 5µJy, and hence
in order to conduct a parallel 1.4GHz imaging follow-up program
with 80% completeness, the 1σ depth of the images should be 1
to 2µJy.
single detection at 850µm and one, or more, upper limits
at other wavelengths (Table 3). The combined redshift dis-
tributions, for the different evolutionary models under con-
sideration, imply that the majority of the sub-mm galax-
ies, approximately 50 -- 90 per cent, lie between z = 2 − 4.
The remainder of the galaxies, ∼ 10 per cent, lie at z < 2.
Fig. 10 also shows that <∼ 50 per cent of the galaxies have
colours that are consistent with z > 4. The distribution of
the high-redshift tail is very sensitive to the choice of evo-
lutionary model, since the individual redshift distributions
for the 27 sources with only a single detection are neces-
sarily very flat over the redshift range z >∼ 2 (see Fig. A3).
For these galaxies, the constraint on their redshift distribu-
tions comes mainly from the flux detected at 850µm and the
ability to reproduce it with the evolution of the 60µm local
luminosity function. More extensive photometry, especially
including shorter sub-mm wavelengths, is crucial to better
constrain these flat distributions, as illustrated by Fig. 7 (see
also Fig.6 in paper I).
4 THE LIKELIHOOD OF MIS-IDENTIFYING
THE RADIO COUNTERPARTS OF HIGH-Z
SUB-MM GALAXIES
The strength of the correlation between the FIR and radio
luminosities in starburst galaxies (Helou et al. 1985, Con-
don 1992) has supported the expectation that all high-z
star forming sub-mm galaxies should also have a detectable
radio-counterpart, provided the radio observations are deep
enough.
Although originally encouraged by this assumption,
some of the deepest VLA searches for radio emission as-
sociated with SCUBA and MAMBO sources have met with
varying degrees of success.
The deep VLA surveys of the Lockman Hole and ELAIS
N2 fields (Ivison et al. 2002) have a 1σ noise limit ≥ 5−9µJy
at 1.4 GHz. These surveys have identified 11/16 and 8/14
sub-mm sources at 1.4GHz, respectively, with peak radio
fluxes above 4σ and integrated fluxes above 3σ, after reject-
ing 5 and 1 sub-mm sources from the parent samples, re-
spectively, because of their high sub-mm noise levels. From
these, there are 9/16 and 5/14 sources in the Lockman Hole
and ELAIS N2 surveys, respectively, with a final S/N>3
when one considers the additional errors due to calibration
and fitting a gaussian to the peak of the radio emission.
If one limits the analysis to sources ≥ 8 mJy at 850µm,
then 4/10 and 4/8 sources have been simultaneously de-
tected with S/N> 3 at 850µm and 1.4GHz. This is in clear
contrast to the claims of previous authors (Fox et al. 2002)
that bright sub-mm sources (> 8mJy at 850µm) should be
detected at flux densities of ∼ 100µJy at 1.4 GHz.
In order to understand the properties of the radio asso-
ciations, we have calculated the probability of detection at
1.4GHz for a mock 8mJy survey, following the same method-
ology as in paper I, but using the extended library of SEDs
(section 2) . The intrinsic 850µm fluxes of the galaxies are
convolved with 1σ ∼ 2.6mJy measurement errors, and 9 per
cent calibration uncertainty. Fig.11 shows the distribution of
intrinsic 1.4GHz fluxes for the detected sources (including
errors) above the 8mJy threshold at 850µm.
We conclude that future surveys at 1.4 GHz must reach
a 3σ sensitivity of 2 -- 5µJy if radio observations are to pro-
vide counterparts to 80% of the high-z sub-mm galaxies with
850µm flux densities ≥ 8 mJy, according to the redshift dis-
tributions assumed in the alternative evolutionary models.
Thus, current radio observations have insufficient sensitivity
to detect the bulk of the sub-mm population with adequate
S/N, and we must wait for the next generation of new and
upgraded facilities (eg. e-VLA, e-MERLIN, SKA).
In the case of the UK 8mJy survey, Fig. 10 shows that
40 -- 50% (Lockman Hole) and 50 -- 60% (ELAIS N2) of the
sources should have an intrinsic 1.4GHz flux lower than the
3σ threshold, in the absence of any radio-loud AGN emis-
sion. These levels agree well with the number of radio coun-
terparts to ≥ 8mJy sources (Scott et al. 2002, Ivison et al.
2002), bearing in mind the actual variation in sensitivity
across the SCUBA wide-area survey fields.
This potential
for misidentifying the correct radio
sources associated with blank-field sub-mm galaxies will in-
evitably lead to incorrect estimates of photometric redshifts
if they are derived in part from the 1.4 GHz fluxes. In the cal-
culation of the cumulative redshift distribution (section 3.3)
we have treated all of the published associations of 1.4 GHz
and 850µm sources as the correct radio counterparts to the
sub-mm galaxies.
Despite the rapid acceptance of the radio -- sub-mm flux
density ratio, α1.4
345 (CY99, CY00), as the preferred method
with which to estimate the redshift distribution of the sub-
mm galaxy population, this diagnostic ratio loses its power
to discriminate between photometric redshifts at z > 2
(CY99, CY00). We have already demonstrated in paper I
(Fig. 11) the detrimental effect of the large dispersion in
radio -- sub-mm colours of the local template galaxies on the
accuracy of the derived photometric redshifts of the sub-mm
blank field sources. Furthermore, the existing radio interfer-
ometers will remain limited in their ability to map the large
areas ∼ 1 -- 40 sq. deg. that will be surveyed by the next
c(cid:13) 2002 RAS, MNRAS 000, 1 -- ??
Breaking the "Redshift Deadlock" -- II
15
(Downes et al. 1999) and the unprecedented depth of
the multi-wavelength HDF surveys (Williams et al. 1996;
Thompson et al. 1999; Richards 1999; Hornschmeier et al.
2000). It is only after the accurate subtraction of the ellip-
tical galaxy, 3−586, at z ∼ 1, however, that a faint lensed
IR counterpart (K = 23.5, H − K = 1.5, I − K > 5.1) to
HDF850.1 has been discovered (Dunlop et al. 2002).
This IR detection now gives greater confidence to the
likelihood that the coincident radio emission, with a flux
density of 16±4µJy at 1.4 GHz, from the combined data of
MERLIN and the VLA, is genuinely associated with the
sub-mmm source (Dunlop et al. 2002). Furthermore this
strengthens the suggestion that a marginal 3.3σ detection
at 8.4 GHz (VLA 3651+1226, Richards, priv. comm.) that
may also associated with HDF850.1.
The redshift probability distribution for HDF850.1,
based on the multi-wavelength 170µm -- 1.4 GHz data, indi-
cates z is comprised between 3.3 and 5.4 with a 86% to 90%
confidence depending on the adopted prior (Fig.12). This
high photometric redshift supports the earlier decision to
reject the z ∼ 1 elliptical galaxy 3−586, as the optical coun-
terpart. Unfortunately the IR counterpart, HDF850.1K, is
too faint to attempt IR spectroscopy with even the 10-m
class of telescopes.
The most efficient means to determine the spectroscopic
redshifts of optically-obscured sub-mm galaxies is the de-
tection of millimetre and centimetre wavelength molecular
lines using large single-dish telescopes (e.g. GBT or LMT)
with broad-band receivers (Townsend et al. 2001, Carilli &
Blain 2002). In this way, it is possible to eliminate the neces-
sity for an optical/IR counterpart, or a position accurate to
< 3 arcsec. Given the predicted accuracy of the photomet-
ric redshift for HDF850.1, and taking all of the uncertainties
discussed in section 2 into account, we demonstrate in Fig.12
that the J=1 -- 0 CO line in HDF850.1 has a 86-90% probabil-
ity of detection in the K-band receiver (18.0 -- 26.5 GHz) on
the GBT. Observations to conduct searches for CO emission
in high-z sub-mm galaxies, guided only by their photometric
redshifts, are currently underway on the GBT.
6 CONCLUSIONS
We have derived meaningful photometric-redshift probabil-
ity distributions for individual sub-mm galaxies using Monte
Carlo simulations that take into account a range of measure-
ment errors and model-dependent uncertainties. From the
individual redshift probability distributions, we have mea-
sured the cumulative redshift distribution for the sub-mm
population of galaxies identified in blank field surveys. We
find that the multi-wavelength FIR -- radio data for 50 sub-
mm galaxies detected (> 3.5σ) at 850µm are consistent with
∼ 50−90% of the population lying at redshifts z = 2−4, with
only a small fraction, ∼ 10% of the population at z < 2. We
also show that up to 50% of the galaxies have colours that
are consistent with z > 4. The possibility that the sub-mm
galaxy population contains a significant fraction undergoing
a major burst of star formation at such early epochs has im-
portant consequences for models of galaxy formation. This
suggestion is still highly dependent on the prior assumption
of the evolutionary model for the sub-mm population which
Figure 12. Redshift probability distributions for HDF850.1. The
different line types and colours correspond to the different evo-
lutionary models considered. Model led1 is not plotted, since it
excludes sources at z > 3.5 due to the strong prior decline of
the density and luminosity of sources. The unshaded area corre-
sponds to the redshift range where the 12CO (J = 1 − 0) line lies
in the observable K-band window (18.5 -- 26.0 GHz), accessible
with the 100m GBT. There is an 86-90% probability of detecting
the J=1-0 CO-line in this GBT window
generation of sub-mm and FIR experiments (e.g. BLAST,
Herschel, SIRTF) down to the necessary sensitivity levels
(3σ ∼ 2 − 5µJy at 1.4GHz), making the use of α345
1.4 an im-
practical measure of redshift for the majority of the popula-
tion. Deep interferometric radio observations, however, re-
main a practical way to pinpoint the source of the sub-mm
emission with sufficient positional accuracy to investigate
galaxy morphologies and exploit optical/IR spectroscopic
line diagnostics.
This
the
limitation of
radio -- sub-mm technique
prompted the design and construction of a Balloon-borne
Large Aperture Submillimetre Telescope, BLAST (Devlin et
al. 2001). The primary scientific goal of BLAST is to break
the current "redshift deadlock" by providing photometric
redshifts with sufficient accuracy, without the prior neces-
sity for secure optical, IR or radio counterparts. BLAST will
use rest-frame sub-mm and FIR colours to derive redshifts
with an accuracy ∆z ∼ ±0.5 for > 1000 galaxies in the red-
shift range 0 < z < 6 detected in wide-area (> 0.5 −10 deg2)
surveys at 250, 350 and 500µm (Hughes et al. 2002).
This redshift accuracy is sufficient to guide the choice of
possible front-end spectrographs, and their receiver tunings,
on large single-dish ground-based mm and cm telescopes
(e.g. 100-m GBT -- Green Bank Telescope and 50-m LMT --
Large Millimetre Telescope) in the effort to detect redshifted
CO emission-lines. We conclude this paper with the descrip-
tion of one application of this technique.
5 BROAD-BAND MM -- CM SPECTROSCOPY
Until recently the brightest sub-mm source in the Hubble
Deep Field, HDF850.1 (Hughes et al. 1998), has avoided
the detection of its X-ray, optical, IR or radio counterpart,
despite an accurate interferometric position at 236 GHz
c(cid:13) 2002 RAS, MNRAS 000, 1 -- ??
16
Itziar Aretxaga et al.
affects most strongly those sub-mm sources detected only at
850µm.
Shorter-wavelength sub-mm data (250 − 500µm) from a
future balloon-borne experiment, BLAST, will provide pow-
erful additional constraints (∆z ∼ ±0.5) on the redshift dis-
tribution of all the SCUBA and MAMBO galaxies at z > 2.
A photometric-redshift accuracy of ±0.5 for an individual
galaxy will be sufficient to provide robust statistical mea-
surements of the redshift distribution of the entire popula-
tion, the global history of dust-obscured star formation and
the clustering properties of sub-mm galaxies.
We have argued that the large dispersion in the radio
to sub-mm SEDs (or 850µm/1.4GHz colour) of the template
galaxies naturally translates into an imprecise measure of
the photometric redshift, and that radio observations can-
not be used efficiently to measure the individual redshifts
of thousands of galaxies detected in future wide-area sub-
mm surveys. The sub-mm/radio spectral index is still able
to discriminate between the z < 2 and z > 2 regime.
We have an on-going programme to conduct centimetre-
wavelength observations of CO J=1-0 molecular line tran-
sitions in high-z sub-mm galaxies on the 100-m GBT. The
observations, however, first require an accurate photometric-
redshift probability distribution to guide the choice of re-
ceiver and tuning parameters. We illustrate the power of
our photometric-redshift technique with the example of
HDF850.1, one of the best-studied sub-mm galaxies (Hughes
et al. 1998; Downes et al. 1999; Dunlop et al. 2002).
Within the next few years, the combination of ground-
based, balloon-borne and satellite experiments will provide
accurately calibrated rest-frame sub-mm -- FIR data for a sig-
nificant population of sub-mm galaxies. The centimetre and
millimetre wavelength spectroscopic redshift confirmation of
their photometric redshifts will calibrate the method out-
lined in this paper. Thus we can finally look forward to
breaking the "redshift deadlock" that currently prevents an
accurate understanding of the nature of the sub-mm galaxy
population and their evolutionary history.
7 ACKNOWLEDGMENTS
This work has been partly supported by CONACYT grants
32180-E and 32143-E. We thank the anonymous referee for
the critical comments and careful reading of this manuscript.
REFERENCES
Barger A.J., Cowie, L.L, Richards, E.A., 2000, A.J., 119, 2092
Bertoldi F. et al., 2000, A&A, 360, 92.
Blain A.W. et al., 2002, Physics Reports,
in press (astro-
ph/020228)
Borys, C, Chapman S.C, Halpern M., Scott D., 2002, MNRAS,
330, 63.
Carilli C.L. & Blain , 2002, ApJ, 569 605.
Carilli C.L. & Yun M.S., 1999, Ap.J., 513, L13 (CY99)
Carilli C.L. & Yun M.S., 2000, Ap.J., 530, 618 (CY00)
Chapin E. et al, 2003, MNRAS, in preparation.
Chapman S.C. et al. 2000, MNRAS, 319, 318.
Chapman S.C., Scott D., Borys C., & Fahlman G.G., 2002, MN-
RAS, 330, 92.
Chapman S.C., Smail I., Ivison R.J., Helou G., Dale D.A., La-
gache G., 2002b, ApJ, in press (astro-ph/0203068)
Condon J.J., 1992, ARAA, 30, 575
Dannerbauer H. et al., 2002, ApJ, in press (astro-ph/0201104)
Devlin M. et al., 2001, in J. Lowenthal & D.H. Hughes eds., Deep
Millimetre Surveys: Implications for Galaxy Formation and
Evolution World Scientific, p. 59 (astro-ph/0012327)
Dey A., Graham J.R., Ivison R.J., Smail I., Wright G.S. & Liu
M.C., 1999, ApJ, 519, 610.
Downes D., et al., 1999, A&A, 347, 809
Dunlop J.S., 2001, in FIRSED2000, eds. I.M. van Bremmel et al.,
Elsevier New Astronomy Reviews (astro-ph/0101297).
Dunlop J.S. et al., MNRAS, 2002, submitted
Dunne L., Eales S.A., Edmunds M., Ivison R., Alexander P., &
Clements D.L., 2000, MNRAS, 315, 115
Cowie L., Barger A.J. & Kneib J., 2002, AJ, 123, 2197.
Eales S.A., et al., 1999, Ap.J., 515, 518
Eales S.A., et al., 2000, A.J., 120, 2244
Fardal M.A., Katz N., Weinberg D.H., Dav´e R., Hernquist L.,
2002, ApJ, submitted (astro-ph/0107290)
Fox M.J., et al., 2002, MNRAS, 331, 839.
Frayer D.T., et al., 2000, A.J., 120, 1668
Gear W.K., Lilly S. J., Stevens J.A., Clements D.L., Webb T.M.,
Eales S.A., Dunne, L., 2000, MNRAS, 316, 51.
Helou P. et al. 1985, ApJ, 440, 35
Hughes D.H., Robson E.I., Dunlop J.S., Gear W.K., 1993, MN-
RAS, 263, 607
Hughes D.H., et al., 1998, Nature, 394, 241
Hughes D.H., et al., 2002, MNRAS, submitted, (paper I, astro-
ph/0111547)
Hornschmeier A., et al., 2000, Ap.J., 541, 49
Ivison R.I. et al., 1998, MNRAS, 298, 583
Ivison R.J., et al., 2000, MNRAS, 315, 209
Ivison R.J., et al., 2002, MNRAS, in press (astro-ph/0206432)
Klaas U. et al. 2001. A&A, 379, 823
Lilly S.J., et al., 1999, Ap.J., 518, 441
Lutz D., et al., 2001, A&A, 378, 70.
Richards E.A., et al., 1998, Ap.J., 116, 1039
Richards E.A., 1999, Ap.J., 513, L9
Richards E.A., 2000, Ap.J., 533, 611
Saunders D.B. Rowan-Robinson M., Lawrence A., Efstathiou G.,
Kaiser N., Ellis R. S., Frenk C. S., 1990, MNRAS, 242, 318
Scott D. et al., 2000, A&A, 357, 5.
Scott S., et al., 2002, MNRAS, 331, 817.
Smail I., et al., 1997, MNRAS, 490, L5
Smail I., et al., 1999, MNRAS, 308, 1061
Smail I., Ivison R.J., Blain A.W., Kneib J.-P., 2002, MNRAS,
331, 495.
Thompson R., et al., 1999, A.J., 117, 17
Townsend et al. 2001, MNRAS, 328, L17
Webb T.M.A., et al., 2002, ApJ, submitted (astro-ph/0201181).
Williams R.E., et al., 1996, A.J., 112, 1335
Yun M.S. & Carilli C.L., 2002, Ap.J., in press (YC02, astro-
ph/0112074).
APPENDIX A: REDSHIFT ESTIMATES FOR
INDIVIDUAL SUB-MM GALAXIES
A catalogue of redshift probability distributions and SEDs
for the galaxies detected in more than 2 bands (Tables 1 and
2) is included. We also include the corresponding colour-
colour-redshift or colour-flux-redshift plots from which they
have been derived. All estimates are derived for model le2.
For galaxies detected in just one band (Table 3) we
present a selection of the results derived for 6 out of the
46 galaxies included in Table 3. The general uniformity of
their redshift distributions is a fair representation of those
galaxies not included in this appendix.
c(cid:13) 2002 RAS, MNRAS 000, 1 -- ??
|
0710.3826 | 1 | 0710 | 2007-10-20T09:03:32 | R fluids | [
"astro-ph"
] | A theory of collisionless fluids is developed in a unified picture, where nonrotating figures with anisotropic random velocity component distributions and rotating figures with isotropic random velocity component distributions, make adjoints configurations to the same system. R fluids are defined and mean and rms angular velocities and mean and rms tangential velocity components are expressed, by weighting on the moment of inertia and the mass, respectively. The definition of figure rotation is extended to R fluids. The generalized tensor virial equations are formulated for R fluids and further attention is devoted to axisymmetric configurations where, for selected coordinate axes, a variation in figure rotation has to be counterbalanced by a variation in anisotropy excess and vice versa. A microscopical analysis of systematic and random motions is performed under a few general hypotheses, by reversing the sign of tangential or axial velocity components of an assigned fraction of particles, leaving the distribution function and other parameters unchanged (Meza 2002). The application of the reversion process to tangential velocity components, implies the conversion of random motion rotation kinetic energy into systematic motion rotation kinetic energy. The application of the reversion process to axial velocity components, implies the conversion of random motion translation kinetic energy into systematic motion translation kinetic energy, and the loss related to a change of reference frame is expressed in terms of systematic (imaginary) motion rotation kinetic energy. A procedure is sketched for deriving the spin parameter distribution (including imaginary rotation) from a sample of observed or simulated large-scale collisionless fluids i.e. galaxies and galaxy clusters. | astro-ph | astro-ph |
R fluids
R. Caimmi∗
December 19, 2018
Abstract
A theory of collisionless fluids is developed in a unified picture,
(σ11 = σ22 = σ33) random velocity component distributions and ro-
where nonrotating (fΩ1 = fΩ2 = fΩ3 = 0) figures with anisotropic
tating (fΩ1 6= fΩ2 6= fΩ3) figures with isotropic (σ11 6= σ22 6= σ33) ran-
dom velocity component distributions, make adjoints configurations
to the same system. R fluids are defined as ideal, self-gravitating
fluids satisfying the virial theorem assumptions, in presence of sys-
tematic rotation around each principal axis of inertia. To this aim,
mean and rms angular velocities and mean and rms tangential veloc-
ity components are expressed, by weighting on the moment of inertia
and the mass, respectively. The figure rotation is defined as the mean
angular velocity, weighted on the moment of inertia, with respect to
a selected axis. The generalized tensor virial equations (Caimmi &
Marmo 2005) are formulated for R fluids and further attention is de-
voted to axisymmetric configurations where, for selected coordinate
axes, a variation in figure rotation has to be counterbalanced by a
variation in anisotropy excess and vice versa. A microscopical analy-
sis of systematic and random motions is performed under a few general
∗Dipartimento di Astronomia, Universit`a di Padova, Vicolo Osservatorio 2, I-35122
Padova, Italy - email: [email protected]
1
hypotheses, by reversing the sign of tangential or axial velocity com-
ponents of an assigned fraction of particles, leaving the distribution
function and other parameters unchanged (Meza 2002). The appli-
cation of the reversion process to tangential velocity components, is
found to imply the conversion of random motion rotation kinetic en-
ergy into systematic motion rotation kinetic energy. The application
of the reversion process to axial velocity components, is found to im-
ply the conversion of random motion translation kinetic energy into
systematic motion translation kinetic energy, and the loss related to
a change of reference frame is expressed in terms of systematic (imag-
inary) motion rotation kinetic energy. A number of special situations
are investigated with further detail. It is found that a R fluid always
admits an adjoint configuration where figure rotation occurs around
only one principal axis of inertia (R3 fluid), which implies that all
the results related to R3 fluids (Caimmi 2007) may be extended to
R fluids. Finally, a procedure is sketched for deriving the spin pa-
rameter distribution (including imaginary rotation) from a sample of
observed or simulated large-scale collisionless fluids i.e. galaxies and
galaxy clusters.
keywords - galaxies: clusters - galaxies: haloes - stars: stellar systems.
1
Introduction
11 = σ2
33), where σ2
Due to particle shocks, collisional fluids (e.g., stars, gas clouds) exhibit an
isotropic stress tensor (σ2
pp is the rms random ve-
locity component on the axis, xp. The absence of particle shocks (leaving
aside extreme situations, such as high-density galactic nuclei) makes a dif-
ferent situation in collisionless fluids (e.g., galaxies, galaxy clusters), where
the stress tensor is - in general - anisotropic (σ2
33). The shape
of the body is determined by systematic rotation, which is quantified by a
spin parameter (null for nonrotating configurations), and/or by the difference
between stress tensor diagonal components, or any equivalent anisotropy in-
dicator, related to the rotation and an equatorial principal axis of inertia,
respectively (null for configurations where the random velocity component
distribution is isotropic). A description of collisionless fluids based on the
equivalence of systematic and random motions with respect to the shape,
appears to be highly rewarding and it would provide further insight on the
11 6= σ2
22 6= σ2
22 = σ2
2
properties of stellar and galaxy systems.
In an earlier attempt (Caimmi 1996) the stress tensor has been expressed
as the sum of two terms, one related to a random (isotropic) velocity com-
ponent distribution, and one other to anisotropic internal motions within
the system. Further investigation has been devoted to the simplest situation
where the system is made of two equal components, which are rotating at
the same rate but in opposite sense. Then it has been recognized that the
anisotropy excess may be related to real rotation, if the shape is flattened,
and to imaginary rotation, if the shape is elongated, with respect to the
rotation axis.
A latter approach (Caimmi & Marmo 2005) has been restricted to homeoidally
striated density profiles, for which the tensor virial equations were formulated
and generalized to unrelaxed configurations. The kinetic-energy tensor has
been expressed as the sum of two terms, one related to systematic rotation
obeying an assigned law, and one other to the remaining motions e.g., ran-
dom motions, streaming motions, radial motions. Finally, an expression of
the spin parameter in terms of the anisotropy excess, has shown the role of
systematic and remaining motions in flattening or elongating the shape.
The above mentioned results have been improved and extended in sub-
sequent work (Caimmi 2006, hereafter quoted as C06), where imaginary ro-
tation has been related to negative anisotropy excess. Then sequences of
configurations for which the generalized tensor virial equations hold, have
been determined for homeoidally striated Jacobi ellipsoids including prolate
shapes induced by imaginary rotation. The results from numerical simula-
tions on the stability of rapidly rotating spherical configurations (Meza 2002)
have been interpreted in the light of the theory. To this respect, the key argu-
ment is that the reversion (from clockwise to counterclockwise or vice versa)
of tangential velocity components related to an assigned fraction of partices,
preserves the potential energy, the kinetic energy, and the distribution func-
tion (Lynden-Bell 1960, 1962; Meza 2002).
The study on homeoidally striated Jacobi ellipsoids has been extended to
a more general class of bodies (R3 fluids) in a recent paper (Caimmi 2007,
hereafter quoted as C07), where the contribution of radial and tangential
velocity components on the equatorial plane was investigated with further
detail. In addition, mean and rms (weighted on the moment of inertia) angu-
lar velocity have been defined, and related to systematic and random motion
tangential kinetic-energy tensor components, respectively. Also for R3 fluids,
3
it has been realized that the effect of (positive or negative) anisotropy excess
is equivalent to additional (real or imaginary) figure rotation.
The current attempt is aimed to extend the above mentioned results to a
still more general class of bodies, R fluids, defined as ideal, self-gravitating,
collisionless fluids where rotation occurs around each principal axis of iner-
tia.
It will be found that R fluids always admit an adjoint configuration
where figure rotation occurs around a single principal axis, that is a R3 fluid.
Accordingly, all the results which hold for R3 fluids may be extended to R
fluids.
The work is organized as follows. A number of basic definitions are pro-
vided in Sect. 2, including the inertia tensor, the angular-velocity tensor, and
the angular-momentum tensor. The generalized tensor virial equations for
R fluids are formulated in Sect. 3. The microscopical analysis of systematic
and random motions , for a collisionless fluid made of N identical parti-
cles, is performed in Sect. 4, where a velocity component reversion process
is defined, and a number of special situations are analysed in detail with
respect to kinetic energy changes from random to systematic motions and
vice versa. A procedure aimed to the derivation of the spin parameter distri-
bution (including imaginary rotation) from an assigned sample of observed
or simulated objects, is outlined in Sect. 5. Some concluding remarks are
reported in Sect. 6.
2 Angular-velocity and angular-momentum ten-
sor
In the special case of solid bodies, rotation is rigid and occurs around a single
axis which, in turn, can remain fixed or change its direction. Accordingly,
the angular momentum and the rotation kinetic energy read (e.g., Landau &
Lifchits 1966, Chap. VI, §§ 31-33; hereafter quoted as LL66):
Jr =
Trot =
3Xs=1
1
2
I ′
rsΩs
;
r = 1, 2, 3 ;
3Xr=1
3Xs=1
I ′
rsΩrΩs
;
4
(1)
(2)
where ~J = (J1, J2, J3) is the angular-momentum vector,
the angular-velocity vector, and I ′ the inertia tensor:
−→
Ω = (Ω1, Ω2, Ω3)
I ′
rs =ZS
ρ(x1, x2, x3)"δrs
3Xr=1
r − xrxs# d3S ;
x2
(3)
related to the density profile, ρ, within the volume, S, being δrs the Kronecker
33, are the
symbol. The diagonal components of the inertia tensor, I ′
moments of inertia with respect to the axes, x1, x2, x3, respectively.
22, I ′
11, I ′
In addition, the inertia tensor is symmetric of second rank, which implies
the existence of a reference frame, (O′ X1 X2 X3), where the inertia tensor is
diagonal (LL66, Chap. VI, § 32):
I ′
rs = δrsI ′
rr
;
(4)
the coordinate axes coincide with the principal axes of inertia, and the di-
agonal components define the principal moments of inertia. Accordingly,
Eqs. (1) and (2) reduce to:
3Xs=1
2(cid:16)I ′
1
Jr =
I ′
rrΩr
;
r = 1, 2, 3 ;
Trot =
11Ω2
1 + I ′
22Ω2
2 + I ′
33Ω2
3(cid:17) ;
where, in addition (LL66, Chap. VI, § 32):
rr ≤ I ′
I ′
ss + I ′
tt
;
r = 1, 2, 3 ;
s = 2, 3, 1 ;
t = 3, 1, 2 ;
(7)
with regard to the body under consideration.
The inertia tensor has been defined in a different way, as (e.g., Chan-
drasekhar 1969, Chap. 2, § 9; Binney & Tremaine 1987, Chap. 4, § 3):
Irs =ZS
ρ(x1, x2, x3)xrxs d3S ;
and the combination of Eqs. (3) and (8) yields:
I ′
rs = δrs
Irr − Irs
;
3Xr=1
I ′
rr = Iss + Itt
I ′
rs = −Irs
;
;
r 6= s 6= t ;
r 6= s ;
5
(5)
(6)
(8)
(9)
(10)
(11)
or:
ss + I ′
tt − I ′
rr
;
r 6= s 6= t ;
2Irr = I ′
Irs = −I ′
rs
;
r 6= s ;
(12)
(13)
which translates one formulation into the other (e.g., Bett et al. 2007).
In the general case of (collisional or collisionless) fluids, rotation could
occur different from solid-body, and around each principal axis of inertia.
Let (O x1 x2 x3) be a generic reference frame and (O′ X1 X2 X3) a reference
frame where the origin coincides with the centre of inertia, and the coordinate
axes coincide with the principal axes of inertia. Let the coordinate axes, X1,
−→
X2, X3, be defined as the principal axes. Let
Ω3, be the angular-
velocity vectors (to be specified later) related to the principal axes of inertia.
−→
Ωr, on the coordinate axis, xs. The
Let Ωrs be the component of the vector,
(3 × 3) tensor, Ωrs, is defined as the angular-velocity tensor of the system
under consideration, with respect to the reference frame, (O x1 x2 x3). Let
−→ω1, −→ω2, −→ω3, be the angular-velocity vectors related to the coordinate axes, x1,
x2, x3. The following relation holds:
−→
Ω1,
−→
Ω2,
ωs =
3Xr=1
Ωrs
;
s = 1, 2, 3 ;
(14)
and the angular-velocity tensor, ωrs = Ωrs, can formally be defined. Simi-
larly, the (3 × 3) angular-momentum tensor is expressed as:
J ′
rs = I ′
rsωrs
;
J ′
s =
3Xr=1
I ′
rsωrs
;
(15)
(16)
where the inertia tensor is related to the reference frame, (O x1 x2 x3).
In the special case where the reference frame, (O x1 x2 x3), coincides with
(O′ X1 X2 X3), then Ωrs = δrsΩrr. Accordingly, Eqs. (14)-(16) reduce to:
ωs = Ωs = Ωss = ωss
J ′
rs = δrsI ′
J ′
s = Js = I ′
rrΩrr
ssΩs
;
;
;
(17)
(18)
(19)
where I ′
tt = Irr + Iss, r 6= s 6= t, represents the moment of inertia with respect
to the principal axis of inertia, xt. From this point on, it shall be intended
6
that the origin coincides with the centre of inertia, and the coordinate axes
coincide with the principal axes of inertia.
The rotation kinetic-energy tensor is defined as:
(Trot)rs =
1
2
I ′
rsΩrΩs =
1
2
δrsI ′
rrΩ2
r
;
(20)
where the diagonal components of the angular-velocity tensor are expressed
as (C07):
1
I ′
rr ZS(cid:12)(cid:12)(cid:12)
Ωr = fΩr =
Ωr(x1, x2, x3, t)(cid:12)(cid:12)(cid:12) =
(cid:12)(cid:12)(cid:12)
wr = (x2
s + x2
t )1/2 ;
−→
−→
Ωr(x1, x2, x3, t)(cid:12)(cid:12)(cid:12) w2
vφr(x1, x2, x3, t)
;
wr
r ρ(x1, x2, x3, t) d3S; r 6= s 6= t;
(21)
(22)
(23)
and Ωr(x1, x2, x3, t) is the mean value related to all the particles at the time, t,
within the infinitesimal volume element, d3S = dx1 dx2 dx3, centred on the
point, P(x1, x2, x3), vφr is the tangential velocity component on the (O xs xt)
principal plane, and the moment of inertia, I ′
rr = Iss + Itt, reads:
I ′
rr =ZS
w2
r ρ(x1, x2, x3, t) d3S ;
r 6= s 6= t ;
(24)
as expected from the theorem of the mean, in connection with Eq. (21).
Similarly, the mean square diagonal components of the angular-velocity
tensor are expressed as:
w2
r ρ(x1, x2, x3, t) d3S; r 6= s 6= t; (25)
(26)
r) =
Ω2
r = g(Ω2
1
I ′
rr ZSh−→
and the related variance reads:
Ωr(x1, x2, x3, t)i2
(cid:16)σfΩrfΩr(cid:17)2
= g(Ω2
r) − (fΩr)2 ;
as known from statistics.
At this stage, it may be useful to extend and generalize the definition of
figure rotation.
Figure rotation. Given a R fluid, the figure rotation is defined as the mean
angular velocity, weighted on the moment of inertia, with respect to a selected
principal axis.
7
In terms of tangential velocity components, the counterparts of Eqs. (21),
(25), and (26) read:
1
1
−→
φr ) =
M ZS(cid:12)(cid:12)(cid:12)
Ωr(x1, x2, x3, t)(cid:12)(cid:12)(cid:12) wrρ(x1, x2, x3, t) d3S ;
gvφr =
M ZSh−→
Ωr(x1, x2, x3, t)i2
g(v2
(cid:16)σfvφrfvφr(cid:17)2
= g(v2
φr ) − (gvφr )2 ;
w2
r ρ(x1, x2, x3, t) d3S; r 6= s 6= t;
r 6= s 6= t ;
(27)
(28)
(29)
and the combination of Eqs. (25) and (28); (21) and (27); (26) and (29);
yields:
r) ;
φr ) = I ′
M g(v2
rr g(Ω2
rr(fΩr)2 ;
M(gvφr )2 = I ′
M(cid:16)σeφreφr(cid:17)2
rr(cid:16)σfΩrfΩr(cid:17)2
= I ′
(30a)
(30b)
(30c)
;
which relate tangential velocity components on the (O xs xt) principal plane,
to angular velocity components on the xr principal axis.
3 The generalized tensor virial equations for
R fluids
Let R fluids be defined as (collisional or collisionless) ideal self-gravitating
fluids where figure rotation occurs around all the three principal axes of
inertia. Let (O x1 x2 x3) be a reference frame where the origin coincides with
the centre of inertia, and the coordinate axes coincide with the principal axes
of inertia. Then the mean radial velocity components must necessarily equal
zero:
vwr = 0 ;
(v2
wr ) = (σwrwr )2
;
(31)
where vwr is the radial velocity component on the (O xs xt) principal plane,
perpendicular to the xr principal axis. Let positive and negative radial ve-
locity components be defined as directed outwards and inwards, respectively.
The same holds for the mean tangential velocity components:
vφr = 0 ;
(v2
φr ) = (σφrφr )2
;
(32)
8
even in presence of systematic rotation. Let positive and negative tangential
velocity components be defined as rotating counterclockwise and clockwise,
respectively.
The kinetic-energy tensor may be expressed as the sum of two contribu-
tions: one, related to systematic motions, and one other, related to random
motions (C07). The result is:
Tkskt = (Tsys)kskt + (Trdm)kskt
;
k = w, φ ;
(33)
where the terms on the right-hand side, using Eqs. (30) and (31), can be
expressed as:
(Tsys)wswt = 0 ;
1
2
(Trdm)wswt =
δstM (σwsws)2
(Tsys)φsφt =
(Trdm)φsφt =
1
2
1
2
δstI ′
δstI ′
ss(fΩs)2 ;
ss(cid:16)σfΩsfΩs(cid:17)2
;
;
(34a)
(34b)
(35a)
(35b)
keeping in mind that nondiagonal components are null in the case under
discussion, only diagonal components shall be considered from this point on.
The combination of Eqs. (26) and (35) yields:
Tφrφr =
1
2
r) ;
I ′
rr g(Ω2
(36)
which depends on the density profile via the moment of inertia, I ′
rr, and
the tangential velocity component distribution via the mean square angular
r), regardless from the fraction of systematic and random motions.
In terms of the contributions related to the axial components of the
velocity, g(Ω2
kinetic-energy tensor, Tss and Ttt, Eqs. (33), (35), and (36) read:
(Tφrφr )ℓℓ =
1
2
[(Tsys)φrφr ]ℓℓ =
[(Trdm)φrφr ]ℓℓ =
2
r) ;
Iℓℓ g(Ω2
Iℓℓ g(Ωr)
Iℓℓh g(Ω2
1
2
1
2
r) − (fΩr)2i
9
ℓ = s, t ;
;
ℓ = s, t ;
(37a)
(37b)
;
ℓ = s, t ;
(37c)
where Eq. (10) has been used.
The invariance of a vector with respect to a change of the reference frame,
implies the validity of the relations (C07):
s ) + (v2
wr) + (v2
φr) = (v2
(v2
(vwr)2 + (vφr )2 = (vs)2 + (vt)2 ;
(σwrwr)2 + (σφr φr)2 = (σss)2 + (σtt)2
t ) ;
(38)
(39)
(40)
;
where the velocity components on the xs and xt principal axes are labelled
by the indices, s and t, respectively.
The combination of Eqs. (26), (29), and (38)-(40) yields:
(σwrwr )2 = (σss)2 + (σtt)2 −
which makes Eqs. (33) and (34) translate into:
(Twrwr )ℓℓ = [(Trdm)wrwr]ℓℓ =
1
2
Mσ2
ℓℓ −
1
2
[(Tsys)wrwr ]ℓℓ = 0 ;
ℓ = s, t ;
;
I ′
rr
(41)
M h g(Ω2
Iℓℓh g(Ω2
r) − (fΩr)2i
r) − (fΩr)2i ; ℓ = s, t ; (42a)
(42b)
in terms of the contributions related to the axial components of the kinetic-
energy tensor, Tss and Ttt.
The generalized tensor virial equations of the second order can be formu-
lated, extending the procedure used for R3 fluids (C07). The result is:
Irrh(fΩs)2 + (fΩt)2i + Mζrrσ2 + (Epot)rr = 0 ;
11 + σ2
22 + σ2
σ2 = σ2
33 ;
ζpp =
(eTrdm)pp
Trdm
=
σ2
pp
σ2
;
p = 1, 2, 3 ;
ζ11 + ζ22 + ζ33 = eTrdm
Trdm
= eσ2
σ2 = ζ ;
(43)
(44)
(45a)
(45b)
where mean angular velocity components on the xr principal axis are due
to systematic rotation around xs and xt axes, (Epot)rr is the self potential-
energy tensor, ζrr may be conceived as generalized anisotropy parameters
10
(Caimmi & Marmo 2005; C06; C07), and eTrdm is the effective random kinetic
energy i.e. the right amount needed for an instantaneous configuration to
satisfy the usual tensor virial equations of the second order, defined by the
effective anisotropy parameters (C07):
=
ζpp
ζ
(eTrdm)pp
eTrdm
eζpp =
eζ11 + eζ22 + eζ33 = 1 ;
;
p = 1, 2, 3 ;
(46a)
(46b)
and the condition, ζ = 1, or ζpp = eζpp, p = 1, 2, 3, makes Eqs. (43) reduce to
their standard counterparts. To get further insight, a microscopical analysis
is needed.
In the special case of axisymmetric configurations, I11 = I22, (Epot)11 =
(Epot)22, and the combination of the related tensor virial equations, expressed
by Eq. (43), yields:
p = 1, 2 ;
q = 2, 1 ;
(47)
Ipph(fΩq)2 − (fΩp)2i = Mσ2(ζqq − ζpp) ;
[(fΩq)2 − (fΩp)2], is counterbalanced by an
where a figure rotation excess,
anisotropy excess, (ζqq − ζpp);
in particular, a null figure rotation excess
implies a null anisotropy excess and vice versa. Accordingly, a flattening
on the (O xp xr) principal plane, induced by the figure rotation excess, has
to be counterbalanced by an elongation on the xq principal axis, induced by
the anisotropy excess, to yield an axisymmetric configuration with respect
to the xr principal axis.
4 Microscopical analysis of systematic and ran-
dom motions
Given a collisionless R fluid, let N be the total number of particles and
m the mean particle mass in absence of mass segregation i.e.
local and
global mean particle mass coincide. For simplicity, the equivalent description
(C07) involving N identical particles of mass, m, shall be considered. With
regard to the (O xs xt) principal plane, let vφr be the tangential velocity
component on the above mentioned plane. It is worth noting (Meza 2002)
that the distribution function is independent of the sign of vφr , and the
11
whole set of possible configurations are characterized by an equal amount of
both kinetic and potential energy. Numerical simulations show that spherical
systems, even if rapidly rotating, are dynamically stable after reversion of
the tangential velocity component in an assigned fraction of particles (Meza
2002).
For sake of simplicity, let the initial configuration be nonrotating (vφr = 0)
and with isotropic random velocity component distribution (ζ11 = ζ22 = ζ33).
In the case under discussion of identical particles, m(i) = m, 1 ≤ i ≤ N,
the centre of inertia velocity components, vCr, equal the related arithmetic
means:
vCr =
= vr
;
(48)
m
=
v(i)
r
NXi=1
Nm
m(i)v(i)
r
m(i)
NXi=1
NXi=1
and the moments of inertia, I ′
rr, reduce to:
I ′
rr =
NXi=1
r i2
m(i)hw(i)
= m
r i2
NXi=1hw(i)
= MR2
Gr
;
(49)
where RGr = [w2
r ]1/2 is the curl radius with respect to the xr axis.
The weighted mean, mean square, and rms tangential velocity compo-
nents, expressed by Eqs. (27)-(29), read:
m(i)v(i)
φr =
v(i)
φr = vφr
;
m
M
NXi=1
φri2
m(i)hv(i)
=
m
M
φri2
NXi=1hv(i)
= (v2
φr) − (vφr )2 = (σφrφr )2
NXi=1
NXi=1
1
M
1
M
φr ) =
gvφr =
g(v2
(cid:16)σeφreφr(cid:17)2
and Eqs. (30) reduce to:
= (v2
φr ) ;
(50)
(51)
(52)
(53a)
(53b)
(53c)
;
;
φr ) = R2
(v2
(vφr )2 = R2
(σφrφr )2 = R2
r) ;
Gr g(Ω2
Gr(fΩr)2 ;
Gr(cid:16)σfΩrfΩr(cid:17)2
12
which relate weighted angular velocities around the xr axis to mean tangential
velocity components on the (O xs xt) principal plane.
At this stage, let the tangential velocity component of a fraction, n/N,
of particles, be reversed in equal sense (from clockwise to counterclockwise
or vice versa), according to the following assumptions.
(i) Both the number, n, of particles where the tangential velocity component
has been reversed, and the number, N − n, of particles which remain
unchanged, are sufficiently large, 1 ≪ n ≪ N, 0 ≤ n ≤ Int(N/2).
(ii) The fraction, nk/Nk, of particles where the tangential velocity compo-
nent has been reversed, within a generic volume element, Sk, is inde-
pendent of the volume element, nk/Nk = n/N.
(iii) The system is made of identical particles, m(i) = m, M = mN.
(iv) After tangential velocity components have been reversed in nk particles
within a generic volume element, Sk, on a total of Nk, a second set of
nk particles (among the remaining Nk −nk) exists, where the tangential
velocity component of any particle equals its counterpart belonging to
the first set.
Obviously, mean square tangential velocity components, (v2
In the following, the above process shall be quoted as "the reversion process".
φr ), are left un-
changed by the reversion process. On the contrary, mean tangential velocity
components after the reversion process read:
vφr =
1
N
NXi=1
v(i)
φr =
(54)
1
N
2nXi=1
v(i)
φr +
NXi=2n+1
v(i)
φr ;
where the first sum within brackets relates to particles where the reversion
process has occurred and their counterparts with equal tangential velocity
components, while the second sum comprises the remaining particles and
necessarily equals the mean tangential velocity component before the occur-
rence of the reversion process, which is null in the case under discussion.
Accordingly, Eq. (54) reduces to:
vφr =
2n
N
(vφr )n =
(vφr )n ;
1
2n
2nXi=1
v(i)
φr =
1
n
nXi=1
v(i)
φr
;
13
(55a)
(55b)
keeping in mind that the first sum is performed on couples of particles with
equal tangential velocity components.
The validity of Eqs. (54) and (55) still maintains if tangential velocity
components, vφr , are replaced by axial velocity components, vr. The combi-
nation of Eqs. (48) and (55a) yields:
vCr = vr =
2n
N
(vr)n ;
(56)
which is the velocity component of the centre of inertia with respect to the
xr principal axis, after the reversion process.
The total kinetic energy is left unchanged by the reversion process but,
on the other hand, a fraction of random motion kinetic energy is turned into
systematic motion kinetic energy. In the following, the reversion process shall
be discussed with further details for a number of different situations.
4.1 Tangential velocity component reversion
Performing the reversion process on a given fraction of particles, n/N, with
respect to tangential velocity components, implies the conversion of random
(rotation) motion kinetic energy into systematic (rotation) motion kinetic
energy, as:
∆(Trdm)φrφr = −∆(Tsys)φrφr = −
1
2
M(vφr )2 = −
2n
N
nm[(vφr )n]2 ;
(57)
where the remaining parameters are left unchanged.
The occurrence of the reversion process implies the following energy
changes:
Trdm → Trdm −
2n
N
nm[(vφr )n]2 ;
(Trdm)φrφr → (Trdm)φrφr −
nm[(vφr )n]2 ;
2n
N
nm[(vφr )n]2 ;
ℓ = s, t ;
1
2
2n
N
(Trdm)ℓℓ → (Trdm)ℓℓ −
Tsys → 0 +
2n
N
nm[(vφr )n]2 ;
(Tsys)φrφr → 0 +
2n
N
nm[(vφr )n]2 ;
14
(58)
(59)
(60)
(61)
(62)
(Tsys)ℓℓ → 0 +
1
2
2n
N
nm(vφr )2
n ;
ℓ = s, t ;
(63)
while the contributions from random radial motions along the equatorial
plane, (Trdm)wrwr , and the rotation axis, (Trdm)rr, remain unchanged.
With the system being relaxed, ζ = 1, in the case under discussion, the
generalized and effective anisotropy parameters, ζpp and eζpp, coincide with
their counterparts related to the usual tensor virial equations, and Eqs. (45a)
and (46a) take the explicit form (C06):
ζℓℓ =
ζrr =
(1/3)Trdm − (2n/N)(n/2)m[(vφr )n]2
Trdm − (2n/N)nm[(vφr )n]2
(1/3)Trdm
Trdm − (2n/N)nm[(vφr )n]2
;
;
ℓ = s, t ;
(64a)
(64b)
where the special case, n = 0, relates to the initial configuration, character-
ized by isotropic random velocity component distributions (ζpp = 1/3) and
no figure rotation.
In the extreme case where the reversion process is completed, n = N/2,
the changes expressed by Eqs. (58)-(63) take the form:
Trdm → Trdm −
N
2
m[(vφr )N/2]2 ;
(Trdm)φrφr → (Trdm)φrφr −
N
2
m[(vφr )N/2]2 ;
(Trdm)ℓℓ → (Trdm)ℓℓ −
N
4
m[(vφr )N/2]2 ;
ℓ = s, t ;
Tsys → 0 +
N
2
m[(vφr )N/2]2 ;
(Tsys)φrφr → 0 +
(Tsys)ℓℓ → 0 +
N
4
m[(vφr )N/2]2 ;
N
2
m[(vφr )N/2]2 ;
ℓ = s, t ;
(65)
(66)
(67)
(68)
(69)
(70)
similarly, Eqs. (64) take the form:
ζℓℓ =
ζrr =
(1/3)Trdm − (N/4)m[(vφr )N/2]2
Trdm − (N/2)m[(vφr )N/2]2
(1/3)Trdm
Trdm − (N/2)m[(vφr )N/2]2
;
;
ℓ = s, t ;
(71a)
(71b)
15
in any case, the anisotropy excess, ζℓℓ − ζrr < 0, is counterbalanced by figure
rotation.
4.2 Axial velocity component reversion
Performing the reversion process on a given fraction of particles, n/N, with
respect to axial velocity components, implies the conversion of random (trans-
lation) motion kinetic energy into systematic (translation) motion kinetic
energy, as:
∆(Trdm)rr = −∆(Tsys)rr = −
1
2
M(vr)2 = −
2n
N
nm[(vr)n]2 ;
(72)
where the remaining parameters are left unchanged.
The reversion process implies the following energy changes:
(Trdm)rr → (Trdm)rr −
2n
N
nm[(vr)n]2 ;
Trdm → Trdm −
2n
N
nm[(vr)n]2 ;
Tsys → 0 +
2n
N
(Tsys)rr → 0 +
nm[(vr)n]2 ;
2n
N
nm[(vr)n]2 ;
(73)
(74)
(75)
(76)
while the contributions from random motions along the xs and xt principal
axis, (Trdm)ss and (Trdm)tt, remain unchanged.
With the system being relaxed, ζ = 1, in the case under discussion, the
generalized and effective anisotropy parameters, ζpp and eζpp, coincide with
their counterparts related to the usual tensor virial equations, and Eqs. (45a)
and (46a) take the explicit form:
ζℓℓ =
ζrr =
(1/3)Trdm
;
Trdm − (2n/N)nm[(vr)n]2
(1/3)Trdm − (2n/N)nm[(vr)n]2
Trdm − (2n/N)nm[(vr)n]2
ℓ = s, t ;
(77a)
;
(77b)
where the special case, n = 0, relates to the initial configuration, character-
ized by isotropic random velocity component distributions (ζpp = 1/3) and
no figure rotation.
16
In the extreme case where the reversion process is completed, n = N/2,
the changes expressed by Eqs. (73)-(76) take the form:
Trdm → Trdm −
N
2
m[(vr)N/2]2 ;
(Trdm)rr → (Trdm)rr −
N
2
m[(vr)N/2]2 ;
Tsys → 0 +
N
2
m[(vr)N/2]2 ;
(Tsys)rr → 0 +
N
2
m[(vr)N/2]2 ;
(78)
(79)
(80)
(81)
similarly, Eqs. (77) take the form:
ζℓℓ =
ζrr =
Trdm − (N/2)m[(vr)N/2]2
(1/3)Trdm − (N/2)m[(vr)N/2]2
Trdm − (N/2)m[(vr)N/2]2
(1/3)Trdm
;
ℓ = s, t ;
(82a)
;
(82b)
in any case, the anisotropy excess, ζℓℓ − ζrr > 0, is counterbalanced by (imag-
inary) figure rotation.
4.3 Change of reference frame and imaginary rotation
Performing the reversion process on a given fraction of particles, n/N, implies
the conversion of random motion (translation) kinetic energy into systematic
motion (translation) kinetic energy, with respect to the xr principal axis,
according to Eqs. (72)-(76). The related kinetic-energy tensor component of
the centre of inertia, by use of Eqs. (56) and (72) reads:
(TC)rr =
1
2
M(vr)2 =
2n
N
nm[(vr)n]2 = −∆(Trdm)rr
;
(83)
which, in the case under discussion, coincides with the kinetic energy of
the centre of inertia, TC, being (TC)ss = (TC)tt = 0. In the centre of inertia
reference frame after the reversion process, the random motion kinetic-energy
tensor component related to the xr axis, (T ′
rdm)rr, by use of Eqs. (56) and (83),
takes the expression:
(T ′
rdm)rr =
1
2
m
NXi=1hv(i)
r − vCri2
= (Trdm)rr −
1
2
M(vr)2 = (Trdm)rr − (TC)rr
;
(84)
17
where the kinetic energy, TC = (TC)rr, is hidden by the change of reference
frame (e.g., LL66, Chap. II, § 8).
Let the i-th particle be at the distance, w(i)
]2}1/2, from
the xr principal axis, with velocity component, v(i)
r . The imaginary angu-
lar velocity (Caimmi 1996; C06; C07), iΩ(i)
r , can be defined in such a way
the translational kinetic energy along the xr axis is counterbalanced by the
imaginary rotational kinetic energy around the xr axis, as:
r = {[x(i)
s ]2 + [x(i)
t
1
2
m[v(i)
r ]2 +
v(i)
r
w(i)
r
Ω(i)
r =
1
2
mhw(i)
r i2
r i2hiΩ(i)
= 0 ;
;
(85)
(86)
where the index, i, is related to the i-th particle, and the factor, i, is the
imaginary unit. In this view, the velocity components on the xr principal
axis may be translated into imaginary tangential velocity components on the
(O xs xt) principal plane, as:
(ivφr )2 = (iwrΩr)2 = −v2
r
;
(87)
according to Eqs. (85) and (86).
Let imaginary rotation around the xr axis be imparted to the particles
where the reversion process has been occurred, and their counterparts with
equal imaginary tangential velocity components, as prescribed by Eq. (87)
particularized to the mean axial velocity component, vr, expressed by Eq. (56).
The related increment in imaginary kinetic energy reads:
∆(Tsys)φrφr =
1
2
M(ivφr )2 =
2n
N
nm[(ivφr )n]2 ;
(88)
and the combination of Eqs. (72), (83), (87), and (88) yields:
∆(Tsys)φrφr = −∆(Tsys)rr = −(TC)rr = ∆(Trdm)rr
;
(89)
the above results may be reduced to a single statement.
Theorem 1. Given a R fluid with isotropic random velocity distribution
and no figure rotation, let the axial velocity component reversion process
be performed on a given fraction of particles, n/N, with respect to the xr
18
principal axis. Then turning to the centre of inertia reference frame with a
kinetic energy loss, ∆(Tsys)rr = (TC)rr, is equivalent to put the initial con-
figuration into imaginary rotation around the xr principal axis, with square
mean tangential velocity component, (ivφr )2 = −(vr)2, with a kinetic energy
gain, ∆(Tsys)φrφr = −(TC)rr.
Accordingly, Eqs. (73)-(76) are replaced by the following:
Tsys → 0 −
2n
N
nm[(vφr )n]2 ;
(Tsys)φrφr → 0 −
(Tsys)ℓℓ → 0 −
1
2
2n
N
2n
N
nm[(vφr )n]2 ;
nm[(vφr )n]2 ;
(90)
(91)
(92)
while the contributions from random motions remain unchanged, and the
random velocity distribution remains isotropic (ζ11 = ζ22 = ζ33 = 1/3).
In the extreme case where the reversion process is complete, n = N/2,
and the maximum amount of available imaginary rotation has been attained,
the changes expressed by Eq. (90)-(92) take the form:
Tsys → 0 −
N
2
m[(vφr )N/2]2 ;
(Tsys)φrφr → 0 −
(Tsys)ℓℓ → 0 −
N
4
m[(vφr )N/2]2 ;
N
2
m[(vφr )N/2]2 ;
(93)
(94)
(95)
the above results may be reduced to a single statement.
Theorem 2. Given a R fluid with isotropic random velocity distribution
and no figure rotation, let the axial velocity component reversion process be
performed on a given fraction of particles, n/N, with respect to the xr princi-
pal axis, and the reference frame changed into the centre of inertia reference
frame. Then the resulting configuration with anisotropy excess, ζℓℓ − ζrr > 0,
Eqs. (77), is equivalent to the initial configuration with null anisotropy ex-
cess, ζℓℓ − ζrr = 0, and imaginary rotation around the xr principal axis, with
square mean tangential velocity component, (ivφr )2 = −(vr)2.
19
4.4 Anisotropy excess and imaginary rotation
Given a nonrotating (fΩr = 0) isotropic (ζ11 = ζ22 = ζ33 = 1/3) configuration,
let the tangential velocity component on the (O xs xt) principal plane be
reversed in such a way Eqs. (65)-(71) hold. Let an equal amount of real and
imaginary tangential velocity component on the (O xs xt) principal plane, be
imparted to each particle for a total contribution equal to (2/3)frTrdm and
−(2/3)frTrdm, respectively, where fr is a positive real number, which leaves
the total energy unchanged. Let the reversion process be repeated for real
tangential velocity components, to attain a null (real) figure rotation. The
related changes, with respect to the initial configuration, are:
2
3
Trdm →(cid:18)1 +
(Trdm)φrφr →(cid:18)2
2(cid:18)2
(Trdm)ℓℓ →
3
3
fr(cid:19) Trdm ;
+
+
2
3
2
3
fr(cid:19) Trdm ;
fr(cid:19) Trdm ;
1
2
3
Tsys → 0 −
frTrdm ;
(Tsys)φrφr → 0 −
(Tsys)ℓℓ → 0 −
1
2
2
3
2
3
frTrdm ;
frTrdm ;
(96)
(97)
(98)
(99)
(100)
(101)
and the related anisotropy parameters read:
ζℓℓ =
ζrr =
[(1/3) + (1/3)fr]Trdm
[1 + (2/3)fr]Trdm
=
(1/3)Trdm
[1 + (2/3)fr]Trdm
=
3 + 2fr
1 + fr
3 + 2fr
1
;
;
ℓ = s, t ;
(102a)
(102b)
where the anisotropy excess, ζℓℓ − ζrr = fr/(3 + 2fr) > 0, is counterbalanced
by imaginary rotation, according to an initial configuration with isotropic
velocity component distribution and no figure rotation. As application of the
above results, two significative examples shall be taken into consideration.
First example. Nonrotating systems flattened by anisotropic velocity com-
ponent distribution (σ11 = σ22 > σ33).
20
Let tangential velocity components on the (O x1 x2) principal plane be re-
versed in a convenient fraction of particles, n/N, to yield a convenient amount
of figure rotation together with isotropic velocity component distribution
(σ′
22 = σ33), as sketched in Fig. 1.
11 = σ′
The combination of Eqs. (52), (53), and (55a) yields:
1
R2
Gr
=
4n2
N 2 [(vφr )n]2 ;
1
(σφrφr )2 =
R2
Gr
Gr ((v2
1
R2
φr ) −
4n2
N 2 [(vφr )n]2) ;
(103)
(104)
(fΩr)2 =
(cid:16)σfΩrfΩr(cid:17)2
while the mean square velocity components are left unchanged by the oc-
currence of the reversion process. The substitution of Eqs. (103) and (104)
into (37b) and (37c) shows the dependence of systematic and random motion
tangential kinetic-energy tensor components on the reversion process.
In the case under discussion (r = 3), the random velocity component
distribution has to be isotropic after the reversion process, which makes
Eqs. (60) and (63) reduce to:
(Trdm)ℓℓ =
(Tsys)ℓℓ =
M
2
M
2 hσ2
2 (cid:16)σ2
M
ℓℓ − σ2
33(cid:17)i =
ℓℓ −(cid:16)σ2
33(cid:17) ;
ℓℓ − σ2
4n2
N 2 M[(vφr )n]2 =
1
2
ℓ = 1, 2 ;
2n
N
nm[(vφr )n]2 =
σ2
33 ;
ℓ = 1, 2 ;
(105)
(106)
M
2
(vφr )2 =
M
2 (cid:16)σ2
ℓℓ − σ2
33(cid:17) ; (107)
which defines the configuration after the reversion process.
Second example. Systems elongated by anisotropic velocity component
distribution (σ11 = σ22 < σ33) with no figure rotation.
Let a convenient amount of real and imaginary figure rotation, vφ3 and ivφ3,
with respect to the x3 principal axis, be imparted to the system. Then the
kinetic energy remains unchanged. Concerning real rotation, let the reversion
process be performed on one half particles, leaving imaginary figure rotation
together with isotropic velocity component distribution (σ′
22 = σ33), as
sketched in Fig. 2. Accordingly, Eqs. (103) and (104) hold for imaginary tan-
gential velocity components on the (O x1 x2) principal plane and imaginary
figure rotation around the x3 principal axis.
11 = σ′
21
x3
✻
x3
✻
✵
O
r
x1
✠
✲
x2
O
r
✲
x2
x1
✠
σ11 = σ22 > σ33 ;
(vφ3 )2 = 0 ;
σ
′
11 = σ
′
22 = σ33 ;
(vφ3 )2 = σ
2
11 − σ
2
33 = σ
2
22 − σ
2
33 ;
Figure 1: After reversion of tangential velocity components in a convenient
fraction of particles, with respect to the (Ox1x2) principal plane, a configu-
ration flattened by anisotropic velocity component distribution (σ11 = σ22 >
σ33) with no figure rotation (vφ3 = 0), left picture, is turned into a configu-
ration flattened by figure rotation [vφ3 = (σ2
33)1/2] with
isotropic velocity component distribution (σ′
22 = σ33), right picture.
The symbol, ✵, denotes figure rotation around the related principal axis.
33)1/2 = (σ2
11 = σ′
11 − σ2
11 − σ2
22
x3
✻
x3
✻
✵
O
r
x1
✠
✲
x2
O
r
✲
x2
x1
✠
σ11 = σ22 < σ33 ;
(ivφ3 )2 = 0 ;
σ
′
11 = σ
′
22 = σ33 ;
(ivφ3 )2 = σ
2
11 − σ
2
33 = σ
2
22 − σ
2
33 ;
Figure 2: After imparting a convenient amount of real and imaginary fig-
ure rotation, vφ3 and ivφ3, with respect to the x3 principal axis, to the
system, and reversing real tangential velocity components on one half par-
ticles, a configuration elongated by anisotropic velocity component distri-
bution (σ11 = σ22 < σ33) with no figure rotation (vφ3 = 0),
left pic-
ture, is turned into a configuration elongated by (imaginary) figure rotation
[ivφ3 = (σ2
33)1/2] with isotropic velocity component
distribution (σ′
22 = σ33), right picture. The symbol, ✵, denotes figure
rotation around the related principal axis.
33)1/2 = (σ2
11 − σ2
11 − σ2
11 = σ′
23
In the case under discussion (r = 3), an isotropic velocity component
distribution after the reversion process implies the validity of Eqs. (105),
(106), and (107), where the tangential velocity components are imaginary
(σℓℓ < σ33), and the configuration after the reversion process is completely
defined.
4.5 Tangential velocity component reversion in the gen-
eral case
In the general case of anisotropic velocity component distribution (σ11 6=
0), with regard to the (O xs xt) principal plane, let the tangential velocity
component reversion process be applied to one half particles in such a way
σ22 6= σ33) and figure rotation around each principal axis (fΩ1 6= fΩ2 6= fΩ3 6=
no figure rotation around the xr principal axis occurs, fΩr = 0. Keeping in
mind Eqs. (50)-(53), the related energy changes read:
(Trdm)φrφr → (Trdm)φrφr +
(Tsys)φrφr → (Tsys)φrφr −
1
2
M(vφr )2 ;
1
2
M(vφr )2 = 0 ;
vφr = (vφr )N/2 =
1
N
NXi=1
v(i)
φr
;
∆(Tsys)φrφr = −∆(Trdm)φrφr = −
M(vφr )2 = −
MR2
(108)
(109)
(110)
(113)
1
2
1
Mσ2
ℓℓ
2
1
2
Gr(fΩr)2 ; (111)
(112)
;
ℓ = 1, 2 ;
[(Trdm)φrφr ]ℓℓ + [(Trdm)wrwr ]ℓℓ =
(Trdm)rr =
1
2
Mσ2
rr
;
where the rms velocity components, σ2
configuration.
ℓℓ and σ2
rr, are related to the initial
The changes in anisotropy paramenters, ζpp = σ2
pp/σ2, read:
ζℓℓ →
ζrr →
ζℓℓ + (1/2)∆ζℓℓ
1 + ∆ζℓℓ
ζrr
;
1 + ∆ζℓℓ
24
;
ℓ = s, t ;
(114a)
(114b)
∆ζℓℓ =
(vφr )2
σ2
;
ℓ = s, t ;
(114c)
where the rms velocity, σ2, is related to the initial configuration.
11, σ2
22, σ2
The application of the above procedure to the x1 and x2 principal axes,
makes the transition from an initial configuration with rms velocity compo-
nents, σ2
to a final configuration with rms velocity components, (σ′
33, and figure rotation around the principal axes, fΩ1, fΩ2, fΩ3,
and figure rotation around the principal axes, 0, 0, fΩ3. The related energy
changes read:
11)2, (σ′
22)2, (σ′
33)2,
M(vφ2)2 ;
M(vφ1)2 ;
Mh(vφ1)2 + (vφ2)2i
(Trdm)11 → (Trdm)11 +
(Trdm)22 → (Trdm)22 +
(Trdm)33 → (Trdm)33 +
1
4
1
4
1
4
Trdm → Trdm +
(Trdm)pp =
1
2
1
2
Mσ2
Mh(vφ1)2 + (vφ2)2i
;
pp ;
p = 1, 2, 3 ;
;
(115)
(116)
(117)
(118)
(119)
(120a)
(120b)
(120c)
(120d)
and the related changes in anisotropy parameters are:
ζ11 →
ζ22 →
ζ33 →
;
ζ11 + (1/2)∆ζ11
1 + ∆ζ11 + ∆ζ22
ζ22 + (1/2)∆ζ22
1 + ∆ζ11 + ∆ζ22
ζ33 + (1/2)[∆ζ11 + ∆ζ22]
;
1 + ∆ζ11 + ∆ζ22
;
∆ζ11 =
(vφ2)2
σ2
;
∆ζ22 =
(vφ1)2
σ2
;
;
the above results may be restricted to a single statement.
Theorem 3. Given a R fluid, a convenient application of the tangential
velocity component reversion process makes an adjoint configuration where
figure rotation occurs around a single principal axis, that is a R3 fluid.
Accordingly, the results valid for R3 fluids (C07) may be extended to the
general case of R fluids.
25
5 Discussion
As suggested in earlier attempts (Caimmi 1996; C06; C07), the equivalence
between a variation in figure rotation and in anisotropy excess, may provide
a useful tool for the description of collisionless fluids. The discussion here
shall be focused on the spin parameter (Peebles 1969, 1971):
λ2 = −
J 2E
G2M 5
;
(121)
where G is the gravitation constant, M the total mass, J the total angular
momentum, and E the total energy. The above formulation includes four
possibilities, namely (i) real rotation (J 2 ≥ 0) and bound system (E < 0),
which is the sole currently used in literature; (ii) imaginary rotation (J 2 < 0)
and bound system (E < 0); (iii)real rotation (J 2 ≥ 0) and unbound system
(E ≥ 0); (iv) imaginary rotation (J 2 < 0) and unbound system (E ≥ 0).
Accordingly, the spin parameter attains real values in cases (i) and (iv), and
imaginary values in cases (ii) and (iii).
In the light of the current model, oblate-like and prolate-like configura-
tions belong to cases (i) and (ii) outlined above, while cases (iii) and (iv) rep-
resent unbound structures for which the virial equations do not hold. Then
the comparison with observations and/or computations, must be restricted
to bound configurations.
The spin parameter distribution is usually fitted using a lognormal dis-
tribution (e.g., van den Bosh 1998; Gardner 2001; Ballin & Steinmetz 2005;
Hernandez et al. 2007) or, in general, dependent on log λ (e.g., Bett et al.
2007), in dealing with real rotation. The inclusion of imaginary rotation
would imply use of λ2 instead of λ as independent variable, allowing for both
positive (real rotation) and negative (imaginary rotation) values.
The following procedure should be followed for calculating λ2 from obser-
vations and/or computations: (1) determine the inertia tensor and the prin-
cipal axes of inertia for an assigned matter distribution; (2) determine the
potential-energy tensor; (3) determine the anisotropy parameters using the
generalized virial equations; (4) perform the reversion process with respect
to two principal axes of inertia to leave figure rotation around the third one
(R3 fluid); (5) convert the anisotropy excess into (real or imaginary) figure
rotation to obtain isotropic velocity component distribution (ζ11 = ζ22 = ζ33);
(6) evaluate the spin parameter; (7) act as already done for all the sample
26
objects; (8) determine the distribution of the spin parameter, P (λ2), with re-
spect to the sample of adjoints configurations, where the velocity component
distribution is isotropic.
The related results could provide further insight on the formation and
the evolution of large-scale collisionless fluids, such as galaxies and galaxy
clusters.
6 Conclusion
A theory of collisionless fluids has been developed in a unified picture, where
nonrotating (fΩ1 = fΩ2 = fΩ3 = 0) figures with isotropic (σ11 = σ22 = σ33) ran-
dom velocity component distributions and rotating (fΩ1 6= fΩ2 6= fΩ3) figures
with anisotropic (σ11 6= σ22 6= σ33) random velocity component distribu-
tions, make adjoints configurations to the same system. R fluids have been
defined as ideal, self-gravitating fluids satisfying the virial theorem assump-
tions (e.g., LL66, Chap. II, § 10; C07), in presence of figure rotation around
each principal axis of inertia.
To this aim, mean and rms angular velocities and mean and rms tangential
velocity components have been expressed, by weighting on the moment of
inertia and the mass, respectively. The figure rotation has been defined as
the mean angular velocity, weighted on the moment of inertia, with respect
to a selected axis.
The generalized tensor virial equations (Caimmi & Marmo 2005) have
been formulated for R fluids and further attention has been devoted to ax-
isymmetric configurations where, for selected coordinate axes, a variation in
figure rotation has to be counterbalanced by a variation in anisotropy excess
and vice versa.
A microscopical analysis of systematic and random motions has been
performed under a few general hypotheses, by reversing the sign of tangential
or axial velocity components of an assigned fraction of particles, leaving the
distribution function and other parameters unchanged (Meza 2002).
The application of the reversion process to tangential velocity compo-
nents, has been found to imply the conversion of random motion rotation
kinetic energy into systematic motion rotation kinetic energy. The applica-
tion of the reversion process to axial velocity components, has been found to
imply the conversion of random motion translation kinetic energy into sys-
27
tematic motion translation kinetic energy, and the loss related to a change of
reference frame has been expressed in terms of systematic motion (imaginary)
rotation kinetic energy.
A number of special situations have been investigated with further detail.
It has been found that a R fluid always admits an adjoint configuration
where figure rotation occurs around only one principal axis of inertia (R3
fluid), which implies that all the results related to R3 fluids (Caimmi 2007)
may be extended to R fluids.
Finally, a procedure has been sketched for deriving the spin parameter
distribution (including imaginary rotation) from a sample of observed or
simulated large-scale collisionless fluids i.e. galaxies and galaxy clusters.
References
[1] Ballin, J., Steinmetz, M.: 2005, ApJ, 627, 647.
[2] Bett, P., Eke, V., Frenk, C.S., et al.: 2007, MNRAS, 376, 215.
[3] Binney, J., & Tremaine, S.: 1987, Galactic Dynamics, Princeton Univer-
sity Press, Princeton. (BT87)
[4] Caimmi, R.: 1996, AN, 317, 401.
[5] Caimmi, R.: 2006, AN, 327, 925. (C06)
[6] Caimmi, R. 2007, SerAJ, 174, 13. (C07)
[7] Caimmi, R., Marmo, C.: 2005, AN, 326, 465.
[8] Chandrasekhar, S.: 1969, Ellipsoidal Figures of Equilibrium, Yale Uni-
versity Press, New Haven.
[9] Gardner, J.P.: 2001, ApJ, 557, 616.
[10] Hernandez, X., Park, C., Cervantes-Sodi, B., Choi, Y.: 2007, MNRAS,
375, 163.
[11] Landau, L., Lifchitz, E.: 1966, Mecanique, Mir, Moscow. (LL66)
[12] Lynden-Bell, D.: 1960, MNRAS, 120, 204.
28
[13] Lynden-Bell, D.: 1962, MNRAS, 124, 1.
[14] Meza, A.: 2002, A&A, 395, 25.
[15] Peebles, P.J.E.: 1969, ApJ, 155, 393.
[16] Peebles, P.J.E.: 1971, A&A, 11, 377.
[17] van den Bosh, F.C.: 1998, ApJ, 507, 601.
29
|
astro-ph/9905127 | 1 | 9905 | 1999-05-11T18:16:35 | Python I, II, and III CMB Anisotropy Measurement Constraints on Open and Flat-Lambda CDM Cosmogonies | [
"astro-ph"
] | We use Python I, II, and III cosmic microwave background anisotropy data to constrain cosmogonies. We account for the Python beamwidth and calibration uncertainties. We consider open and spatially-flat-Lambda cold dark matter cosmogonies, with nonrelativistic-mass density parameter Omega_0 in the range 0.1--1, baryonic-mass density parameter Omega_B in the range (0.005--0.029) h^{-2}, and age of the universe t_0 in the range (10--20) Gyr. Marginalizing over all parameters but Omega_0, the combined Python data favors an open (spatially-flat-Lambda) model with Omega_0 simeq 0.2 (0.1). At the 2 sigma confidence level model normalizations deduced from the combined Python data are mostly consistent with those drawn from the DMR, UCSB South Pole 1994, ARGO, MAX 4 and 5, White Dish, and SuZIE data sets. | astro-ph | astro-ph |
KSUPT-99/1, KUNS-1552
February 1999
Python I, II, and III CMB Anisotropy Measurement Constraints on Open and
Flat-Λ CDM Cosmogonies
Gra¸ca Rocha1,2, Rados law Stompor3,4, Ken Ganga5, Bharat Ratra1, Stephen R. Platt6,
Naoshi Sugiyama7, and Krzysztof M. G´orski8,9
ABSTRACT
We use Python I, II, and III cosmic microwave background anisotropy data
to constrain cosmogonies. We account for the Python beamwidth and calibration
uncertainties. We consider open and spatially-flat-Λ cold dark matter cosmogonies,
with nonrelativistic-mass density parameter Ω0 in the range 0.1 -- 1, baryonic-mass
density parameter ΩB in the range (0.005 -- 0.029)h−2 , and age of the universe t0 in
the range (10 -- 20) Gyr. Marginalizing over all parameters but Ω0, the combined
Python data favors an open (spatially-flat-Λ) model with Ω0 ≃ 0.2 (0.1). At the 2 σ
confidence level model normalizations deduced from the combined Python data are
mostly consistent with those drawn from the DMR, UCSB South Pole 1994, ARGO,
MAX 4 and 5, White Dish, and SuZIE data sets.
Subject headings: cosmic microwave background -- cosmology: observations -- large-scale
structure of the universe
1.
Introduction
Ganga et al. (1997a, hereafter GRGS) developed a technique to account for uncertainties,
such as those in the beamwidth and the calibration, in likelihood analyses of cosmic microwave
1Department of Physics, Kansas State University, Manhattan, KS 66506.
2Centro de Astrof´ısica da Universidade do Porto, Rua das Estrelas s/n, 4100 Porto, Portugal.
3Center for Particle Astrophysics, University of California, Berkeley, CA 94720.
4Copernicus Astronomical Center, Bartycka 18, 00-716 Warszawa, Poland.
5Infrared Processing and Analysis Center, MS 100 -- 22, California Institute of Technology, Pasadena, CA 91125.
6Snow and Ice Research Group, University of Nebraska, Lincoln, NE 68583-0850.
7Department of Physics, Kyoto University, Kitashirakawa-Oiwakecho, Sakyo-ku, Kyoto 606-8502, Japan.
8Theoretical Astrophysics Center, Juliane Maries Vej 30, 2100 Copenhagen Ø, Denmark.
9Warsaw University Observatory, Aleje Ujazdowskie 4, 00-478 Warszawa, Poland.
-- 2 --
background (CMB) anisotropy data. This technique has been used in conjunction with
theoretically-predicted CMB anisotropy spectra in analyses of the Gundersen et al. (1995) UCSB
South Pole 1994 data, the Church et al. (1997) SuZIE data, the MAX 4+5 data (Tanaka et al.
1996; Lim et al. 1996), the Tucker et al. (1993) White Dish data, and the de Bernardis et al.
(1994) ARGO data (GRGS; Ganga et al. 1997b, 1998; Ratra et al. 1998, 1999a, hereafter R99a).
A combined analysis of all these data sets is presented in Ratra et al. (1999b, hereafter R99b).
In this paper we present a similar analysis of CMB anisotropy data from the Python I, II,
and III observations performed at the South Pole (Dragovan et al. 1994, hereafter D94; Ruhl et
al. 1995b, hereafter R95; Platt et al. 1997, hereafter P97). The Python detectors and telescope
are described by Ruhl (1993) and D94; also see Ruhl et al. (1995a) and Alvarez (1996). In what
follows we review the information needed for our analysis.
Python I, II, and III CMB data were taken in a frequency band centered at 90 GHz with four
bolometric detectors centered at the corners of a 2.◦75 by 2.◦75 square on the sky. The beam profiles
are well-approximated by a Gaussian of FWHM 0.◦75 ± 0.◦05 (one standard deviation uncertainty).
Observations were centered at α = 23.h37, δ = −49.◦44 (J2000.0). Python I and II data were taken
at a single telescope elevation. Python III data were taken at this fiducial elevation as well as two
additional elevations offset 2.◦75/3 on the sky above and below the fiducial elevation. The reduced
Python data are shown in Figure 1.
All of the Python measurements were made by switching the four beams horizontally across
the sky in a three-point pattern by rotating a vertical flat mirror at 2.5 Hz. This chopping pattern
was then combined with slow (typically 0.1 Hz) azimuthal beam switching of the entire telescope
to produce a four-beam response to a sky signal.
The chopper throw and azimuthal telescope beam switching were both 2.◦75 on the sky for
the Python I and II observations. Python I (hereafter I) resulted in 16 data points, 8 each from
the lower and upper rows of detectors (D94). Python II observed two sets of points on the sky
(R95), although one of the detectors did not work during this season. The first set of observations,
hereafter IIA, overlapped the I points and yielded 7 measurements from the lower row of detectors
and 8 from the upper row. The second set, hereafter IIB, measured the same number of points on
the sky as the first set, but these were offset in azimuth relative to the I points by −2.◦75/2 on the
sky.
Two series of measurements were made at each of the three elevations observed during the
Python III season (P97). For each series, the physical throw of the chopper was the same at all
elevations. Because the actual throw on the sky depends on elevation, the Python III beam was
smeared to a Gaussian FWHM of 0.◦82 ± 0.◦05 (one standard deviation uncertainty) when forming
the four-beam pattern to account for the imperfect overlap of the beams caused by this effect.
This procedure also accounted for the relative pointing uncertainty and fluctuations in the chopper
throw.
The first series of Python III measurements, hereafter IIIL, used the same chopper and beam
-- 3 --
switch parameters as Python I and II. The IIIL data consists of two sets of points taken at the
I elevation but offset in right ascension by −2.◦75/3 and −2 × 2.◦75/3 on the sky. Two sets of
measurements were also made at each of the lower and upper elevations, but these were offset in
right ascension by 0.◦0 and +2 × 2.◦75/3 on the sky relative to I.
The second series of Python III measurements, hereafter IIIS, were made with both the
chopper throw and telescope beam switch reduced to 2.◦75/3. For each of the three telescope
elevations, the IIIS data consists of points separated horizontally by 2.◦75/3 on the sky for each of
the 2 rows of detectors. Figure 1 shows the points observed by IIIL and IIIS at each elevation.
Together, I, II, and III densely sample a 5.◦5 by 22◦ region of the sky.
The 1 σ absolute calibration uncertainty in the Python data is 20% (D94; R95; P97) and is
accounted for in our analysis. The absolute pointing uncertainty is 0.◦1 (D94; R95) and is not
accounted for in our analysis.
In §2 we summarize the computational techniques used in our analysis. See GRGS and R99a
for detailed discussions. Results are presented and discussed in §3. Conclusions are given in §4.
2. Summary of Computation
The zero-lag window function for the Python observations are shown in Figure 2 and the
zero-lag window function parameters are in Table 1.
In this paper we focus on a spatially open CDM model and a spatially flat CDM model
with a cosmological constant Λ. These low density models are largely consistent with current
observational constraints.10 For recent discussions see Park et al. (1998), Retzlaff et al. (1998),
Croft et al. (1999), and Peebles (1999).
The models have Gaussian, adiabatic primordial energy-density power spectra. The
flat-Λ model CMB anisotropy computations use a scale-invariant energy-density perturbation
power spectrum (Harrison 1970; Peebles & Yu 1970; Zel'dovich 1972), as predicted in the
simplest spatially-flat inflation models (Guth 1981; Kazanas 1980; Sato 1981). The open model
computations use the energy-density power spectrum (Ratra & Peebles 1994, 1995; Bucher,
Goldhaber, & Turok 1995; Yamamoto, Sasaki, & Tanaka 1995) predicted in the simplest
open-bubble inflation models (Gott 1982). The computation of the CMB anisotropy spectra is
described by Stompor (1994) and Sugiyama (1995).
As discussed in R99a, the spectra are parameterized by their quadrupole-moment amplitude
10 While not considered in this paper, a time-variable cosmological "constant" dominated spatially-flat model is
also largely consistent with current data (e.g., Peebles & Ratra 1988; Sugiyama & Sato 1992; Ratra & Quillen 1992;
Frieman & Waga 1998; Ferreira & Joyce 1998; Wang & Steinhardt 1998; Carroll 1998; Hu et al. 1999; Huterer &
Turner 1999; Liddle & Scherrer 1999; Starobinsky 1998).
-- 4 --
Qrms−PS, the nonrelativistic-mass density parameter Ω0, the baryonic-mass density parameter ΩB,
and the age of the universe t0. The spectra are computed for a range of Ω0 spanning the interval
0.1 to 1 in steps of 0.1, for a range of ΩBh2 [the Hubble parameter h = H0/(100 km s−1 Mpc−1)]
spanning the interval 0.005 to 0.029 in steps of 0.004, and for a range of t0 spanning the
interval 10 to 20 Gyr in steps of 2 Gyr.
In total 798 spectra were computed to cover the
cosmological-parameter spaces of the open and flat-Λ models. Figure 2 shows examples of the
CMB anisotropy spectra used in our analysis. Other examples are in Figure 2 of R99a and Figure
1 of R99b.
While it is of interest to also consider other cosmological parameters, such as tilt or gravity
wave fraction or a time-variable cosmological "constant" (instead of a constant Λ), to make the
problem tractable we have focussed on the four parameters mentioned above. We emphasize
however that the results of the analysis are model dependent. For instance, a time-variable Λ
model would likely lead to a different constraint on Ω0 than that derived below in the constant Λ
model.
Following GRGS, for each of the 798 spectra considered the "bare" likelihood function is
computed at the nominal beamwidth and calibration, as well as at a number of other values of
the beamwidth and calibration determined from the measurement uncertainties. The likelihood
function used in the derivation of the central values and limits is determined by integrating
(marginalizing) the bare likelihood function over the beamwidth and calibration uncertainties
with weights determined by the measured probability distribution functions of the beamwidth
and the calibration. See GRGS for a more detailed discussion. When marginalizing over the
beamwidth uncertainty we have checked in a few selected cases that the five-point Gauss-Hermite
quadrature summation approximation to the integral agrees extremely well with the three-point
Gauss-Hermite approximation used by GRGS (and for most of the analysis in this paper). The
likelihoods are a function of four parameters mentioned above: Qrms−PS, Ω0, ΩBh2, and t0. We
also compute marginalized likelihood functions by integrating over one or more of these parameters
after assuming a uniform prior in the relevant parameters. The prior is set to zero outside the
ranges considered for the parameters. GRGS and R99a describe the prescription used to determine
central values and limits from the likelihood functions. In what follows we consider 1, 2, and 3 σ
highest posterior density limits which include 68.3, 95.4, and 99.7% of the area.
3. Results and Discussion
Table 2 lists the derived values of Qrms−PS and bandtemperature δTl for the flat bandpower
spectrum, for various combinations of the I, II, and III data. These numerical values account
for the beamwidth and calibration uncertainties. The last two δTl entries in Table 2 are quite
consistent with those derived by P97; the small differences reflect the different methods used to
account for beamwidth and calibration uncertainties here and in P97.
-- 5 --
For the flat bandpower spectrum the combined I, II, and III data average 1 σ δTl error bar
is ∼ 25%11 : Python data results in a very significant detection of CMB anisotropy, even after
accounting for beamwidth and calibration uncertainties. Note that the calibration uncertainty,
20%, is the most important contributor to this error bar.
A number of other interesting conclusions follow from the entries in Table 2. Comparing the
I+II and IIIL results, which are from experiments which probe almost identical angular scales, we
see that the IIIL amplitude is ∼ 1 σ higher than the I+II amplitude. Comparing the result from
the analysis of the coadded I and IIA data (which are from experiments with identical window
functions) and the result from the full analysis of the I and IIA data (which takes into account all
the spatial correlations), we see that the deduced amplitudes are almost identical. This is probably
mostly a reflection of the fact that the individual I and IIA amplitudes are almost identical (see
Table 2).
As discussed in R99a and R99b, the four-dimensional posterior probability density distribution
function L(Qrms−PS, Ω0, ΩBh2, t0) is nicely peaked in the Qrms−PS direction but fairly flat in
the other three directions. Marginalizing over Qrms−PS results in a three-dimensional posterior
distribution L(Ω0, ΩBh2, t0) which is steeper, but still relatively flat. As a consequence, limits
derived from the four- and three-dimensional posterior distributions are generally not highly
statistically significant. We therefore do not show contour plots of these functions here.
Marginalizing over Qrms−PS and one other parameter results in two-dimensional posterior
probability distributions which are more peaked. See Figures 3 and 4. As in the ARGO (R99a)
and combination (R99b) data set analyses, in some cases these peaks are at an edge of the
parameter range considered.
Figure 3 shows that the two-dimensional posterior distributions allow one to distinguish
between different regions of parameter space at a fairly high formal level of confidence. For
instance, the open model near Ω0 ∼ 0.75, ΩBh2 ∼ 0.03, and t0 ∼ 20 Gyr, and the flat-Λ model
near Ω0 ∼ 0.6, ΩBh2 ∼ 0.03, and t0 ∼ 20 Gyr, are both formally ruled out at ∼ 3 σ confidence.
However, we emphasize, as discussed in R99a and R99b, care must be exercised when interpreting
the discriminative power of these formal limits, since they depend sensitively on the fact that the
uniform prior has been set to zero outside the range of the parameter space we have considered.
Figure 4 shows the contours of the two-dimensional posterior distribution for Qrms−PS and
Ω0, derived by marginalizing the four-dimensional distribution over ΩBh2 and t0. These are shown
for the combined Python I, II, and III data, the DMR data, and three combinations of data from
the SP94, ARGO, MAX 4+5, White Dish, and SuZIE experiments (R99b), for both the open and
flat-Λ models. Constraints on these parameters from the combined Python data are consistent
with those from the DMR data for the flat-Λ models, panel a), while for the open model, panel b),
11 For comparison, the corresponding DMR 1 σ δTl error bar is ∼ 10 − 12% (depending on model, G´orski et al.
1998).
-- 6 --
>
∼ 0.2 (0.35). The combined Python data amplitudes are a
consistency at 2 σ (1 σ) requires Ω0
little higher than those derived from the other small-scale data combinations, panels c) − h), but
at 2 σ confidence the various amplitudes are mostly consistent.
Figure 5 shows the one-dimensional posterior distribution functions for Ω0, ΩBh2, t0, and
Qrms−PS, derived by marginalizing the four-dimensional posterior distribution over the other three
parameters. From these one-dimensional distributions, the combined I, II, and III data favors an
open (flat-Λ) model with Ω0 = 0.19 (0.10), or ΩBh2 = 0.005 (0.005), or t0 = 10 (13) Gyr, amongst
the models considered. At 2 σ confidence the combined Python data formally rule out only small
regions of parameter space. From the one-dimensional distributions of Figure 5, the data requires
Ω0 < 0.71 or > 0.8 (Ω0 < 0.55 or > 0.63), or ΩBh2 < 0.028 (ΩBh2 < 0.028), or t0 < 20 Gyr (t0
< 20 Gyr) for the open (flat-Λ) model at 2 σ. As discussed in R99a and R99b, care is needed
when interpreting the discriminative power of these formal limits. These papers also discuss a
more conservative Gaussian posterior distribution limit prescription. Using this more conservative
prescription, we find only an upper 1 σ limit on Ω0 ( <
∼ 0.5) in the open model.
While the statistical significance of the constraints on cosmological parameters is not high,
it is reassuring that the combined Python data favor low-density, low ΩBh2, young models,
consistent with some of the indications from the combinations of CMB anisotropy data considered
by R99b, and the indications from most recent non-CMB observations (see discussion in R99b).
The peak values of the one-dimensional posterior distributions shown in Figure 5 are listed
in the figure caption for the case when the four-dimensional posterior distributions are normalized
such that L(Qrms−PS = 0 µK) = 1. With this normalization, marginalizing over the remaining
parameter the fully marginalized posterior distributions are 2 × 10106(1 × 10106) for the open
(flat-Λ) model and the combined Python data. This is qualitatively consistent with the indication
from panels a) and b) of Figure 5 that the most-favored open model is somewhat more favored
than the most-favored flat-Λ one.
4. Conclusion
The combined Python I, II, and III data results derived here are mostly consistent with
those derived from the DMR, SP94, ARGO, MAX 4+5, White Dish and SuZIE data. The
combined Python data significantly constrains Qrms−PS (for the flat bandpower spectrum
Qrms−PS = 40 +12
−8 µK at 1 σ) and weakly favors low-density, low ΩBh2, young models.
We acknowledge helpful discussions with D. Alvarez, M. Dragovan, G. Griffin, J. Kovac, and
J. Ruhl. This work was partially carried out at the Infrared Processing and Analysis Center
and the Jet Propulsion Laboratory, California Institute of Technology, under a contract with the
National Aeronautics and Space Administration. KG also acknowledges support from NASA ADP
grant NASA-1260. BR and GR acknowledge support from NSF grant EPS-9550487 with matching
-- 7 --
support from the state of Kansas and from a K∗STAR First award. GR also acknowledges support
from a PRAXIS XXI program of FCT (Portugal) grant. RS acknowledges support from NASA
AISRP grant NAG-3941 and help from Polish Scientific Committee (KBN) grant 2P03D00813.
-- 8 --
Table 1: Numerical Values for the Zero-Lag Window Function Parametersa
Python I/II
Python IIIL
Python IIIS
le−0.5
53
52
128
le
91.7
87.7
171
lm
73
72
176
le−0.5 pI(Wl)
99
98
230
1.34
1.30
0.623
aThe value of l where Wl is largest, lm, the two values of l where Wl
le = I(lWl)/I(Wl), and I(Wl) = P∞
l=2(l + 0.5)Wl/{l(l + 1)}.
e−0.5 = e−0.5Wlm, le−0.5 , the effective multipole,
-- 9 --
Table 2: Numerical Values for Qrms−PS and δTl
Bandpower Spectrum
from Likelihood Analyses Assuming a Flat
Data Seta
Qrms−PS
b Ave. Abs. Err.c Ave. Frac. Err.d
I
IIA
C(I+IIA)
I+IIA
IIBf
II
C(I+IIA)+IIB
I+II
IIIL
IIIS
III
I+II+IIIL
I+II+III
(µK)
39 56
28
39 57
27
40 57
29
40 56
29
9.3 20
0
31 44
23
34 47
26
34 46
25
41 55
32
42 56
33
41 54
33
39 51
31
40 52
32
(µK)
14
15
14
14
10
10
11
11
11
11
10
10
10
36%
38%
35%
35%
110%
34%
31%
31%
28%
27%
25%
27%
25%
b
δTl
(µK)
61 87
43
60 88
43
62 89
45
61 87
45
14 31
0
48 68
35
53 73
40
52 72
39
64 84
49
66 87
51
64 83
51
60 80
47
63 81
50
LRe
2 × 1014
9 × 1012
3 × 1031
2 × 1031
1
1 × 1012
1 × 1030
1 × 1030
3 × 1033
2 × 1029
6 × 1068
3 × 1067
1 × 10105
aC(...+...) means that the data from the two sets of observations have been coadded prior to analysis, II refers to the
combined IIA and IIB data, and III refers to the combined IIIL and IIIS data.
bFor each data set, the first of the three entries is where the posterior probability density distribution function peaks
and the vertical pair of numbers are the ±1 σ (68.3% highest posterior density) values.
cAverage absolute error on Qrms−PS in µK.
dAverage fractional error, as a fraction of the central value.
eLikelihood ratio.
f IIB does not have a 2 σ highest posterior density detection; the appropriate equal tail 2 σ upper limits are 50 µK
(Qrms−PS) and 77 µK (δTl).
-- 10 --
REFERENCES
Alvarez, D.L. 1996, Princeton University Ph.D. Thesis
Bucher, M., Goldhaber, A.S., & Turok, N. 1995, Phys. Rev. D, 52, 3314
Carroll, S.M. 1998, Phys. Rev. Lett., 81, 3067
Church, S.E., Ganga, K.M., Ade, P.A.R., Holzapfel, W.L., Mauskopf, P.D., Wilbanks, T.M., &
Lange, A.E. 1997, ApJ, 484, 523
Croft, R.A.C., Weinberg, D.H., Pettini, M., Hernquist, L., & Katz, N. 1999, ApJ, in press
de Bernardis, P., et al. 1994, ApJ, 422, L33
Dragovan, M., Ruhl, J.E., Novak, G., Platt, S.R., Crone, B., Pernic, R., & Peterson, J.B. 1994,
ApJ, 427, L67 (D94)
Ferreira, P.G., & Joyce, M. 1998, Phys. Rev. D, 58, 023503
Frieman, J.A., & Waga, I. 1998, Phys. Rev. D, 57, 4642
Ganga, K., Ratra, B., Church, S.E., Sugiyama, N., Ade, P.A.R., Holzapfel, W.L., Mauskopf, P.D.,
& Lange, A.E. 1997b, ApJ, 484, 517
Ganga, K., Ratra, B., Gundersen, J.O., & Sugiyama, N. 1997a, ApJ, 484, 7 (GRGS)
Ganga, K., Ratra, B., Lim, M.A., Sugiyama, N., & Tanaka, S.T. 1998, ApJS, 114, 165
G´orski, K.M., Ratra, B., Stompor, R., Sugiyama, N., & Banday, A.J. 1998, ApJS, 114, 1
Gott, J.R. 1982, Nature, 295, 304
Gundersen, J.O., et al. 1995, ApJ, 443, L57
Guth, A. 1981, Phys. Rev. D, 23, 347
Harrison, E.R. 1970, Phys. Rev. D, 1, 2726
Hu, W., Eisenstein, D.J., Tegmark, M., & White, M. 1999, Phys. Rev. D, 59, 023512
Huterer, D., & Turner, M.S. 1999, Phys. Rev. Lett., submitted
Kazanas, D. 1980, ApJ, 241, L59
Liddle, A.R., & Scherrer, R.J. 1999, Phys. Rev. D, 59, 023509
Lim, M.A., et al. 1996, ApJ, 469, L69
Park, C., Colley, W.N., Gott, J.R., Ratra, B., Spergel, D.N., & Sugiyama, N. 1998, ApJ, 506, 473
Peebles, P.J.E. 1999, PASP, 111, 274
Peebles, P.J.E., & Ratra, B. 1988, ApJ, 325, L17
Peebles, P.J.E., & Yu, J.T. 1970, ApJ, 162, 815
Platt, S.R., Kovac, J., Dragovan, M., Peterson, J.B., & Ruhl, J.E. 1997, ApJ, 475, L1 (P97)
-- 11 --
Ratra, B., Ganga, K., Stompor, R., Sugiyama, N., de Bernardis, P., & G´orski, K.M. 1999a, ApJ,
510, 11 (R99a)
Ratra, B., Ganga, K., Sugiyama, N., Tucker, G.S., Griffin, G.S., Nguyen, H.T., & Peterson, J.B.
1998, ApJ, 505, 8
Ratra, B., & Peebles, P.J.E. 1994, ApJ, 432, L5
Ratra, B., & Peebles, P.J.E. 1995, Phys. Rev. D, 52, 1837
Ratra, B., & Quillen, A. 1992, MNRAS, 259, 738
Ratra, B., Stompor, R., Ganga, K., Rocha, G., Sugiyama, N., & G´orski, K.M. 1999b, ApJ, 517, in
press (R99b)
Retzlaff, J., Borgani, S., Gottlober, S., Klypin, A., & Muller, V. 1998, New Astron., 3, 631
Ruhl, J. 1993, Princeton University Ph.D. Thesis
Ruhl, J.E., Dragovan, M., Novak, G., Platt, S.R., Crone, B.K., & Pernic, R.W. 1995a, Astro.
Lett. and Comm., 32, 249
Ruhl, J.E., Dragovan, M., Platt, S.R., Kovac, J., & Novak, G. 1995b, ApJ, 453, L1 (R95)
Sato, K. 1981, Phys. Lett. B, 99, 66
Starobinsky, A.A. 1998, JETP Lett., 68, 757
Stompor, R. 1994, A&A, 287, 693
Stompor, R. 1997, in Microwave Background Anisotropies, ed. F.R. Bouchet, R. Gispert, B.
Guiderdoni, & J. Tran Thanh Van (Gif-sur-Yvette: Editions Frontieres), 91
Sugiyama, N. 1995, ApJS, 100, 281
Sugiyama, N., & Sato, K. 1992, ApJ, 387, 439
Tanaka, S.T., et al. 1996, ApJ, 468, L81
Tucker, G.S., Griffin, G.S., Nguyen, H.T., & Peterson, J.B. 1993, ApJ, 419, L45
Wang, L., & Steinhardt, P.J. 1998, ApJ, 508, 483
Yamamoto, K., Sasaki, M., & Tanaka, T. 1995, ApJ, 455, 412
Zel'dovich, Ya. B. 1972, MNRAS, 160, 1P
This preprint was prepared with the AAS LATEX macros v4.0.
-- 12 --
Figure Captions
Fig. 1. -- Measured thermodynamic temperature differences (with ±1-σ error bars). Panel a) shows
Python I (open circles), IIA (filled squares, offset horizontally from true positions for clarity), IIB
(open squares), and IIIL (crosses) data, while panel b) shows Python IIIS data (crosses).
Fig. 2. -- CMB anisotropy multipole moments l(l + 1)Cl/(2π) × 1010 (solid lines, scale on left
axis, note that these are fractional anisotropy moments and thus dimensionless) as a function of
multipole l, for selected models normalized to the DMR maps (G´orski et al. 1998; Stompor 1997).
Panels a) − c) show selected flat-Λ models. The heavy lines are the Ω0 = 0.1, ΩBh2 = 0.005, and
t0 = 12 Gyr case, which is close to where the combined Python data likelihoods (marginalized over
all but one parameter at a time) are at a maximum. Panel a) shows five ΩBh2 = 0.005, t0 = 12
Gyr models with Ω0 = 0.1, 0.3, 0.5, 0.7, and 0.9 in descending order at the l ∼ 200 peaks. Panel
b) shows seven Ω0 = 0.1, t0 = 12 Gyr models with ΩBh2 = 0.029, 0.025, 0.021, 0.017, 0.013, 0.009,
and 0.005 in descending order at the l ∼ 200 peaks. Panel c) shows six Ω0 = 0.1, ΩBh2 = 0.005
models with t0 = 20, 18, 16, 14, 12, and 10 Gyr in descending order at the l ∼ 200 peaks. Panels
d) − f ) show selected open models. The heavy lines are the Ω0 = 0.2, ΩBh2 = 0.005, and t0 = 10
Gyr case, which is close to where the combined Python data likelihoods (marginalized over all but
one parameter at a time) are at a maximum. Panel d) shows five ΩBh2 = 0.005, t0 = 10 Gyr
models with Ω0 = 1, 0.8, 0.6, 0.4, and 0.2 from left to right at the peaks (the peak of the Ω0 = 0.2
model is off scale). Panel e) shows seven Ω0 = 0.2, t0 = 10 Gyr models with ΩBh2 = 0.029, 0.025,
0.021, 0.017, 0.013, 0.009, and 0.005 in descending order at l ∼ 400. Panel f ) shows six Ω0 = 0.2,
ΩBh2 = 0.005 models with t0 = 20, 18, 16, 14, 12, and 10 Gyr in descending order at l ∼ 400.
Also shown are the Python zero-lag window functions Wl (scale on right axis): I/II (long-dashed
lines), IIIL (short-dashed lines), and IIIS (dotted lines). See Table 1 for Wl-parameter values.
Fig. 3. -- Confidence contours and maxima of the combined Python data two-dimensional posterior
probability density distribution functions, as a function of the two parameters on the axes of each
panel (derived by marginalizing the four-dimensional posterior distributions over the other two
parameters). Dashed lines (crosses) show the contours (maxima) of the open case and solid lines
(solid circles) show those of the flat-Λ model. Panel a) shows the (ΩBh2, Ω0) plane, and panel b)
shows the (t0, Ω0) plane.
Fig. 4. -- Confidence contours and maxima of the two-dimensional (Qrms−PS, Ω0) posterior
probability density distribution functions. Panels a), c), e), & g) in the left column show the flat-Λ
model and panels b), d), f ), & h) in the right column show the open model. Note the different scale
on the vertical (Qrms−PS) axes of pairs of panels in each row. Heavy lines show the ±1 and ±2 σ
confidence limits and solid circles show the maxima of the two-dimensional posterior distributions
derived from the combined Python I, II, and III data. Shaded regions show the two-dimensional
posterior distribution 1 σ (denser shading) and 2 σ (less dense shading) confidence regions for the
DMR data (G´orski et al. 1998; Stompor 1997) in panels a) & b); for the SP94, ARGO, MAX 4
-- 13 --
and 5, White Dish and SuZIE data combination (R99b) in panels c) & d); for the previous data
combination excluding SuZIE (R99b) in panels e) & f ); and for the SP94Ka, MAX 4 ID, and
MAX 5 HR data combination (R99b) in panels g) & h). The DMR results are a composite of
those from analyses of the two extreme data sets: i) galactic frame with quadrupole included and
correcting for faint high-latitude galactic emission; and ii) ecliptic frame with quadrupole excluded
and no other galactic emission correction (G´orski et al. 1998). In panels c) − h) crosses show the
maxima of the appropriate non-Python data two-dimensional posterior distributions.
Fig. 5. -- One-dimensional posterior probability density distribution functions for Ω0, ΩBh2, t0,
and Qrms−PS (derived by marginalizing the four-dimensional one over the other three parameters)
in the open and flat-Λ models. These have been renormalized to unity at the peaks. Dotted
vertical lines show the confidence limits derived from these one-dimensional posterior distributions
and solid vertical lines in panels g) and h) show the ±1 and ±2 σ confidence limits derived by
projecting the combined Python I, II, and III data four-dimensional posterior distributions. The
2 σ DMR (marginalized and projected) confidence limits in panels g) and h) are a composite of
those from the two extreme DMR data sets (see caption of Figure 4). When the four-dimensional
posterior distributions are normalized such that L(Qrms−PS = 0 µK) = 1, the peak values of the
one-dimensional distributions shown in panels a) − h) are 2 × 10106, 3 × 10106, 6 × 10107, 7 × 10107,
1 × 10105, 2 × 10105, 9 × 10104, and 6 × 10104, respectively.
Figure 1a)
Figure 1b)
Figure 2
Figure 3
Figure 4
Figure 5.1
Figure 5.2
|
astro-ph/9602005 | 1 | 9602 | 1996-02-01T20:28:07 | Analysis of the Type Ia Supernova SN1994D | [
"astro-ph"
] | We present an analysis of the observed light curves and spectra of the Type Ia supernova SN1994D in the galaxy NGC 4526. The sensitivity of theoretical light curves and spectra to the underlying hydrodynamical model is discussed.
The calculations are consistent with respect to the explosion mechanism, the optical and infrared light curves, and the spectral evolution, leaving the description of the nuclear burning front and the structure of the white dwarf as the only free parameters.
The explosions are calculated using a one-dimensional Lagrangian code including a nuclear network. Subsequently, the light curves are constructed. Detailed NLTE-spectra are computed for several instants of time using the density, chemical, and luminosity structure resulting from the light curve code.
Different scenarios have been compared to the observations including detonations, deflagrations, delayed detonations, tamped detonations and Helium detonations. Only one of our delayed detonations models shows good agreement with the observations of SN1994D both for the B, V, R and I colors and the spectral evolution. The classical deflagration model W7 does not reproduce the high Si velocities as observed early on. The deflagration velocity is close to the laminar deflagration $(v =0.03 c_s)$, and the transition from the deflagration to the detonation occurs at $\rho_{tr} = 2.~10^7 g/cm^{3}$. The initial central density of the white dwarf (WD) is $2.7~10^9 g/cm^{3}$.
The distance to SN1994D is determined to $16.2 \pm 2 Mpc$. The explosion took | astro-ph | astro-ph |
ANALYSIS OF THE TYPE IA SUPERNOVA SN1994D
P. H OFLICH
Harvard University, Center for Astrophysics
60 Garden Str., Cambridge, MA 02138, USA
1. Introduction
On Mar. 7, 1994, Treffers, Filippenko and Van Dyk (1994) discovered the
Type Ia supernova (SN Ia) 1994D in the SO galaxy NGC 4526, a member
of the Virgo cluster which was was about 12 days before maximum light.
Subsequently, both the light curve in Johnson's UBVRI colors and spectra
a have been monitored by several groups (e.g. Filippenko et al. private
communication, Smith et al., 1995) making this supernova one of the best
observed until now. In addition, starting at Mar 12, IR-spectra are reported
to be taken at the Royal Greenwich Observatory showing a P-Cygni line at
about 1.0 µm which, likely, is due to He I (Meikle 1995).
It is generally accepted that SN Ia are thermonuclear explosions of
carbon-oxygen white dwarfs (WD) (Hoyle & Fowler 1960). However, details
of the scenario are still under debate. For discussions of various theoretical
aspects see e.g. Wheeler & Harkness 1990, Canal 1995, Nomoto et al. 1995,
Hoflich & Khokhlov 1995, Woosley & Weaver 1995). What we observe as
a supernova event is not the explosion itself but the light emitted from
a rapidly expanding envelope produced by the stellar explosion. As the
photosphere recedes, deeper layers of the ejecta become visible. A detailed
analysis of the light curves (LC) and spectra gives us the opportunity to
determine the density, velocity and composition structure of the ejecta.
2. Comparison of light curves and spectra
Our analysis is based on observations of SN1994D by the supernova group
at the Center for Astrophysics. For the first time, a detailed analysis of a
Type Ia supernova is presented which is consistent both with respect to the
explosion mechanism, the optical and infrared light curves, and the spectral
evolution. The only free parameters are the initial structure of the WD and
2
P. H OFLICH
Figure 1. Observed V and B LCs for SN1994D in comparison to the theoretical LCs
of the deflagration W7, the delayed detonation N21, the helium detonation HeD10, the
pulsating delayed detonation PDD3, and the envelope model DET2env4.
the description of the burning front. Our approach provides a direct link
between observational properties and both the explosion mechanism and,
maybe, the progenitor evolution.
The explosions are calculated using one-dimensional Lagrangian hy-
dro with artificial viscosity (Khokhlov, 1991) and radiation-hydro codes
(Hoflich & Khokhlov 1995) including a nuclear network. Subsequently, the
LC are constructed. Spectra are computed for several instants of time us-
ing the density, chemical, and luminosity structure resulting from the light
curve code which includes the solution of the radiation transport implic-
itly via the moment equations, expansion opacities, a detailed Monte-Carlo
γ−ray transport, and a detailed equation of state. Our NLTE code for calcu-
lating synthetic spectra solves the relativistic radiation transport equations
in comoving frame consistently with the statistical equations and ioniza-
tion due to γ radiation for the most important elements (He, C, O, Ne, Na,
Mg, Si, S, Ca, Fe). About 300,000 additional lines are included assuming
LTE-level populations and an equivalent-two-level approach for the source
functions. For more details on technical aspects and a discussion of the
relation between observational properties and the underlying model, see
Khokhlov et al. (1993), Hoflich (1995), Hoflich & Khokhlov (1995), Hoflich
et al. (1995, this volume) and references therein.
Although not presented in detail for all models available (see Nomoto
et al., 1984, Khokhlov et al. 1993, Hoflich et al. 1994, Hoflich, 1995, Hoflich
& Khokhlov, 1995), we have compared the observed optical and infrared
light curves of SN1994D for deflagration (W7, DF1), pulsating delayed
detonation (PDD1a-c,3-9), delayed detonation (N21, N32, M35-39), helium
detonations (HeD2-12), and envelope models (DET2env2-4) being a crude
representation of a merger scenario.
The pulsating delayed detonation models, helium detonation and enve-
lope models can be ruled out because of the shape of their V and B light
ANALYSIS OF THE TYPE IA SUPERNOVA SN1994D
3
Figure 2.
detonation model M36.
Some Abundances as a function of the expansion velocity for the delayed
Figure 3. Observed LCs of SN1994D in comparison to the theoretical LCs of M36.
curves (Fig.1). W7 and N32 that fit observations of several SNe Ia (Muller &
Hoflich, 1994) provide better fits to SN1994D, but they fail to reproduce the
IR light curves (Hoflich 1995) and, less critical, the post-maximum decline
in V. One may conclude that SN1994D is no 'standard' Type Ia supernova,
but it is close to standard. We find best agreement between theory and ob-
servation for a the delayed detonation model M36 with ρtr = 2.4 107g/cm3
and a deflagration velocity of 0.03 times the sound velocity (Figs. 2 & 3).
The initial central density of the WD is 2.7 109g/cm3, i.e. about 20 %
lower than in our delayed detonation models previously considered. The
lower density may be understood in terms of a higher accretion rate on
the progenitor. During the explosion, 0.6M⊙ of 56N i are produced. Similar
4
P. H OFLICH
Figure 4. Synthetic spectrum at day 15 for M36 compared to observations at Mar. 16th.
models with ρtr of 2.0 107 and 3.0 107g/cm3 and Ni productions of 0.51
and 0.67 M⊙, respectively, are not able to reproduce the observations.
From the vertical shift of the light curves, the time of explosion can be
determined to be between JD 244 9414.5 and 244 9415.5, meaning SN1994D
was discovered just about 3 to 4 days after the explosion! Our models
are consistent with no interstellar reddening. Taking the uncertainties in
our models into account, we determine the distance of NGC 4526 to be
16 ± 2 M pc. The explosion took place between March 3rd and 4th, 1994.
The NLTE-analysis of the spectra confirmed the results from the LC
analysis and give an insight to several new aspects. Taking into account
that we do not allow for any artificial adjustment to provide better fits,
M36 reproduces the observed spectra reasonably well (Fig. 4), including the
evolution with time (Hoflich, 1995). The agreement of the velocity shifts
both of the elements of partial burning such as Si, S, O and also the iron
rich elements indicate that the chemical and density profiles of M36 seem
to resemble SN1994D. In the spectrum from March 11, the Doppler shift
of the absorption minimum of Si II indicates photospheric expansion ve-
locities of ≈ 15000km/sec with wings reaching out to more than 22000
km/sec. This strongly suggest the presence of a significant amount of Si
at high velocities. Our delayed detonation model M36 provide a Si mass
fraction of 15 % up to the outer layers ( ≈ 25000km/sec, Fig. 1). For com-
parison, the classical deflagration model W7 does not show Si at velocities
ANALYSIS OF THE TYPE IA SUPERNOVA SN1994D
5
≤ 16400km/sec because the smaller flame speed in the outer layers and,
consequently, the larger pre-expansion of the WD envelope during the ex-
plosion. In our previous analysis, we had to report the serious problem that
the Ti features are too strong and, artificially, we had to reduce Ti in the
outer layers. In recent calculations for delayed detonation models (Hoflich
et al. 1995), we find that the Ti problem vanishes if we use a larger nuclear
network and start with a metallicity according to a population II star.
Finally, we want to mention some of the limitations of our study. Despite
similarities in both the spectra and light curves, the photospheric expansion
velocities inferred from observed spectra show strong individual variations
(e.g. Branch 1987 and Barbon et al. 1990). Different models are needed to
fit the light curves (Muller & Hoflich 1994, Hoflich et al. 1994). Therefore,
the results of our study must not be generalized. Although we treat several
elements in full NLTE, detailed atomic models shall be implemented for the
iron group elements, namely Ti, Co, Ni. Our approach to the line scattering
by using an equivalent level approach for the 'LTE-lines' seems to work, but
it must be regarded only as a first step towards a more consistent treatment.
Finally, M36 produces only 0.003 M⊙ of He in the region of α rich freeze-out
whereas, a much larger amount is needed to produce a strong He feature at
1.µ. For a more detailed discussion of the latter problem, see Meikle 1995.
3. References
Barbon R.,Bennetti S.,Cappalaro E.,Rosino L.,Turatto M. (1990), A&A, 237, 79
Branch D. (1987), ApJ, 316, L81
Canal R. (1995),Les Houches Lectures, eds. J. Audouze et al., Elsevier, in press
Hoflich P. (1995), ApJ , 443, 533
Hoflich P., Khokhlov A. (1995), ApJ , in press
Hoflich P., Khokhlov A., Wheeler C. (1994), ApJ , 444, 831
Hoflich P., Khokhlov A., Nomoto K., Thielemann F.K, Wheeler C.J. (1995), this
volume
Hoyle P., Fowler , (1960) ApJ , 132, 565
Khokhlov A., Muller E., Hoflich P. (1993), A&A, 270, 23
Khokhlov A., (1991), A&A, 245, 114
Meikle P. (1995), this volume
Muller E., Hoflich P., (1994), A&A, 281, 51
Nomoto K., Yamaoka H., Shigeyama T., Iwamoto K., (1995), in: Supernovae, ed.
R.A. McCray, Cambridge University Press, in press
Nomoto K., Thielemann F.-K., Yokoi K. (1984), ApJ, 286, 644
Smith et al., (1994), ApJ, in preparation combined observations of CTIO and CfA
Treffers, Filippenko, Van Dyke (1994), IAU-circular 5946
Wheeler J. C., Harkness R .P., (1990), Rep. Prog. Phys., 531467
Woosley S. E., Weaver T. A. (1995),Les Houches Lectures, eds. J. Audouze et al.,
Elsevier, in press
|
astro-ph/0511124 | 1 | 0511 | 2005-11-04T09:52:00 | Comments on Long-Wavelength Backreaction and Dark Energy | [
"astro-ph",
"gr-qc",
"hep-ph",
"hep-th"
] | In this brief note we comment on a recent attempt by Martineau and Brandenberger (astro-ph/0510523) to explain the acceleration of the Universe using the back-reaction of long-wavelength perturbations associated with isocurvature perturbation modes. | astro-ph | astro-ph | Comments on Long-Wavelength Backreaction and Dark Energy
DFPD-05/A/21
Edward W. Kolb∗
Particle Astrophysics Center, Fermi National Accelerator Laboratory,
Batavia, Illinois 60510-0500, USA
and Department of Astronomy and Astrophysics, Enrico Fermi Institute,
University of Chicago, Chicago, Illinois 60637-1433 USA
Sabino Matarrese†
Dipartimento di Fisica "G. Galilei" Universit`a di Padova,
INFN Sezione di Padova, via Marzolo 8, Padova I-35131, Italy
Antonio Riotto‡
CERN, Theory Division, CH-1211 Geneva 23, Switzerland
(Dated: October 20, 2019)
Abstract
In this brief note we comment on a recent attempt by Martineau and Brandenberger [1] to
explain the acceleration of the Universe using the back-reaction of long-wavelength perturbations
associated with isocurvature perturbation modes.
PACS numbers: 98.80.Cq, 95.35.+d, 4.62.+v
5
0
0
2
v
o
N
4
1
v
4
2
1
1
1
5
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
∗Electronic address: [email protected]
†Electronic address: [email protected]
‡Electronic address: [email protected]
1
In this brief note we would like to comment on a work by Martineau and Brandenberger
(MB) [1], who have recently presented a solution to the dark energy problem making use of
the gravitational backreaction of long wavelength (super-Hubble) fluctuation modes on the
background metric.
The issue of the physical influence of the super-Hubble modes on cosmological observables,
such as the local Hubble expansion rate, has already been posed in the literature [2, 3, 4,
5, 6, 7]. Sometime ago, some of us [8] have shown that if adiabaticity holds, then at any
order in perturbation theory super-Hubble perturbations do not have any impact on local
physical observables (if gradients are neglected). Let us briefly summarize here the line of
reasoning. Our starting point is the Arnowitt-Deser-Misner (ADM) formalism, with metric
ds2 = −N 2 dt2 + Ni dt dxi + γij dxi dxj.
(1)
The three-metric γij, and the lapse and the shift functions N and Ni, describe the evo-
lution of the spacelike hypersurfaces. In the ADM formalism the equations simplify consid-
erably if we set N i = 0. We may then express the spatial metric in the form
γij = exph2α(t, xi)i hij(xi),
(2)
where the conformal factor exp[α(t, xi)] may be interpreted as the spatially dependent scale
factor. The time-independent three-metric, hij(xi), (with unit determinant) describes the
three geometry of the conformally transformed space. Within linear perturbation theory,
α(t, xi) would contribute only to scalar perturbations, whereas hij contains another scalar,
as well as vector and tensor perturbations. Vector and tensor perturbations are necessary
to satisfy Einstein equations beyond linear order, however in the long wavelength limit they
do not affect the equations for the scalar perturbations. Since exp[α(t, xi)] is interpreted as
a scale factor, we can define the local Hubble parameter H(t, xi) = α(t, xi)/N(t, xi).
Now consider a Universe filled with ideal fluid(s) described by an energy-momentum
tensor of the form Tµν = (ρ + P ) uµuν + P gµν, where ρ and P are the energy density and
pressure, respectively. The four-velocity vector can be chosen locally to be uµ = (1,~0). In
the case of a single fluid, this amounts to saying that a volume measured by a local observer
is comoving with the energy flow of the fluid. In the multi-fluid case, the volume comoving
with the total energy density ρ is not the same as the individual volumes comoving with the
individual energy density components. However, one can show that on super-Hubble scales
2
all comoving volumes become equivalent, and therefore our choice of the four velocity is well
justified.
We consider the effect on the local expansion rate of the Universe and set
exphα(t, xi)i = a(t) exph−ψ(t, xi)i
(3)
and choose to work in the synchronous gauge for which N = 1 (together with N i = 0). In
this gauge the field equations look just like the familiar homogeneous Friedmann-Robertson-
Walker equations point by point. As mentioned, for a generic set of fluids we may safely take
the four velocity to be uµ = (1,~0) on super-Hubble scales, from which we define the local
expansion rate of tangential surfaces orthogonal to the fluid flow, i.e., the local expansion
rate, to be Θ ≡ Dµuµ. In the synchronous gauge this quantity coincides with 3H.
H(xi, t) = 1
Using Eq. (3), we find that at any order in perturbation theory the local expansion rate is
3Θ(xi, t) = a/a − ψ. We infer that the local physical expansion rate is influenced
by long-wavelength perturbations only if the gravitational potential is time dependent (if we
ignore spatial gradients in ψ). Let us analyze the case of an adiabatic fluid. From Einstein's
equations for a globally flat space we deduce (on super-Hubble scales) that locally
3H 2 = 8πGρ = −8πG P + 2 H.
(4)
If the pressure P is a unique function of the energy density, we may expand ρ = ρ0 + ∆ρ,
where ρ0 is the energy density entering the homogeneous Einstein equations from which
the homogeneous scale factor a(t) is computed. Inserting this expansion into Eq. (4) and
eliminating the pressure in favor of the energy density ρ ∝ H 2, we find a differential equation
of the generic form G(cid:16) ψ, ψ(cid:17) = 0, which does not contain any term proportional to ψ.
Therefore, ψ = ψ(xi) is a solution of Einstein equations at any order in perturbation theory
and ψ = 0 holds on super-Hubble scales. From this generic argument we conclude that there
is no influence of very long wavelength modes on the physical local expansion rate of the
Universe if adiabaticity holds.
In the adiabatic case, the influence of infrared modes is not locally measurable. There is
a simple explanation of this result. When adiabaticity holds, the pressure is a well defined
function of the energy density. This means that the Hubble rate is only a function of the
unique available physical clock, the energy density. Indeed, suppose that the local expansion
3
rate is H(t, xi) = H(ρ(t, xi), t). This leads to
∂H
∂t !xi
∂t!xi ∂H
∂ρ !t
∂t!xi ∂H
+ ∂t
∂t !ρ
= ∂ρ
= −4πG(ρ + P ) + ∂H
∂t !ρ
,
(5)
where we have made use of the equation 3H = − ρ/(ρ + P ). Comparing Eq. (5) with
H = −4πG(ρ + P ), we see that (∂H/∂t)ρ = 0, and hence H(t, xi) = H(ρ(t, xi)). The
dependence of the local expansion rate of the Universe on the clock time takes the same
form as in the unperturbed Universe when evaluated at a fixed value of the only clock
available, the energy density ρ. Infrared modes do not have any locally measurable effect
on the expansion rate of the Universe. This result applies, in particular, during inflation if
the energy density is dominated by a single inflaton field since in such a case adiabaticity
is guaranteed. One implication of this result is that long-wavelength modes may not be
responsible for stopping inflation.
This conclusion does not hold when adiabaticity is not attained. Indeed, the right-hand
side of Eq. (5) would contain an extra factor
∂t !xi ∂H
∂P !t
∂P
,
(6)
which makes it impossible for the local expansion rate to be a unique function of the energy
density. The same conclusion was reached using a second-order approach [4, 6, 7].
MB have very recently argued that one can make use of the fluctuations of a scalar field
ϕ and so long as these fluctuations are associated with an isocurvature mode, the effect on
local observables should be measurable, so that at late times the backreaction could mimic
dark energy. This conclusion is reached by inspecting the second-order effective energy
momentum tensor. Their approach is not fully consistent as the metric fluctuations and the
scalar field fluctuations are not properly expanded at second-order to include intrinsically
second-order quantities. However, for the sake of comparison, we will perform the same
approximation.
Following Ref. [1], let us consider a scalar field ϕ with potential V (ϕ) = λϕ4 . As
it was shown in Ref. [9], the equation of state of the zero mode ϕ0 is that of radiation,
implying ϕ0 ∝ a−1, a being the background scale factor. The scalar field may be split
as ϕ = ϕ0 + δϕ and one may consider the long-wavelength limit of the spatially averaged
4
energy density of the scalar field. MB argue that it is dominated at large times by a piece
δρϕ = 1
2V ′′hδϕ2i = −δPϕ. It dominates over the matter contribution, thus leading to an
effective cosmological constant behavior. The way this conclusion is reached is by making
use of Einstein equations which relate the scalar fluctuations to the metric fluctuations
( H + 3H 2)Φ = −4πG V ′δϕ +
δρm
2 ! ,
(7)
where Φ is the metric (scalar) perturbation in the longitudinal gauge, and dropping the piece
proportional to δρm. Since during a matter-dominated epoch H 2 ∼ a−3 ∼ V ′ and assuming
that the gravitational potential Φ is constant, MB conclude that δϕ does not evolve with
time, implying that δρϕ ∼ −δPϕ ∼ hΦ2i/a2. This term will eventually dominate over matter
thus leading to acceleration.
This result is, however suspicious because MB employ the "adiabatic" part of the fluctu-
ation of the scalar field, that is the part of the fluctuation proportional to the gravitational
potential. Let us work in the longitudinal gauge with metric ds2 = −(1 + 2Φ)dt2 + a2(t)(1 −
2Φ)δijdxidxj. As we stressed above, one needs a purely isocurvature component to hope
that super-Hubble modes have any influence on cosmological observables. The first-order
Klein-Gordon equation for the fluctuation of the scalar field reads
δ ϕ + 3Hδ ϕ + V ′′δϕ = 4 ϕ0 Φ − 2V ′ϕ0Φ .
The exact solution of this equation is (we assume a constant Φ as in Ref. [1])
δϕ = q
δϕ∗
ϕ0∗
ϕ0 + t ϕ0Φ ,
(8)
(9)
where the label "∗" indicates that the initial conditions have been fixed at some time t∗.
The first piece represents the pure isocurvature component of the fluctuation, while the
piece proportional to the gravitational potential is the adiabatic component. Its origin can
be understood via the separate Universe approach: when the fluctuation starts having some
nontrivial evolution after inflation, initial conditions are different in different Hubble patches
if adiabatic perturbations set by the gravitational potential Φ are present. The parameter
q ∼ 1 depends upon the polynomial form of the potential. From Eq. (9) we see that
δϕ ∼ ϕ0 ∼ a−1 throughout the all evolution of the Universe. This means that δρϕ ∼ a−4
during all epochs and there is no way the long-wavelength fluctuations of the scalar field
can dominate over nonrelativistic matter.
5
The reason why MB reach a different conclusion is that they solve Eq. (7) to infer the time
behavior of the scalar fluctuation. However, the piece proportional to δρm is dropped. This is
somewhat inconsistent since there are extra components in that equation to account for the
other fluids (and indeed more than one fluid is necessary to have isocurvature perturbation
without which no super-Hubble influence on observables is possible). On the contrary,
solving directly the Klein-Gordon equation shows that super-Hubble fluctuations behave
like a relativistic gas as the zero mode.
To understand this result in another way, we consider a nonlinear generalization of the
metric in the longitudinal gauge, ds2 = −e2Φdt2 + e−2Ψa2(t)δijdxidxj. Again, following
Ref. [1], we take the two gravitational potentials equal, Φ = Ψ (even though this not fully
consistent), and constant in time (there is a mild logarithmic dependence upon time at
second-order, which is, however, not relevant in the considerations of Ref.
[1]). As long
as we are interested in long-wavelength perturbations, we can recast this perturbed metric
in a FRW form by putting dt = eΦdt and a = e−Φa. The Klein-Gordon equation for the
quantum scalar field containing the zero mode and the super-Hubble perturbations reads
ϕ′′ + 3Hϕ′ +
dV
dϕ
= 0,
(10)
where the primes indicate now differentiation with respect to t and H = (da/dt) =
e−2Φ(H − Φ). Eq. (10) can then be rewritten as
ϕ + 3H ϕ + e2Φ dV
dϕ
= 4 ϕ Φ ,
(11)
which expanded to first-order gives exactly the perturbed Klein-Gordon equation, Eq. (8).
For a constant gravitational potential, one concludes that the scalar field (including the
long-wavelength modes) behaves exactly like the zero mode (upon rescaling: λ → e2Φλ).
Therefore, the energy density associated with the long-wavelength "adiabatic" scalar per-
turbations scales like a relativistic gas (as the zero mode) and may never dominate over
matter. Since super-Hubble perturbations may have an effect on observables only if they
are associated with isocurvature perturbations, we conclude that the effect, if any, must
depend upon the isocurvature component of δϕ, i.e., on δϕ∗/ϕ0∗. Indeed, it is easy to see
that the isocurvature perturbation Smϕ = 3 (ζm − ζϕ), where the ζm and ζϕ are the comov-
ing curvature perturbations for matter and the scalar field respectively, is proportional to
δϕ∗/ϕ0∗. However, isocurvature perturbations induce a sizable effect only when the fluids
6
have comparable energy densities, ζ ∼ H ( ρϕ ρm/ ρ2) Smϕ, and it is not clear to us if it suffices
to have a super-Hubble isocurvature component to mimic dark energy.
Acknowledgments
We thank D. H. Chung for useful correspondence. A.R. is on leave of absence from
INFN, Padova (Italy). E.W.K. is supported in part by NASA grant NAG5-10842 and by
the Department of Energy. S.M. acknowledges partial financial support by INAF.
[1] P. Martineau and R. Brandenberger, arXiv:astro-ph/0510523.
[2] V. F. Mukhanov, L. R. W. Abramo and R. H. Brandenberger, Phys. Rev. Lett. 78, 1624 (1997).
[3] W. Unruh, astro-ph/9802323.
[4] G. Geshnizjani and R. H. Brandenberger, Phys. Rev. D 66, 123507 (2002).
[5] F. Finelli, G. Marozzi, G. P. Vacca and G. Venturi, Phys. Rev. D 69, 123508 (2004).
[6] G. Geshnizjani and R. Brandenberger, JCAP 0504, 006 (2005).
[7] R. H. Brandenberger and C. S. Lam, hep-th/0407048.
[8] E. W. Kolb, S. Matarrese, A. Notari and A. Riotto, Mod. Phys. Lett. A 20, 2705
(2005) (e-print arXiv:astro-ph/0410541) See also, E. W. Kolb, S. Matarrese and A. Riotto,
arXiv:astro-ph/0506534, section IIIA.
[9] M. S. Turner, Phys. Rev. D 28, 1243 (1983); Y. Shtanov, J. H. Traschen and R. Brandenberger,
Phys. Rev. D 51, 5438 (1995).
7
|
astro-ph/0010338 | 3 | 0010 | 2001-01-18T14:31:42 | Upper Limits On Periodic, Pulsed Radio Emission from the X-Ray Point Source in Cassiopeia A | [
"astro-ph"
] | The Chandra X-ray Observatory recently discovered an X-ray point source near the center of Cassiopeia A, the youngest known Galactic supernova remnant. We have conducted a sensitive search for radio pulsations from this source with the Very Large Array, taking advantage of the high angular resolution of the array to resolve out the emission from the remnant itself. No convincing signatures of a dispersed, periodic source or of isolated dispersed pulses were found, whether for an isolated or a binary source. We derive upper limits of 30 and 1.3 mJy at 327 and 1435 MHz for the phase-averaged pulsed flux density from this source. The corresponding luminosity limits are lower than those for any pulsar with age less than 10^4 years. The sensitivities of our search to single pulses were 25 and 1.0 Jy at 327 and 1435 MHz. For comparison, the Crab pulsar emits roughly 80 pulses per minute with flux densities greater than 100 Jy at 327 MHz and 8 pulses per minute with flux densities greater than 50 Jy at 1435 MHz. These limits are consistent with the suggestion that the X-ray point source in Cas A adds to the growing number of neutron stars which are not radio pulsars. | astro-ph | astro-ph | Draft version October 26, 2018
Preprint typeset using LATEX style emulateapj v. 04/03/99
1
0
0
2
n
a
J
8
1
3
v
8
3
3
0
1
0
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
UPPER LIMITS ON PERIODIC, PULSED RADIO EMISSION FROM THE X-RAY POINT
SOURCE IN CASSIOPEIA A
M. A. McLaughlin1, J. M. Cordes1, A. A. Deshpande2
,
3, B. M. Gaensler4
,
5,
T. H. Hankins2, V. M. Kaspi6
,
7, & J. S. Kern2
Draft version October 26, 2018
ABSTRACT
The Chandra X-ray Observatory recently discovered an X-ray point source near the center of Cassiopeia
A, the youngest known Galactic supernova remnant. We have conducted a sensitive search for radio
pulsations from this source with the Very Large Array, taking advantage of the high angular resolution
of the array to resolve out the emission from the remnant itself. No convincing signatures of a dispersed,
periodic source or of isolated dispersed pulses were found, whether for an isolated or a binary source.
We derive upper limits of 30 and 1.3 mJy at 327 and 1435 MHz for its phase-averaged pulsed flux
density. The corresponding luminosity limits are lower than those for any known radio pulsar with age
less than 104 years. Our search sensitivities to single pulses were 25 and 1.0 Jy at 327 and 1435 MHz.
For comparison, the Crab pulsar emits roughly 80 pulses per minute with flux densities greater than 100
Jy at 327 MHz and 8 pulses per minute with flux densities greater than 50 Jy at 1435 MHz. These limits
suggest that Cas A belongs to the growing population of young neutron stars that are radio quiet
1.
INTRODUCTION
Cassiopeia A, with an age of only 320 years,
is the
youngest known Galactic supernova remnant. The super-
nova that gave rise to this remnant was probably discov-
ered in 1680, when Flamsteed (1725) observed a 6th mag-
nitude star within 13′ of Cas A's location (Ashworth 1980).
Because its abundances of heavy elements are consistent
with those expected from the explosion of a massive star
(Hughes et al. 2000), Cas A is thought to be the remnant
of a Type II supernova. Given the stellar initial mass func-
tion, a large fraction of Type II supernovae are expected to
leave behind neutron stars. While many searches at vari-
ous wavelengths for the compact remnant in Cas A have
been undertaken, only upper limits had been established
until the recent Chandra discovery of an X-ray point source
(hereafter XPS) near the center of the remnant (Tanan-
baum 1999). This XPS was later confirmed in archival
ROSAT (Aschenbach 1999) and Einstein (Pavlov & Za-
vlin 1999) images. The XPS is within 5′′ of the expansion
center of the remnant (van den Bergh & Kamper 1983).
For a remnant distance of 3.4+0.3
−0.1 kpc, calculated through
radial velocity, proper motion, and age data (Reed et al.
1995), this offset corresponds to a maximum transverse ve-
locity of roughly 200 km s−1, a typical velocity for young
pulsars (Cordes & Chernoff 1998).
Pavlov et al. (2000) find that the X-ray spectrum of the
XPS is fit well by a power law with index 2.6 − 4.1, with
luminosity (2−60)×1034 ergs s−1. Similarly, Chakrabarty
et al. (2001) estimate a power law index of 2.8 − 3.6 and
luminosity of (2 − 16) × 1034 ergs s−1. These indices are
much steeper than the X-ray spectra of Active Galactic
Nuclei (AGN). Furthermore, given the measured surface
density of AGN, the probability of finding an AGN that
close to the center of the SNR is negligible. For these
reasons, Pavlov et al.
(2000) conclude that the XPS is
almost certainly associated with the Cas A remnant. The
fitted power law indices for the XPS are also much steeper
than those measured for the 6 young X-ray pulsars hav-
ing ages less than 104 years, whose spectral indices range
from 1.1 − 1.7 (Chakrabarty et al. 2001).
In addition,
the estimated XPS luminosities are only marginally consis-
tent with measured young pulsar X-ray luminosities, which
range from 3 × 1035 to 4 × 1036 ergs s−1. It is also signif-
icant that, unlike 5 of the 6 young pulsars with ages less
than 104 years, Cas A shows no evidence for a synchrotron
nebula surrounding the point source.
While there is much evidence for and against the neu-
tron star nature of the XPS, the definitive proof would
be the detection of pulsations. No X-ray pulsations have
been found in data taken with the Chandra High Resolu-
tion Camera. However, a time-tagging problem caused the
timing accuracy to be degraded from 16 µs to ≈ 4 ms, al-
lowing searches for only broad pulses with periods greater
than 20 ms (Chakrabarty et al. 2001). The detection of
radio pulsations, however, would be just as conclusive. For
this reason, we have conducted a search for radio pulsa-
tions from the XPS using the Very Large Array (VLA) at
frequencies of 327 and 1435 MHz. There have been several
previous searches for a radio pulsar in Cas A (Davies &
Large 1970, Seiradakis & Graham 1980, Woan & Duffett-
Smith 1993, Lorimer et al. 1998). Our VLA search is
many times more sensitive, as having an accurate position
for the XPS allows us to resolve out the strong emission
1Astronomy Department, Cornell University, Ithaca, NY 14853
2Physics Department, New Mexico Institute of Mining and Technology, Socorro, NM 87801
3Raman Research Institute, Bangalore 560080, India
4Hubble Fellow
5Center for Space Research, Massachusetts Institute of Technology, Cambridge, MA 02139
6Physics Department, McGill University, Montreal, Canada H3A2T8
7on leave from Physics Department and Center for Space Research, Massachusetts Institute of Technology, Cambridge, MA 02139
1
2
from the remnant, which with 327- and 1435-MHz flux
densities of 6400 and 2000 Jy is one of the strongest radio
sources known.
2. DATA
We observed the XPS on three days (23 October
1999, 31 October 1999, and 28 November 1999) with
the phased-array VLA in B array, where the maximum
baseline is 11.4 km. The J2000 coordinates of the XPS
are RA=23h23m27s.94, DEC=+58o48′42′′, with positional
uncertainty of 1′′ (Tananbaum 1999). As the 327- and
1435-MHz synthesized beamwidths are approximately 20
and 4 arcseconds respectively, one beam at each frequency
was sufficient for this search. During each session, the XPS
was observed at center frequencies of 327 and 1435 MHz,
resulting in three pointings of length 20, 40, and 50 min-
utes at 327 MHz, and four pointings of length 40, 40, 45,
and 50 minutes at 1435 MHz. During each observing ses-
sion, we observed the known pulsar PSR B0355+54, which
has a period of 154 ms and a dispersion measure (DM) of
57 pc cm−3. A blank-sky pointing and the flux and phase
calibrator J2350+646 (with flux densities of 27 and 5 Jy
at 327 and 1435 MHz) were also observed.
We used the VLA's High Time Resolution Processor
(Moffett 1997) to record dual circular polarization data at
a sampling frequency of 9.22 kHz. Each scan was started
on a 10-second tick, tied to the VLA's hydrogen maser,
compared to Universal Time through the GPS. At 327
MHz, 14 × 0.125-MHz frequency channels covered a total
bandwidth of 1.75 MHz. These channels were spaced non-
contiguously to avoid known interference frequencies and
covered a total bandwidth range of 3 MHz. At 1435 MHz,
14×4-MHz channels covered a total contiguous bandwidth
of 48 MHz. (As only 50 MHz of bandwidth is available,
some channels were duplicated). The gain in each of the
28 channels (2 polarizations and 14 frequencies) was cal-
culated from observations of the calibrator source.
3. ANALYSIS
For each of the seven XPS data sets, we first summed
right and left circular polarizations and then dedispersed
the data over a range of trial DMs. The Taylor and Cordes
(1993) model predicts a maximum DM of approximately
75 pc cm−3 in the direction of the XPS. We dedispersed
the data with 60 trial DMs from 0 to 300 pc cm−3 to allow
for an underestimate or a contribution from Cas A. These
60 DMs were spaced nonuniformly, with greater spacing
at higher DMs, where the temporal smearing due to dis-
persion across an individual frequency channel is larger.
Each dedispersed time series was Fourier transformed with
32 Mpoint and/or 16 Mpoint Fast Fourier Transforms
(FFTs), depending on the length of the data set. Shorter
FFTs of 8 and 4 Mpoints were done to increase our sensi-
tivity to some binary pulsars. The resulting power spectra
were harmonically summed (up to a maximum of 16 har-
monics) to increase our sensitivity to narrow pulses (Lyne
& Smith 1998). All features in the resulting power spec-
tra with signal-to-noise above a threshold of seven were
saved as pulsar candidates. For each candidate, the corre-
sponding dedispersed time series was folded at the period
corresponding to the fundamental harmonic. These pro-
files were inspected and candidate lists from different days
and at different frequencies were compared. While many
periodic signals were detected, they were predominantly
at low DM, had noiselike pulse profiles, and did not ap-
pear on multiple epochs or at multiple frequencies, and
were therefore attributed to radio frequency interference
(RFI). Figure 1 shows the folded pulse profiles resulting
from applying the search algorithm to PSR B0355+54.
Fig. 1. -- Folded pulse profiles for PSR B0355+54, which has a
period of 156 ms and DM of 57 pc cm−3, at 327 MHz (left) and
1435 MHz (right). Note that these profiles are for 5 and 2 minutes
of data at 327 and 1435 MHz, whereas our longest XPS search scans
were 50 minutes at these same frequencies.
An independent Fourier analysis of the same data, car-
ried out at the New Mexico Institute of Mining and Tech-
nology, was also unsuccessful. This analysis was sensitive
to periods ≥10 ms and covered a smaller DM range, but
with a finer DM spacing. Three filter-bank channels with
high levels of RFI were rejected, decreasing the number of
spurious candidates significantly with only a 12% sensitiv-
ity loss. Power spectra were harmonically summed up to a
maximum of 32 harmonics. Final pulsar candidates were
evaluated by comparing average profiles from two halves of
the bandpass and two halves of profiles folded over a two
period span. These were compared with the corresponding
profiles for non-dedispersed data.
The Crab pulsar, currently the youngest known radio
pulsar, was initially detected through its 'giant' pulses,
not through its time-averaged pulsed emission (Staelin &
Reifenstein 1968). Giant pulses are individual pulses with
amplitudes much greater than the mean pulse amplitude
(Lundgren et al. 1995). As we might expect other young
pulsars to behave similarly to the Crab, we also searched
for aperiodic, dispersed pulses from the XPS. We recorded
all pulses with amplitudes above a signal-to-noise thresh-
old of 4 for each dedispersed time series, enchancing our
sensitivity to broadened pulses by repeating this thresh-
olding with different levels of time-series smoothing. Given
our 4σ threshold, we were sensitive to all single pulses with
flux densities greater than 25/√wms Jy, where wms is the
pulse width in milliseconds, at 327 MHz and 1.0/√wms
Jy at 1435 MHz. Our smoothing algorithm optimized the
search for pulse widths up to 15 ms.
(For comparison,
the widths of the Crab pulsar's giant pulses range from
roughly 0.2 to 0.4 ms.) We found no evidence for isolated
pulses from the XPS. The result of this analysis, along
with the result for PSR B0355+54, is shown in Figure 2.
For comparison, the Crab pulsar emits roughly 80 pulses
per minute with flux densities greater than 100 Jy at 327
MHz and 8 pulses per minute with flux densities greater
than 50 Jy at 1435 MHz.
3
Figure 3 shows the dependence of our search sensitiv-
ity on DM and period at the two observing frequencies,
accounting for the effects of scattering and dispersion on
the intrinsic duty cycle of the pulsar. Assuming values for
period and DM of 30 ms and 75 pc cm−3, we calculate
7σ upper limits of 30 and 1.3 mJy at 327 and 1435 MHz.
As shown in Figure 3, these limits will decrease for longer
period, smaller duty cycle, and lower DM. There are no
hardware limitations to our maximum detectable period,
determined solely by the length of the FFT.
While there is no infrared or optical evidence for a bi-
nary companion (van den Bergh & Pritchet 1986), accel-
eration effects from such a companion would cause pulse
smearing, decreasing our sensitivity. We estimate the
signal-to-noise degradation introduced for various orbital
systems by taking a Gaussian pulse and applying orbital
smearing by iteratively calculating emission times across
the orbit. Smearing times due to time constant, DM,
and scattering are calculated and applied to the orbital-
smeared pulse. We calculate the best harmonic sum for
the resultant waveform and compare this to the best har-
monic sum for a waveform with no orbital smearing to
calculate the loss in signal-to-noise. For a 30 ms period
and duty cycle of 0.2, a 1 solar mass companion in an 8
hour circular orbit with an inclination angle (i.e. angle
between the orbital plane and our line of sight) of 30◦, our
sensitivity decreases by roughly a factor of 2, averaging
over all orbital phases, for a 16 Mpoint FFT. In the case
of our 8 and 4 Mpoint FFTs, these same binary param-
eters would decrease our search sensitivity by factors of
2.4 and 2.8 over the original 16 Mpoint sensitivity to an
unaccelerated pulsar. Although the nominal sensitivities
are less for these shorter FFT lengths, the shorter FFTs
would allow us to detect pulsars in fast binaries where the
binary smearing during a longer FFT is comparable to or
greater than the pulse period. The sensitivity would be
further decreased for smaller spin periods, faster orbits,
and more massive companions. If the orbit was eccentric,
the net smearing, averaged over all orbital phases, would
be less, and the resultant sensitivity would be higher. Ac-
cretion from a binary companion could completely quench
the radio emission.
Fig. 3. -- Contours of pulsed search 7σ sensitivity vs. P and DM
for searches at both 327 and 1435 MHz. The thickest (rightmost)
line corresponds to a sensitivity of 16 mJy at 327 MHz and of 0.8
mJy at 1435 MHz. Minimum detectable flux densities increase by
a factor of 2 with eac h subsequently thinner line. The contours are
not smooth because of the finite spacing of our trial DMs.
Interstellar scintillation modulations of the flux of
the XPS are mostly negligible during our observations.
Fig. 2. -- The number of isolated, dispersed pulses at 327 MHz
above a 4σ threshold is plotted vs. DM for a 5-minute observation
of PSR B0355+54 (left) and a 50-minute observation of the XPS
(right). While a peak at the 57 pc cm−3 DM of PSR B0355+54 is
obvious, there is no evidence for such an excess from the XPS. The
peak at low DM in the right plot is consistent with interference, and
is similar to the number of counts observed in the consecutive blank
sky pointing. I n the absence of radio frequency interference and/or
a pulsar, we would expect a flat distribution.
4. RESULTS
To calculate an upper limit for pulsed radio emission
from the XPS, we must account for the excess system tem-
perature due to the emission from Cas A itself. The Cas A
remnant is 5′ in size (Green 2000). Hence, the ratios of
our beam area to remnant area are 0.003 and 0.0002 at
327 and 1435 MHz. Given Cas A's measured flux density
and spectral index (Green 2000), we calculate remnant
flux densities of 6432 and 2060 Janskys at 327 and 1435
MHz, contributing to an increase in system temperature
of a modest 40 K at 327 MHz and a mere 1 K at 1435 MHz
for the phased array. However, while the phased array re-
solves out emission from the nebula, each single dish will
see the whole remnant, leading to an increase in system
temperature of 475 K at 327 MHz and 230 K at 1435 MHz.
By comparison, a single dish with the same collecting area
as the VLA would suffer increases of system temperature
of 13000 and 6200 K at 327 and 1435 MHz.
We calculate the minimum detectable flux density of our
search using the expression (Cordes & Chernoff 1997)
Smin =
ηTsys
GpNpol∆ν∆tNFFT
√Nh
PNh
l=1 Rl
,
(1)
where η is the signal-to-noise threshold used, Tsys is the
system temperature, G is the telescope gain, Npol is the
number of polarization channels, ∆ν is the total band-
width, ∆t is the sample interval, NFFT is the number of
points in the FFT, Nh is the optimal number of harmon-
ics with which a pulsar can be detected, and Rl ≤ 1 is the
ratio of the lth harmonic to the fundamental. Nh will de-
pend on the pulsar's duty cycle, or ratio of pulse width to
period, with narrower pulses producing more harmonics.
For this calculation, we take the intrinsic duty cycle to be
the minimum of 0.3 and 0.03(Pms/1000.)−0.5, where Pms is
the pulse period in milliseconds (Biggs 1990). Pulses may
then be additionally broadened by propagation effects, in-
cluding scattering and dispersion. The Taylor & Cordes
(1993) model predicts 1.7 ms and 3 µs of pulse broadening
due to scattering in the direction of the XPS and for a
distance of 3.4 kpc at 327 and 1435 MHz, respectively.
XPS, gamma-ray searches with future, more sensitive in-
struments, could be successful. Further measurements of
the spectrum and time variability of the XPS will also help
determine which of the above scenarios is the case.
This search adds to the growing body of evidence that
there are manifestations of young neutron stars other
than the standard Crab-like radio pulsar. Although it
is widely accepted that radio pulsars are born in super-
novae, fewer than 10% of searches targeted at supernova
remnants have found young (age < 25 kyr) pulsars (Kaspi
et a. 1996, Gorham et al. 1996, Lorimer et al. 1998).
Some of this may be due to scattering at low frequencies
in the Galactic plane and the decreased sensitivity due
to the bright remnants, but it also hints at the unrec-
ognized diversity of the neutron star population. While
some neutron stars are detected as radio pulsars, others
may be born as anomalous X-ray pulsars, soft gamma-ray
repeaters, quiescent neutron stars, or other classes of ob-
jects not yet discovered (Kaspi 2000, Gotthelf & Vasisht
2000). The pathways to and percentages of these different
classes of objects are not yet clear. However, future Chan-
dra detections of compact X-ray sources, and large-scale
radio pulsar surveys, like the Parkes Multibeam Survey
(Lyne et al. 2000), will help to answer these questions.
We wish to thank Miller Goss for expediting our ob-
servations and Duncan Lorimer for helpful comments on
the manuscript. We thank Larry Rudnick and Barron Ko-
ralesky for kindly providing their radio images of Cas A
for our analysis. This work was partially supported by
NSF grant AST-9618408. The work was also supported
by the National Astronomy and Ionosphere Center, which
is operated by Cornell University under cooperative agree-
ment with the National Science Foundation (NSF). The
National Radio Astronomy Observatory is a facility of
the National Science Foundation operated under cooper-
ative agreement by Associated Universities, Inc. BMG
acknowledges the support of NASA through Hubble Fel-
lowship grant HST-HF-01107.01-A awarded by the Space
Telescope Science Institute, which is operated by the Asso-
ciation of Universities for Research in Astronomy, Inc., for
NASA under contract NAS 5 -- 26555. VMK acknowledges
support from an Alfred P. Sloan research fellowship, from
an NSF CAREER award (AST-9875897) and an NSERC
Research Grant (RGPIN228738-00).
4
The predicted scintillation bandwidth and timescale (see
Cordes & Rickett 1998), assuming a uniform medium and
a velocity of 50 km s−1 are approximately 0.1 kHz and
30 seconds at 327 MHz and 0.6 MHz and 150 seconds at
1435 MHz, in both cases too small to significantly affect
our quoted upper limits. If this velocity is higher than 50
km s−1, as a young pulsar's velocity is likely to be (Cordes
& Chernoff 1998), the predicted scintillation timescale will
be smaller and will therefore be even less of a factor in our
analysis.
We have also used VLA images of Cas A to search for a
radio continuum point-source at the position of the XPS.
At 1.3 GHz and 4.4 GHz, we find 5σ upper limits on the
flux density from the XPS of 40 mJy and 6 mJy respec-
tively. While not as constraining as those derived from our
VLA and Arecibo pulsed searches described above, these
continuum upper limits are completely independent of the
period, dispersion and scattering of any radio pulsations
from this source.
5. CONCLUSIONS
We have calculated upper limits of 30 and 1.3 mJy at
327 and 1435 MHz for the phase-averaged pulsed flux den-
sity from the XPS in Cas A. We did not detect any single,
dispersed pulses above our sensitivity of 25 and 1.0 Jy (as-
suming a pulse width of 1 ms) at 327 and 1435 MHz. We
now compare our upper limit to those from previous Cas A
searches. Davies & Large (1970) carried out an unsuccess-
ful 408-MHz search for single pulses from Cas A. Seiradakis
& Graham (1980) quoted a 1420-MHz upper limit of 1.6
mJy for pulsed emission from Cas A. However, their limit
does not account for the additional system temperature
contributed by the nebula, significant for their beamwidth
of 9′. More recently, Woan & Duffett-Smith (1993) pub-
lished an upper limit of 80 mJy at 408 MHz and Lorimer et
al. (1998) published an upper limit of 46 mJy at 606 MHz.
For a typical pulsar spectrum, our 1435-MHz upper limit
is an order of magnitude better than these previously pub-
lished limits, due to the known position of the XPS and
the power of the VLA to resolve out the excess emission
from the nebula.
Our upper limit of 1.3 mJy at 1435 MHz translates to a
luminosity of 15 mJy kpc2 for a neutron star at a distance
of 3.4 kpc. This is lower than the mean 1400-MHz lumi-
nosity of 30 mJy kpc2 estimated for young pulsars (Frail
& Moffett 1993) and is lower than the luminosity of any
known radio pulsar with characteristic age less than 104
years (see Figures 1 and 2 of Brazier & Johnston 1999).
This suggests that the XPS in Cas A is either not a pul-
sar, is a radio-quiet pulsar, or is a radio pulsar beamed
away from us. While evidence suggests that pulsar beams
are wider at short periods (see e.g. Biggs (1990)) and that
they are therefore likely to intersect the Earth's line of
sight, individual pulsars show mixed properties. For in-
stance, the Crab pulsar shows narrow radio components
that may be part of a large beam, but the profile of another
young pulsar in the Large Magellanic Cloud, PSR B0540-
69, is very broad (Manchester et al. 1993). Furthermore,
it is difficult to disentangle true beam width from viewing
angle for these pulsars. It is certainly possible that, like
the Geminga pulsar, the XPS shows no radio pulses but
is an X-ray, and perhaps gamma-ray, pulsar. Although
there is no EGRET point source near the position of the
REFERENCES
5
Aschenback, B., 1999, IAU Circ. 7249
Ashworth, W. B., 1980, J. Hist. Astron., 11, 1
Biggs, J. D. 1990, MNRAS, 245, 514
Brazier, K. T. S. & Johnston, S. 1999, MNRAS, 305, 671
Camilo, F. et al. ASP Conf. Ser. 202: IAU Colloq. 177: Pulsar
Astronomy - 2000 and Beyond, 3
Cordes, J. M., & Chernoff, D. F., 1997, ApJ, 482, 971
Cordes, J. M. & Chernoff, D. F. 1998, ApJ, 505, 31
Cordes, J. M., & Rickett, B. J. 1998, ApJ, 507, 846
Davies, J. G., & Large, M. I., 1970, MNRAS, 149, 301
Chakrabarty, D., Pivovaroff, M. J., Hernquist, L. E., Heyl, J. S., &
Narayan, R., 2001, ApJ, 546, in press
Flamsteed, J. 1725, London : H. Meere; in folio; DCC.f.9, DCC.f.10,
DCC.f.11
Frail, D. A, & Moffett, D. A., 1993, ApJ, 408, 67
Gorham, P., Ray, P., Anderson, S., Kulkarni, S., Prince, T. 1996,
ApJ, 458, 257
Gotthelf, E. V., & Vasisht, G., 2000, ASP Conf. Ser. 202: IAU Colloq.
177: Pulsar Astronomy - 2000 and Beyond, 699
Green D. A., 2000, 'A Catalogue of Galactic Supernova Remnants
(2000 August version)', Mullard Radio Astronomy Observatory,
Cavendish
Laboratory,
Cambridge, United Kingdom (available on the World-Wide-Web
at "http://www.mrao.cam.ac.uk/surveys/snrs/")
Hughes, J. P., Rakowski, C. E., Burrows, D. N., & Slane, P. O.
2000,ApJ, 528, L109
Kaspi, V., Manchester, R., Johnston, S., Lyne, A., D'Amico, N. 1996,
AJ, 111, 2028
Kaspi, V. M. 2000, ASP Conf. Ser. 202: IAU Colloq. 177: Pulsar
Astronomy - 2000 and Beyond, 485
Lorimer, D. R, Lyne, A. G., & Camilo, F., 1998, A&A, 331, 1002
Lundgren, S. C. et al. 1995, ApJ, 453, 433
Lyne, A. G., & Smith, F. G., 1998, Pulsar Astronomy, 2nd ed.,
Cambridge University Press, Cambridge
Lyne, A. G. et al. 2000, MNRAS, 312, 69
Manchester, R. N., Mar, D. P., Lyne, A. G., Kaspi, V. M. and
Johnston, S. 1993, 403, L29
Moffett, D. A. 1997, PhD. Thesis
Pavlov, G. G. & Zavlin, V. E., 1999, IAU Circ. 7270
Reed, J. E., Hester, J. J., Fabian, A. C., & Winkler, P. F. 1995, ApJ,
440, 706
Seiradakis, J. H. & Graham, D. A., 1980, A&A, 85, 353
Staelin, D. H. & Reifenstein, E. C., 1968, Science, 162, 1481
Tananbaum, H. 1999, IAU Circ. 7246
Taylor, J. H., Manchester, R. N., & Lyne, A. G. 1993, ApJS, 88, 529
Taylor, J. H. & Cordes, J. M. 1993,ApJ, 411, 674
van den Bergh, S., & Kamper, K. W. 1983, ApJ, 268, 129
van den Bergh, S. & Pritchet, C. J. 1986, ApJ, 307, 723
Woan, G. & Duffett-Smith, P. J., 1993, MNRAS, 260, 693
|
astro-ph/0611374 | 1 | 0611 | 2006-11-13T05:01:19 | Spectropolarimetry of the 3.4 micron absorption feature in NGC 1068 | [
"astro-ph"
] | In order to test the silicate-core/organic-mantle model of galactic interstellar dust, we have performed spectropolarimetry of the 3.4 micron C-H bond stretch that is characteristic of aliphatic hydrocarbons, using the nucleus of the Seyfert 2 galaxy, NGC 1068, as a bright, dusty background source. Polarization calculations show that, if the grains in NGC 1068 had the properties assigned by the core-mantle model to dust in the galactic diffuse ISM, they would cause a detectable rise in polarization over the 3.4 micron feature. No such increase is observed. We discuss modifications to the basic core-mantle model, such as changes in grain size or the existence of additional non-hydrocarbon aligned grain populations, which could better fit the observational evidence. However, we emphasize that the absence of polarization over the 3.4 micron band in NGC 1068 - and, indeed, in every line of sight examined to date - can be readily explained by a population of small, unaligned carbonaceous grains with no physical connection to the silicates. | astro-ph | astro-ph |
Spectropolarimetry of the 3.4 µm absorption feature in NGC 1068
Gemini Observatory Northern Operations Centre, 670 N. A'ohoku Place, Hilo, HI 96720, USA
R. E. Mason
[email protected]
G. S. Wright
UK Astronomy Technology center, Royal Observatory Edinburgh, Blackford Hill, Edinburgh EH9 3HJ, UK
[email protected]
A. Adamson
Joint Astronomy Centre, 660 North A'ohoku Place, Hilo, HI 96720, USA
[email protected]
and
Y. Pendleton
NASA Ames Research Center, Mail Stop 245-3, Moffet Field CA 94035, USA
[email protected]
ABSTRACT
In order to test the silicate-core/organic-mantle model of galactic interstellar dust, we have
performed spectropolarimetry of the 3.4 µm C -- H bond stretch that is characteristic of aliphatic
hydrocarbons, using the nucleus of the Seyfert 2 galaxy, NGC1068, as a bright, dusty background
source. Polarization calculations show that, if the grains in NGC1068 had the properties assigned
by the core-mantle model to dust in the galactic diffuse ISM, they would cause a detectable rise
in polarization over the 3.4 µm feature. No such increase is observed. We discuss modifications
to the basic core-mantle model, such as changes in grain size or the existence of additional
non-hydrocarbon aligned grain populations, which could better fit the observational evidence.
However, we emphasize that the absence of polarization over the 3.4 µm band in NGC1068 --
and, indeed, in every line of sight examined to date -- can be readily explained by a population
of small, unaligned carbonaceous grains with no physical connection to the silicates.
Subject headings: galaxies: individual: NGC1068 -- galaxies: ISM -- galaxies: Seyfert -- dust, extinction
-- polarization
1.
Introduction
In our galaxy and no doubt others, stars and
planets form from components available in the lo-
cal interstellar medium. The organic compounds
observed in interstellar space may therefore be the
first step towards the complex materials that help
make planets habitable. A greater understanding
of the origin and evolution of the organic materi-
als in the ISM, both in our galaxy and in others,
is thus of great interest.
Particularly puzzling are the aliphatic hy-
drocarbons, whose presence in the diffuse ISM
(DISM) of the Milky Way has been established
through observations of C-H bond stretches near
1
3.4 µm, seen in absorption throughout the DISM
of our own galaxy and others (Sandford et al.
1991; Pendleton et al. 1994; Sandford et al. 1995;
Imanishi 2000; Rawlings et al. 2003; Risaliti et al.
2003; Dartois et al. 2004; Mason et al. 2004). Two
very different models exist to explain the pro-
duction of this material: photoprocessing of the
icy dust grain mantles found in dense molecular
clouds (e.g. Li & Greenberg 1997), and hydro-
genation in the DISM of small, amorphous carbon
particles of the kind produced in outflows from
evolved stars (e.g. Mennella et al. 2002). As the
former process would result in the 3.4 µm band of
DISM hydrocarbons arising in organic mantles on
silicate grain cores, it is commonly known as the
"core-mantle" dust model.
Both of these models are to some extent con-
sistent with the observational data. For instance,
laboratory experiments mimicking both dust for-
mation pathways have succeeded in producing ma-
terials with 3.4 µm bands which are a good match
to the astronomical
feature (Greenberg et al.
1995; Mennella et al. 1999, 2002), although the
critically diagnostic 5-10 µm region reveals a
much closer match between observations and hy-
drogenated amorphous carbon (HAC) materi-
als (Chiar et al. 2000; Pendleton & Allamandola
2002). The presence of the 3.4 µm band in a pro-
toplanetary nebula (Chiar et al. 1998) shows that
hydrocarbon solids can be produced in circumstel-
lar regions, while the attribution of the 6 µm ex-
cess absorption in protostars to organic refractory
material (Gibb & Whittet 2002) would,
if con-
firmed, indicate that ice processing can also pro-
duce some fraction of the hydrocarbons. The or-
ganic OCN− band at 4.62 µm provides evidence of
ice processing in such environments (Demyk et al.
1998; Pendleton et al. 1999).
Spectropolarimetry provides opportunities for
clearly discriminating between these scenarios. In
the core-mantle model, the silicate core/organic
mantle grains are responsible for the bulk of the
visual and IR extinction and all of the polariza-
tion in the DISM (Li & Greenberg 1997). As the
polarization is caused by dichroic absorption, we
would therefore expect to observe polarization of
the continuum, with an increase in polarization
over absorption features associated with the grains
(e.g. Hough et al. 1988; Pendleton et al. 1990;
Smith et al. 2000). Specifically, increases in polar-
2
ization must occur simultaneously over the 3.4 µm
"mantle" feature, and the 9.7 µm absorption char-
acteristic of the silicate core, with roughly com-
parable efficiency (Li & Greenberg 2002). This
was first investigated by Adamson et al. (1999),
who compared the polarization over the 3.4 µm
feature towards the Galactic center source, IRS 7,
with that of the 9.7 µm band towards the nearby
Galactic center source, IRS 3. The increase in po-
larization over the 3.4 µm absorption was found
to be .0.08%, much less than the 0.4% expected
on the basis of the silicate feature polarization, in
apparent conflict with the core-mantle model.
However, as pointed out by Li & Greenberg
(2002), the silicate feature polarization used in
that work, while arising toward an object at a
projected distance of about 0.25pc from that to-
wards which the 3.4 µm polarization was mea-
sured (Geballe et al. 1989), does not refer to ex-
actly the same line of sight, admitting the possi-
bility that the silicate feature in the diffuse ISM
towards IRS 7 may be less polarized than ex-
pected. Studies of additional lines of sight (at
lower signal-to-noise ratio and/or spectral resolu-
tion) while reaching similar conclusions to those of
Adamson et al. (1999) , suffer from similar limita-
tions (Nagata et al. 1994; Ishii et al. 2002). More
recently, Chiar et al. (2006) measured the polar-
ization of both the hydrocarbon and silicate fea-
tures towards the Quintuplet Cluster in the Galac-
tic Center, finding the hydrocarbons to be signif-
icantly less polarized than would be expected if
they existed as mantles on the polarized silicate
grains.
To provide another data point against which
the predictions of the core-mantle grain model can
be compared, and to extend this study to a new
and different environment, we have used the bright
nucleus of the Seyfert 2 galaxy NGC1068 as a
source against which to measure the polarization
of the 3.4 µm band and the continuum around it.
Rather than the 10 µm silicate feature, we have
used the degree of continuum polarization to cal-
culate the enhancement in polarization that would
be expected over the 3.4 µm feature from a screen
of elongated, coated silicate grains which polarize
by the selective absorption of one plane of the in-
cident light. In §2, we describe the line of sight to-
wards the nucleus of NGC1068 and its suitability
for this work. The spectropolarimetric observa-
tions and their treatment are outlined in §3. The
calculations that we have carried out are discussed
in §4, and the results presented in §5. This work
is then summarized in §6.
2. Dust and dichroic polarization in NGC 1068
As the archetypal Seyfert 2 galaxy, NGC 1068
(d=16 Mpc; H0 = 70km s−1 Mpc−1; 1′′ = 72
pc) has been the center of much attention. The
unified model of active galactic nuclei (AGN) im-
plies that the galaxy harbors a Seyfert 1 nu-
cleus, obscured from our point of view by a torus
of dust and molecular gas, and 10 µm interfer-
ometry (Jaffe et al. 2004) as well as the detec-
tion of broad emission lines in polarized light
(Antonucci & Miller 1985) have provided strong
evidence for the existence of such a torus. Co-
pious other mid-IR and X-ray data also show
that the nucleus of NGC 1068 is obscured by
large amounts of warm dust (e.g. Bock et al. 2000;
Matt et al. 2000; Tomono et al. 2001; Alloin et al.
2000; Mason et al. 2006). While the classification
of NGC 1068 as a type 2 object suggests that the
obscuring torus is oriented roughly edge-on to our
line of sight, the disk of the galaxy has quite a low
inclination (i ≈ 29 ◦; Garc´ıa-G´omez et al. 2002).
This means that the line of sight to the center of
NGC 1068 samples dust local to the active nu-
cleus, with a negligible contribution from dust in
the disk of the galaxy.
The 3.4 µm absorption band has been observed
in NGC 1068 (Bridger et al. 1994; Wright et al.
1996; Imanishi et al. 1997; Marco & Brooks 2003;
Mason et al. 2004), as has silicate absorption at
9.7 µm (Kleinmann et al. 1976; Roche et al. 1984;
Jaffe et al. 2004; Mason et al. 2006; Rhee & Larkin
2006). The 3.4 µm band has a very similar pro-
file to that observed in Galactic lines of sight;
although the band carrier exists in the central re-
gion of an active galaxy, it apparently undergoes
little extra processing compared to the Galactic
DISM (Dartois et al. 2004; Mason et al. 2004).
NGC 1068 has a nuclear dust spiral which ap-
pears physically connected to the galaxy disk
(Pogge & Martini 2002), so it is quite possible
that the hydrocarbons in the nucleus of NGC 1068
formed in the disk of the galaxy and then mi-
grated to the center. We therefore proceed on the
assumption that the hydrocarbon-containing dust
3
in the nucleus of NGC 1068 initially formed by one
of the mechanisms that has been suggested to be
responsible for Galactic aliphatic hydrocarbons.
The nucleus of NGC 1068 is known to be po-
larized from the UV to the mid-IR, and the mech-
anisms responsible for this polarization have been
the subject of intense study. Antonucci & Miller
(1985) were among the first to observe the UV
and visible polarization of NGC 1068, and they
interpreted the high, wavelength-independent po-
larization as being due to scattering off very small
particles, probably free electrons. Bailey et al.
(1988) later discussed the polarization properties
of NGC 1068 from the UV out to 10µm, and found
a twist in position angle from the optical into the
near-IR. They interpreted this as dust or electron
scattering being replaced by dichroic absorption,
the preferential absorption of one component of
the electric vector, as the most important polariz-
ing mechanism. Their data also show a change
of ∼ 70◦ in position angle between the L and
M bands, consistent with dichroic emission from
aligned dust grains becoming the dominant polar-
izing mechanism at longer wavelengths.
Further evidence for the importance of dichroic
absorption as a polarizing mechanism in the IR
comes from imaging polarimetry. After analyzing
the deviations from the centrosymmetric patterns
expected from scattering alone, Lumsden et al.
(1999) concluded that their data require another
mechanism of constant position angle contribut-
ing to the polarization. The effect of this mecha-
nism grows with increasing wavelength, as would
be expected if absorption were beginning to domi-
nate over scattering, and was estimated to account
for > 90% of the K band polarization. Lums-
den at el. were successful in fitting their observed
JHK polarized flux points with greybody emission
from hot (T ∼ 1200K) dust, reddened by a λ−1.75
extinction curve and scaled by a Serkowski law
appropriate for moderately extinguished Galactic
sources. Furthermore, Packham et al. (1997) in-
terpreted the different aperture dependence of the
J-, H- and K- band polarization as consistent with
a dichroic contribution in the H- and K-bands.
With higher-resolution HST imaging polarimetry,
Simpson et al. (2002) also found that the K-band
polarization has a contribution from dichroism,
but they were able to set tighter limits on the lo-
cation of the dichroic component, to within 1′′ of
the nucleus.
This agrees with the more detailed modeling of
Young et al. (1995), who find that neither electron
nor dust scattering can account for the near-IR
polarization of NGC 1068. However, if a dichroic
component is added, the near-IR polarized flux
spectrum can be reproduced. In this model there
is still some contribution from electron scattering
to the near-IR flux, but this decreases with wave-
length relative to the dichroic component. By the
K band, electron scattering contributes perhaps
10% of the total flux, with dust making up almost
all of the remainder. At L, the scattering contri-
bution in their model would be negligible.
Watanabe et al. (2003) are also successful in
modeling the near-IR polarization of NGC 1068
with polarization from aligned dust grains. In ad-
dition, they raise the possibility that the IR po-
larization could be caused by scattering off large
grains in the torus, and point out that neither
the 70 ◦change in position angle nor the absence
of a centrosymmetric scattering pattern at longer
wavelengths necessarily rules out scattering as the
polarizing mechanism. This has yet to be tested in
any detail, but the implications that it may have
for the conclusions of this study are discussed in
§5.
While the near-IR data point to dust absorption
being the dominant polarizing mechanism in the
L band, spectropolarimetry of the 9.7 µm silicate
feature does not show the pronounced polarization
excess that might be expected if emission or ab-
sorption from aligned silicates were causing the po-
larization (Aitken et al. 1984). This observation
was interpreted as evidence that the mid-IR polar-
ization arises in emission from non-silicate grains
or has a nonthermal origin. However, the lack of
polarization in the silicate feature does not neces-
sarily rule out the presence of aligned silicate-core
grains in the nucleus of NGC 1068. In radiative
transfer calculations of dichroic polarization from
silicate-containing grain mixtures in dusty disks,
Efstathiou et al. (1997) and Aitken et al. (2002)
find numerous configurations in which the polar-
ization over the feature is in fact quite flat, consis-
tent with the observations of Aitken et al. (1984).
Although it appears possible for radiative trans-
fer effects to suppress the silicate feature polariza-
tion, the shorter-wavelength 3.4 µm band should
be much less affected by such interplay between
absorption and emission. The ratio of the optical
depths of the 3.4 and 9.7µm bands in NGC 1068
supports this suggestion. In the five Galactic lines
of sight examined by Sandford et al. (1995) there
is a fair degree of correlation between the depths
of the two features (τ9.7/τ3.4 = 13 − 19, with the
higher values towards the Galactic Centre), but
in NGC 1068 τ9.7/τ3.4 ≈ 5, suggesting that the
3.4 µm band is indeed less affected by underly-
ing emission than is the silicate feature. This fur-
ther suggests that treating the dust producing the
3.4 µm band as a uniform absorbing screen (an
assumption implicit in the calculations outlined in
§ 4), while undoubtedly a major simplification of
the dust geometry and temperature structure in
this AGN, is still a useful approximation. Overall,
imaging polarimetry, spectral modeling and the
near-90 ◦ rotation in position angle all imply that
a model of the L-band polarization of NGC 1068
based on selective dust absorption is a reasonable
representation of the true situation.
3. Observations and Data Reduction
L-band spectropolarimetry of the nucleus of
NGC1068 was obtained on the nights of 2000
September 19 and 2000 October 6 using the IR-
POL2 spectropolarimetry module and CGS4 on
the 3.8m UK Infrared Telescope on Mauna Kea,
Hawaii. The 40 l/mm grating and 0.6′′-wide slit
were used, providing R=1360 at 3.4 µm. The
weather conditions were good during the second
night, but some thin cirrus was present on the
first1.
In order to obtain Stokes q and u parameter
spectra a frame was taken with IRPOL2's half-
wave plate at each of four positions: 0◦, 45◦, 22.5◦,
and 67.5◦. This cycle was repeated 64 times in all.
At each of these waveplate angles, ordinary and
extraordinary beams were extracted from the sky-
subtracted frames, stacked together, and the final,
total spectra combined using the "ratio" method
(see e.g. Tinbergen 1996) to produce Stokes q and
u parameter spectra. This method has the advan-
tage of minimizing the effects of variations in sky
transmission during the observations.
Prior to calculation of the polarization, further
1Spectropolarimetry of the 9.7 µm silicate feature was later
also attempted, but poor weather meant that none of the
data obtained were useful.
4
treatment of the raw q and u spectra was nec-
essary. Firstly, data points between 3.3 and 3.4
µm (observed) were rejected. This part of the
spectrum contains deep absorptions from the hy-
drocarbon cement in IRPOL2's Wollaston prism
(at 3.35-3.4 µm) and from atmospheric methane
(3.31-3.33 µm), and these lines do not ratio out at
all satisfactorily.
A second effect which must be dealt with is
a large-amplitude ripple in the spectrum which
is thought to be caused by multiple reflections
in the waveplate. The ripple was removed using
the FFT technique described by Adamson et al.
(1999). Briefly, the ripple causes peaks to ap-
pear around 0.5 and 1.0 times the Nyquist fre-
quency in the Fourier transform of the spectrum.
These peaks were set to zero, then the transform
was inverted. The instrumental zero point of po-
larization was determined and removed from the
NGC1068 q and u spectra using observations of
the unpolarized star HD18803 (Clayton & Martin
1981), and the resulting polarization spectrum
corrected for the imperfect efficiency of the wave-
plate2. The position angle of polarization (PA)
was calibrated using observations of the young
stellar object, AFGL2519, whose PA at 3-4 µm has
previously been measured by Hough et al. (1989).
An intensity spectrum was also constructed using
only a few frames taken near in time to HD18803,
which was used as a telluric standard.
The polarization, position angle and total flux
spectra of NGC1068 are shown in Figure 1, to-
gether with the polarimetric results of Bailey et al.
(1988) and Lebofsky et al. (1978).
4. Polarization Calculations
To calculate the L-band polarization produced
by coated dust grains, we have taken advantage
of the discrete dipole approximation (DDA) code,
DDSCAT6.1 (Draine 1988; Draine & Flatau 1994,
2004). In the DDA the dust grain is represented
as an array of dipole oscillators on a cubic lattice,
and absorption, scattering and extinction cross-
sections are then calculated from the amplitudes
of the dipoles as they interact with the incident
electric field.
2 as given by
www.jach.hawaii.edu/UKIRT/instruments/irpol/CGS4/cgs4pol.html#5
The optical constants adopted for the grain
cores and mantles are those of "astronomical sili-
cate" (Draine & Lee 1984; Draine 1985; Weingartner & Draine
2001) and the organic refractory material of
Li & Greenberg (1997). The refractive index of
the organic component is derived from the Murchi-
son meteorite and laboratory residues exposed to
the solar UV field aboard the EURECA satel-
lite (Greenberg et al. 1995; Greenberg & Li 1996).
While both of those materials show absorption
bands at 5 - 9 µm that are not seen in the dif-
fuse ISM (Pendleton & Allamandola 2002), the
3.4 µm bands of both bear a good resemblance to
the observed feature.
The existence of polarization caused by dust
absorption indicates that the dust grains must
be aspherical, but beyond that their shapes are
not well-established. We have treated the dust
grains as oblate spheroids with an axial ratio of
2:1, the cores and mantles having equal eccen-
tricities. Based on fits to the Trapezium 9.7µm
silicate feature, Draine & Lee (1984) find that
2:1 oblate spheroids are a reasonable representa-
tion of dust grains, and Hildebrand & Dragovan
(1995) also favor oblate over prolate spheroids.
They conclude that the axial ratio must be <3:1,
with a best-fitting value of 1.5:1. There has
been some dispute over whether oblate or prolate
spheroids give a better fit to the observational data
(Greenberg & Li 1996), but we note that the main
discrepancy in the polarization predicted by these
two shapes comes over the 9.7µm silicate feature.
The differences are small at the wavelengths of in-
terest in this work (see Fig. 6 of Draine & Lee
1984). The elongation of the grains will affect the
efficiency with which they polarize, but has little
effect on the wavelength dependence of the polar-
ization (Greenberg & Li 1996).
We have adopted simple picket-fence alignment
for this work (which is equivalent to the per-
fect spinning alignment approximation for oblate
grains), with the grains' short axes orientated per-
pendicular to the direction of propagation of the
incident light. This has been shown to be quite ad-
equate if only the wavelength dependence of the
polarization is required, rather than its absolute
value (Chlewicki & Greenberg 1990). The func-
tional form used for the grain size distribution is
the MRN a−3.5 power law with limits of 0.005 and
0.25 µm (Mathis et al. 1977).
5
Given this information about the size, shape,
composition and orientation of the grains, DDSCAT
calculates a number of quantities. Those relevant
to this work are
Qabs = Cabs/πa2
Qsca = Csca/πa2
Qext = Qabs + Qsca
Qpol = Qext, k −Qext, ⊥
(1)
(2)
(3)
(4)
in which Qs are efficiencies and Cs cross-sections
for absorption, scattering and extinction, and
Qext, k and Qext, ⊥ refer to extinction efficien-
cies for the two orthogonal incident polarization
states (Draine 1988).
From Qext and Qpol, extinction and polariza-
tion spectra for ensembles of grains can be ob-
tained in the following manner:
τ (λ) = Ngrain Z amax
p(λ) = Ngrain Z amax
amin
amin
n(a)CgeoQext(a, λ)da (5)
n(a)Cgeo
Qpol
2
(a, λ)da (6)
where Cgeo is the geometrical cross-section of the
grain in question, Ngrain the column density of
grains (with n(a), the number of grains in the
interval a, a + da, appropriately normalized) and
amin and amax the limits of the size distribution
to be considered.
5. Results and discussion
We are interested in the magnitude of any in-
crease in polarization that would be expected over
the 3.4 µm absorption band, given a certain con-
tinuum polarization caused by dichroic absorption
by core-mantle grains. To obtain a crude esti-
mate of the polarization that might be expected
over the 3.4 µm band, we can consider the polar-
ization observed over the 9.7 µm silicate band in
galactic lines of sight. For a given grain shape,
size and composition, changes in abundance and
degree of alignment will alter the level of con-
tinuum polarization without affecting the ratio
of feature to continuum polarization, Pf eat/Pcont.
In galactic lines of sight that can be fitted with
pure absorptive polarization, this ratio is typically
about 3-4, with a couple of objects reaching val-
ues of ∼7 (OMC1 BN, SgrA GCS3 II; Smith et al.
6
(2000)). Alternatively, the optical constants of
astronomical silicate give rise to Pf eat/Pcont ≈
12 (Draine & Lee 1984; Hildebrand & Dragovan
1995). The "excess" polarization over an absorp-
tion feature is determined by the strength of the
band, and Li & Greenberg (2002) have shown that
silicates and hydrocarbons polarize to a similar de-
gree per unit optical depth. Observed values of
τ9.7/τ3.4 range from about 13-19 (Sandford et al.
1995), implying Pf eat/Pcont ∼1.15 - 2.0 in the 3.4
µm band. For NGC1068, where the L-band con-
tinuum polarization is about 2.5%, this translates
to peak polarizations of 2.9 - 5% across the 3.4
µm feature. This suggests that a significant ex-
cess polarization should be detectable over the 3.4
µm feature, although the spread in silicate feature
polarizations and τ9.7/τ3.4, and the fact that most
of the sources in Smith et al. (2000) are dominated
by molecular cloud material mean that these num-
bers should be treated with some caution.
To estimate the likely maximum observed
polarization change over the 3.4 µm band in
NGC 1068, the best-fitting linear continuum was
first found for the polarization data. A model
polarization spectrum was then created from the
optical depth spectrum derived from the spec-
tropolarimetry data, assuming that p ∝ τ . The
model spectrum was normalized to the continuum
around 3.55 µm, then scaled by various factors and
the reduced χ2 between it and the binned polar-
ization spectrum calculated for each scaling. The
best fit corresponds to a deviation of -0.12±0.25%
(95% confidence) from the linear continuum; the
data are in fact consistent with a small decrease in
polarization over the 3.4 µm feature. The model
fit giving the upper limit on the change in po-
larization over the 3.4 µm band (i.e., peaking at
0.13% above the continuum) is shown in Figure 2.
The polarization calculated as described in
§ 4 for a population of core-mantle grains with
the MRN size distribution is also shown in Fig.
2. As grain size may affect the magnitude of
the 3.4 µm polarization excess (§5.1), the po-
larization from a 0.2 µm grain, close to the
largest size in the MRN distribution and some-
what larger than the average < a >= 0.087µm
of the finite cylinders in the Li & Greenberg
(1997) model,
The different
curves in the figure result from different values
of fcarb, the fraction of the grain volume that
is also shown.
is contained in the mantle. Various figures have
been suggested for this quantity; Li & Greenberg
(2002) consider fcarb = 0.2, 0.5, 0.75 while ac-
cording to Chlewicki & Greenberg (1990), fcarb ∼
0.9. Greenberg & Li (1996) and Li & Greenberg
(1997) estimate mantle/core volume ratios of
about 2:1. As expected, the polarization at
the peak of the 3.4 µm band increases with the
amount of carbonaceous material in the grains,
from approximately 3.0% for the 25% -mantle
MRN grains, to about 3.6% for the 75%-mantle
particles.
For the core/mantle ratios closest to those most
commonly quoted in the literature, the 3.4 µm po-
larization predicted by the model is clearly well in
excess of the observed polarization. In the case of
the grains with only 25% of the volume contained
in the mantle, the polarization at the peak of the
3.4 µm absorption is 0.29% over the linear contin-
uum fit mentioned above. This is still somewhat
above the limit we derive on the observed polar-
ization; the expected and observed polarizations
are not reconciled even with only a small fraction
of each grain being composed of organic refractory
material. The core-mantle grain model, at least in
the form proposed for the Galactic diffuse ISM,
is unlikely to explain the L-band polarization of
NGC 1068.
5.1. Variations on the basic core-mantle
model
Are there likely to be significant differences in
the grain population(s) in NGC1068 that might di-
minish the amount of polarization expected over
the 3.4 µm band, even while core-mantle grains
are present? The similarity of the band profile
in NGC 1068 and the Galactic diffuse ISM sug-
gests that the composition of the hydrocarbon
material is similar in both galaxies (Dartois et al.
2004; Mason et al. 2004), but other differences
may arise. For instance, as the Li & Greenberg
(1997) model proposes that the entire infrared
polarization arises in the core-mantle grains, we
have so far also assumed this, but if another, non-
hydrocarbon-containing aligned grain component
also contributed to the polarization in NGC1068 it
would diminish the size of the rise in polarization
expected over the 3.4 µm band while contributing
to AV but not τ3.4. In NGC1068, such a second
polarizing grain population could conceivably be,
7
for example, bare silicates arising from destruction
of the mantles on some fraction of the grains.
Values of extinction to the complex, infrared-
emitting regions of the nucleus of NGC1068 may
not be as straightforward to interpret as the ex-
tinction through the large column of cold dust to-
wards the Galactic Center, but estimates range
from ∼15 (Lumsden et al. 1999; Watanabe et al.
2003) to ∼40 (Young et al. 1995), disregarding
values based on τ3.4. Given that τ3.4 ≈ 0.1, this
implies AV /τ3.4 ∼ 150−400, compared with galac-
tic values of ∼150 (towards the galactic center;
Pendleton et al. 1994) or ∼several hundred (to-
wards field stars at various galactic longitudes;
Rawlings et al. 2003). There is therefore no com-
pelling reason to think that grain components
other than those proposed for the galactic diffuse
ISM must exist in NGC1068, but the wide range
of estimates of AV /τ3.4 for both the galactic and
extragalactic lines of sight means that this cannot
be ruled out either.
Another issue that could affect the 3.4 µm band
in both extinction and polarization is that of grain
size.
Increasing grain size tends to increase the
efficiency of continuum extinction while decreas-
ing the strength of absorption features, so large
grains might be able to polarize the L-band con-
tinuum effectively without producing much excess
polarization through the 3.4 µm band. For grains
much larger than about 0.2 µm in radius, scatter-
ing starts to become important (Qabs/Qsca ≈ 5
for a 0.2 µm core-mantle 2:1 oblate spheroid at
3.3 µm) and calculations of the polarization from
such particles in a complex system like NGC1068,
where the inclination of the dusty torus and mul-
tiple scattering effects may be critical, are beyond
the scope of this paper.
It has been suggested that dust grains in AGN
may be biased towards larger sizes than in the dif-
fuse ISM, but the evidence remains contradictory.
Laor & Draine (1993) pointed out that large (∼10
µm) dust grains are likely to survive longer than
small grains close to an AGN and will not produce
a silicate emission feature. The weakness or ab-
sence of silicate emission in type 1 AGN prompted
the inclusion of large grains in some torus models
(van Bemmel & Dullemond 2003), but its recent
discovery in several quasars may argue against
large grains (Hao et al. 2005; Siebenmorgen et al.
2005). Flat L-M′ colors, peculiar EB−V /NH and
these observations
AV /NH ratios in Seyfert 2 nuclei and the absence
of the 2175A feature in reddened Seyfert 1s have
all been interpreted as evidence for large grains
(Imanishi 2001; Maiolino et al. 2001a,b), but geo-
metrical effects may also be able to explain many
of
(Weingartner & Murray
2002).
In the specific case of NGC1068, both
Young et al. (1995) and Watanabe et al. (2003)
were able to fit the observed J- to K-band and
optical polarization with grain sizes no different
from those thought to exist in the galactic DISM,
although Watanabe et al.
suggest that scatter-
ing from large grains in the torus might be a vi-
able alternative. Detailed extinction and polariza-
tion calculations such as those of Watanabe et al.
(2003),
if extended to the 3.4 µm band, could
provide valuable constraints on both the size and
structure of the grains in AGN tori.
Finally, we note that if the grain size distribu-
tion is biased towards larger sizes in NGC1068,
this could be through preferential destruction of
small grains, or large grains may have grown by
coagulation in this warm, dense environment. If
the latter, then they may have lost some of their
former core-mantle nature.
It seems likely that
small hydrocarbon inclusions in large, coagulated
grains would leave their polarization signature on
the 3.4 µm band, but again, further calculations
would be needed to test this.
5.2. Hydrocarbon formation in the diffuse
ISM
The absence of excess polarization over the 3.4
µm band is naturally explained if the feature arises
in a population of grains that does not contribute
significantly to the continuum polarization. Such
a grain population must be small and/or optically
isotropic (small grains being harder to align than
large ones; Lazarian 2003, and references therein),
and have no physical connection to the silicate
grains. Recent laboratory work has provided per-
suasive evidence that small carbon grains like
those thought to be ejected from AGB stars can
be hydrogenated during the later stages of stellar
evolution (Schnaiter et al. 1999) and in the dif-
fuse ISM (Mennella et al. 1999; Munoz Caro et al.
2001; Mennella et al. 2002). Calculations indi-
cate that hydrogenation proceeds at a fast enough
rate in the diffuse ISM to balance the dehydro-
genation caused by UV photons. Conversely,
8
in dense molecular clouds, the reduced abun-
dance of atomic H and the presence of icy man-
tles act to prevent rehydrogenation of the grains,
which can still be efficiently dehydrogenated by
photons and cosmic rays (Mennella et al. 2001,
2003). These and other lines of evidence (e.g.
Shenoy et al. 2003) imply that most hydrocarbon-
containing grains are formed in the diffuse ISM
and that re-formation dominates over the dehy-
drogenation that subsequently occurs, in agree-
ment with the non-detection of the 3.4 µm feature
in dense cloud material. This evolutionary sce-
nario for the aliphatic hydrocarbons is entirely
consistent with the lack of a 3.4 µm polarization
excess in NGC 1068 and several Galactic lines
of sight (Adamson et al. 1999; Ishii et al. 2002;
Chiar et al. 2006).
6. Summary
We have performed L-band spectropolarimetry
of NGC 1068 and shown that the excess polariza-
tion over the 3.4 µm feature is below that which
would be expected on the basis of the silicate-
core/organic-mantle grain model as applied to the
galactic diffuse ISM, consistent with a growing
body of evidence suggesting that the aliphatic
hydrocarbons in the general diffuse ISM are not
formed by processing of the ice mantles that form
on silicate grains in molecular clouds. The coated
grain model could still be valid in NGC 1068
if there also exists an extra, non-hydrocarbon
aligned grain population, or possibly if the grain
size distribution is biased to larger sizes than in the
diffuse ISM of our galaxy (detailed calculations of
both the continuum and 3.4 µm feature polariza-
tion from micron-sized dust grains may be a useful
way of constraining the size and/or composition of
the carbonaceous grain population in AGN). Al-
ternatively, reaction of small carbon grains with
atomic hydrogen in the diffuse ISM would be ex-
pected to produce a population of small, nonpolar-
izing hydrocarbon-containing grains which would
naturally explain the lack of polarization of the
3.4 µm feature. Such a model is also successful
in accounting for the non-detection of the 3.4 µm
band in molecular clouds, which is otherwise diffi-
cult to explain.
7. Acknowledgments
We would like to thank R. Antonucci, P. Hirst,
M. Kishimoto, T. Roush, M. Smith and A. Tie-
lens for taking the time to comment, and A. Li for
providing the optical constants in tabular form.
We are grateful to the anonymous referee for com-
ments that strengthened the paper. We thank
the Department of Physical Sciences, University
of Hertfordshire for providing IRPOL2 for the
UKIRT. The United Kingdom Infrared Telescope
is operated by the Joint Astronomy center on be-
half of the U.K. Particle Physics and Astronomy
Research Council. Supported by the Gemini Ob-
servatory, which is operated by the Association of
Universities for Research in Astronomy, Inc., on
behalf of the international Gemini partnership of
Argentina, Australia, Brazil, Canada, Chile, the
United Kingdom, and the United States of Amer-
ica.
REFERENCES
Adamson, A. J., Whittet, D. C. B., Chrysosto-
mou, A., Hough, J. H., Aitken, D. K., Wright,
G. S., & Roche, P. F. 1999, ApJ, 512, 224
Aitken, D. K., Briggs, G., Bailey, J. A., Roche,
P. F., & Hough, J. H. 1984, Natur, 310, 660
Aitken, D. K., Efstathiou, A., McCall, A., &
Hough, J. H. 2002, MNRAS, 329, 647
Alloin, D., Pantin, E., Lagage, P. O., & Granato,
G. L. 2000, A&AP, 363, 926
Antonucci, R. R. J., & Miller, J. S. 1985, ApJ,
297, 621
Bailey, J., Axon, D. J., Hough, J. H., Ward, M. J.,
McLean, I., & Heathcote, S. R. 1988, MNRAS,
234, 899
Bock, J. J., Neugebauer, G., Matthews, K., Soifer,
B. T., Becklin, E. E., Ressler, M., Marsh, K.,
Werner, M. W., Egami, E., & Blandford, R.
2000, AJ, 120, 2904
Bridger, A., Wright, G. S., & Geballe, T. R. 1994,
in ASSL Vol. 190: Astronomy with Arrays, The
Next Generation, 537
Chiar, J. E., Adamson, A. J., Whittet, D. C. B.,
Chrysostomou, A., Hough, J. H., Kerr, T. H.,
9
Mason, R. E., Poche, P. F., & Wright, G. 2006,
ArXiv Astrophysics e-prints
Chiar, J. E., Pendleton, Y. J., Geballe, T. R., &
Tielens, A. G. G. M. 1998, ApJ, 507, 281
Chiar, J. E., Tielens, A. G. G. M., Whittet,
D. C. B., Schutte, W. A., Boogert, A. C. A.,
Lutz, D., van Dishoeck, E. F., & Bernstein,
M. P. 2000, ApJ, 537, 749
Chlewicki, G., & Greenberg, J. M. 1990, ApJ, 365,
230
Clayton, G. C., & Martin, P. G. 1981, AJ, 86, 1518
Dartois, E., Marco, O., Munoz-Caro, G. M.,
Brooks, K., Deboffle, D., & d'Hendecourt, L.
2004, A&A, 423, 549
Demyk, K., Dartois, E., D'Hendecourt, L., Jour-
dain de Muizon, M., Heras, A. M., & Breitfell-
ner, M. 1998, A&A, 339, 553
Draine, B. T. 1985, ApJS, 57, 587
-- . 1988, ApJ, 333, 848
Draine, B. T., & Flatau, J. 1994, J. Opt. Soc Am.,
11, 1491
Draine, B. T., & Flatau, P. J. 2004, User Guide
for the Discrete Dipole Approximation Code
DDSCAT (Version 6.1), astro-ph/0309069
Draine, B. T., & Lee, H. M. 1984, ApJ, 285, 89
Efstathiou, A., McCall, A., & Hough, J. H. 1997,
MNRAS, 285, 102
Garc´ıa-G´omez, C., Athanassoula, E., & Barber`a,
C. 2002, A&A, 389, 68
Geballe, T. R., Baas, F., & Wade, R. 1989, A&A,
208, 255
Gibb, E. L., & Whittet, D. C. B. 2002, ApJL, 566,
L113
Greenberg, J. M., & Li, A. 1996, A&A, 309, 258
Greenberg, J. M., Li, A., Mendoza-Gomez, C. X.,
Schutte, W. A., Gerakines, P. A., & de Groot,
M. 1995, ApJL, 455, L177
Hao, L., Spoon, H. W. W., Sloan, G. C., Mar-
shall, J. A., Armus, L., Tielens, A. G. G. M.,
Sargent, B., van Bemmel, I. M., Charmandaris,
V., Weedman, D. W., & Houck, J. R. 2005,
ApJ, 625, L75
Hildebrand, R. H., & Dragovan, M. 1995, ApJ,
450, 663
Hough, J. H., Sato, S., Tamura, M., Yamashita,
T., McFadean, A. D., Rouse, M. F., Kaifu, N.,
Suzuki, H., Nagata, T., Gatley, I., & Bailey, J.
1988, MNRAS, 230, 107
Hough, J. H., Whittet, D. C. B., Sato, S., Ya-
mashita, T., Tamura, M., Nagata, T., Aitken,
D. K., & Roche, P. F. 1989, MNRAS, 241, 71
Imanishi, M. 2000, MNRAS, 319, 331
-- . 2001, AJ, 121, 1927
Imanishi, M., Terada, H., Sugiyama, K., Moto-
hara, K., Goto, M., & Maihara, T. 1997, PASJ,
49, 69
Ishii, M., Nagata, T., Chrysostomou, A., &
Hough, J. H. 2002, AJ, 124, 2790
Jaffe, W., Meisenheimer, K., Rottgering, H. J. A.,
Leinert, C., Richichi, A., Chesneau, O., Fraix-
Burnet, D., Glazenborg-Kluttig, A., Granato,
G.-L., Graser, U., Heijligers, B., Kohler, R.,
Malbet, F., Miley, G. K., Paresce, F., Pel, J.-
W., Perrin, G., Przygodda, F., Schoeller, M.,
Sol, H., Waters, L. B. F. M., Weigelt, G.,
Woillez, J., & de Zeeuw, P. T. 2004, Nature,
429, 47
Maiolino, R., Marconi, A., & Oliva, E. 2001a,
A&A, 365, 37
Maiolino, R., Marconi, A., Salvati, M., Risaliti,
G., Severgnini, P., Oliva, E., La Franca, F., &
Vanzi, L. 2001b, A&A, 365, 28
Marco, O., & Brooks, K. J. 2003, A&A, 398, 101
Mason, R. E., Geballe, T. R., Packham, C., Lev-
enson, N. A., Elitzur, M., Fisher, R. S., & Perl-
man, E. 2006, ApJ, 640, 612
Mason, R. E., Wright, G., Pendleton, Y., & Adam-
son, A. 2004, ApJ, 613, 770
Mathis, J. S., Rumpl, W., & Nordsieck, K. H.
1977, ApJ, 217, 425
Matt, G., Fabian, A. C., Guainazzi, M., Iwasawa,
K., Bassani, L., & Malaguti, G. 2000, MNRAS,
318, 173
Mennella, V., Baratta, G. A., Esposito, A., Ferini,
G., & Pendleton, Y. J. 2003, ApJ, 587, 727
Mennella, V., Brucato, J. R., Colangeli, L., &
Palumbo, P. 1999, ApJL, 524, L71
-- . 2002, ApJ, 569, 531
Mennella, V., Munoz Caro, G. M., Ruiterkamp,
R., Schutte, W. A., Greenberg, J. M., Brucato,
J. R., & Colangeli, L. 2001, A&A, 367, 355
Munoz Caro, G. M., Ruiterkamp, R., Schutte,
W. A., Greenberg, J. M., & Mennella, V. 2001,
A&A, 367, 347
Nagata, T., Kobayashi, N., & Sato, S. 1994, ApJL,
Kleinmann, D. E., Gillett, F. C., & Wright, E. L.
423, L113+
1976, ApJ, 208, 42
Laor, A., & Draine, B. T. 1993, ApJ, 402, 441
Lazarian, A. 2003, JQSRT, 79, 881
Lebofsky, M. J., Kemp, J. C., & Rieke, G. H. 1978,
ApJ, 222, 95
Li, A., & Greenberg, J. M. 1997, A&A, 323, 566
-- . 2002, ApJ, 577, 789
Lumsden, S. L., Moore, T. J. T., Smith, C., Fu-
jiyoshi, T., Bland-Hawthorn, J., & Ward, M. J.
1999, MNRAS, 303, 209
Packham, C., Young, S., Hough, J. H., Axon,
D. J., & Bailey, J. A. 1997, MNRAS, 288, 375
Pendleton, Y. J., & Allamandola, L. J. 2002,
ApJS, 138, 75
Pendleton, Y. J., Sandford, S. A., Allamandola,
L. J., Tielens, A. G. G. M., & Sellgren, K. 1994,
ApJ, 437, 683
Pendleton, Y. J., Tielens, A. G. G. M., Tokunaga,
A. T., & Bernstein, M. P. 1999, ApJ, 513, 294
Pendleton, Y. J., Tielens, A. G. G. M., & Werner,
M. W. 1990, ApJ, 349, 107
10
Pogge, R. W., & Martini, P. 2002, ApJ, 569, 624
Weingartner, J. C., & Murray, N. 2002, ApJ, 580,
Rawlings, M. G., Adamson, A. J., & Whittet,
D. C. B. 2003, MNRAS, 341, 1121
Rhee, J. H., & Larkin, J. E. 2006, ApJ, 640, 625
Risaliti, G., Maiolino, R., Marconi, A., Bas-
sani, L., Berta, S., Braito, V., Della Ceca, R.,
Franceschini, A., & Salvati, M. 2003, ApJL,
595, L17
Roche, P. F., Whitmore, B., Aitken, D. K., &
Phillips, M. M. 1984, MNRAS, 207, 35
Sandford, S. A., Allamandola, L. J., Tielens,
A. G. G. M., Sellgren, K., Tapia, M., & Pendle-
ton, Y. 1991, ApJ, 371, 607
Sandford, S. A., Pendleton, Y. J., & Allamandola,
L. J. 1995, ApJ, 440, 697
Schnaiter, M., Henning, T., Mutschke, H., Kohn,
B., Ehbrecht, M., & Huisken, F. 1999, ApJ, 519,
687
Shenoy, S. S., Whittet, D. C. B., Chiar, J. E.,
Adamson, A. J., Roberge, W. G., & Hassel,
G. E. 2003, ApJ, 591, 962
Siebenmorgen, R., Haas, M., Krugel, E., & Schulz,
B. 2005, A&A, 436, L5
Simpson, J. P., Colgan, S. W. J., Erickson, E. F.,
Hines, D. C., Schultz, A. S. B., & Trammell,
S. R. 2002, ApJ, 574, 95
Smith, C. H., Wright, C. M., Aitken, D. K., Roche,
P. F., & Hough, J. H. 2000, MNRAS, 312, 327
Tinbergen, J. 1996, Astronomical polarimetry
(Cambridge, New York: Cambridge University
Press)
Tomono, D., Doi, Y., Usuda, T., & Nishimura, T.
2001, ApJ, 557, 637
van Bemmel, I. M., & Dullemond, C. P. 2003,
A&A, 404, 1
Watanabe, M., Nagata, T., Sato, S., Nakaya, H.,
& Hough, J. H. 2003, ApJ, 591, 714
Weingartner, J. C., & Draine, B. T. 2001, ApJ,
548, 296
88
Wright, G. S., Bridger, A., Geballe, T. R., &
Pendleton, Y. 1996, in ASSL Vol. 209: New
Extragalactic Perspectives in the New South
Africa
Young, S., Hough, J. H., Axon, D. J., Bailey, J. A.,
& Ward, M. J. 1995, MNRAS, 272, 513
This 2-column preprint was prepared with the AAS LATEX
macros v5.2.
11
Fig. 2. -- Observed and predicted polarization over
the 3.4 µm feature in NGC1068 for an MRN distri-
bution of coated silicate particles with different thick-
nesses of mantle. Also shown are the polarization of
a 0.2 µm grain (fcarb= 0.25), close to the largest size
in the MRN distribution, and the limit on the change
in polarization over the feature based on a model fit
to the observed polarization (see text). The column
density of grains (eq. 6) was chosen to reproduce the
observed polarization around 3.55 µm, a region ex-
pected to be unaffected by the 3.4 µm feature.
Fig.
1. -- Top: L-band polarization spectrum of
NGC 1068. The binned spectrum was obtained by
calculating the standard error of the q and u spec-
tra in 0.010 µm bins and is offset by 3% for clar-
ity. Also shown are the measurements of Bailey et al.
(1988,
square,
offset by 0.01 µm in wavelength). Middle: continuum-
subtracted flux spectrum, showing the peak of the
3.4 µm absorption feature. Bottom: position angle of
polarization, with the measurements of Bailey et al.
(1988) and Lebofsky et al. (1978), as above.
triangles) and Lebofsky et al. (1978,
12
|
0801.0740 | 2 | 0801 | 2008-05-14T19:39:24 | Non-thermal X-rays from the Ophiuchus galaxy cluster and dark matter annihilation | [
"astro-ph",
"hep-ph"
] | We investigate a scenario where the recently discovered non-thermal hard X-ray emission from the Ophiuchus cluster originates from inverse Compton scattering of energetic electrons and positrons produced in weakly interacting dark matter pair annihilations. We show that this scenario can account for both the X-ray and the radio emission, provided the average magnetic field is of the order of 0.1 microGauss. We demonstrate that GLAST will conclusively test the dark matter annihilation hypothesis. Depending on the particle dark matter model, GLAST might even detect the monochromatic line produced by dark matter pair annihilation into two photons. | astro-ph | astro-ph |
Non-thermal X-rays from the Ophiuchus galaxy cluster and dark matter annihilation
Stefano Profumo1, ∗
1Santa Cruz Institute for Particle Physics and Department of Physics,
University of California, Santa Cruz CA 95064
We investigate a scenario where the recently discovered non-thermal hard X-ray emission from the
Ophiuchus cluster originates from inverse Compton scattering of energetic electrons and positrons
produced in weakly interacting dark matter pair annihilations. We show that this scenario can ac-
count for both the X-ray and the radio emission, provided the average magnetic field is of the order
of 0.1 µG. We demonstrate that GLAST will conclusively test the dark matter annihilation hypoth-
esis. Depending on the particle dark matter model, GLAST might even detect the monochromatic
line produced by dark matter pair annihilation into two photons.
PACS numbers: 95.35.+d,, 98.80.Cq, 95.85.Nv, 95.85.Pw
Clusters of galaxies are the largest bound dark matter
(DM) structures in the universe. As such, they are nat-
ural targets for the search for observational signatures
of particle DM [1]. If DM is in the form of weakly in-
teracting massive particles (WIMPs) [2, 3], DM pair an-
nihilations generically produce γ-rays as well as a non-
thermal energetic electron-positron (e±) population. The
latter, in turn, is expected to yield secondary emissions at
soft γ-ray, X-ray and radio frequencies via inverse Comp-
ton scattering, bremsstrahlung and synchrotron radia-
tion, opening up the possibility of a multi-wavelength
approach to particle DM detection [1, 4]. A generic fea-
ture of the broad-band DM annihilation spectrum is a
significant hard X-ray component [1, 5].
Interestingly, the discovery of a non-thermal hard X-
ray emission from the Ophiuchus cluster, detected with
relatively robust statistical significance in a 3 Ms obser-
vation with the IBIS/ISGRI and JEM-X instruments on
board INTEGRAL, was recently reported in Ref. [6]. The
Ophiuchus cluster is a nearby (z ≃ 0.028 [7]) rich cluster
with a high temperature plasma (kT ∼ 10 keV), featur-
ing the second brightest emission in the 2-10 keV band.
In addition to X-ray observations, the steep-spectrum ra-
dio source MSH 17-203 [8] was associated to the Ophi-
uchus cluster [7], indicating the presence of relativistic
electrons. However, the cluster was not detected at γ-
ray frequencies by EGRET [9].
Several clusters are known to host extended radio
emissions [10], suggesting the existence of energetic
non-thermal electrons that radiate at radio frequencies
through synchrotron emission. The same electron pop-
ulation should also produce hard X-rays via inverse-
Compton (IC) scattering off cosmic microwave back-
ground photons. Up to now, however, firm evidence
for extended non-thermal hard X-ray emission in clus-
ters was still missing, with a few controversial [11, 12]
exceptions, including BeppoSAX [13] and INTEGRAL
[14] observations of the Coma cluster, BeppoSAX obser-
vations of Abell 2256 [15] and Chandra observations of
the Perseus cluster [16]. In addition, under the assump-
tion of negligible contamination from obscured AGNs,
[17] reports ∼ 2σ detections, using the BeppoSAX PDS
instrument, of hard X-ray non-thermal components from
a few more merging clusters, including Abell 2142, 2199,
3376, Virgo and the Ophiuchus cluster itself.
While ordinary astrophysical mechanisms,
including
merger shocks, can be invoked to explain the non-thermal
electrons presumably responsible for the observed non-
thermal activity in galaxy clusters, in the present anal-
ysis we propose and investigate a novel scenario where
WIMP annihilations produce, or significantly contribute
to, said non-thermal population responsible for the hard
X-ray detection in the Ophiuchus galaxy cluster.
In a
model independent approach, we determine the param-
eters of the particle DM setups that provide the best
fits to the INTEGRAL X-ray data, and we compute the
resulting multi-wavelength spectra. We then compare
these spectra with the radio data and with the γ-ray
limits and future prospects for the soon-to-be-launched
Gamma-Ray Large Area Telescope (GLAST). The high-
lights of our analysis are: (1) the DM hypothesis will
conclusively be probed with GLAST; (2) the radio emis-
sion can in principle also be fitted with the synchrotron
emission from DM-annihilation-produced e±, as long as
the average magnetic field in the cluster is of the order of
0.1 µG; (3) GLAST might be able detect the monochro-
matic γ-rays produced in direct DM pair annihilation
into two photons.
The flux of e± produced by WIMP pair annihilations
depends on the particle DM setup and on the DM den-
sity distribution. We define a source function Qe(Ee, ~x),
which gives the number of e± per unit time, energy and
volume element produced locally in space, as
Qe(Ee, ~x) = hσvi0 Xf
dN f
e
dEe
(Ee) Bf Npairs(~x).
(1)
In the equation above, hσvi0 is the WIMP annihilation
rate at zero temperature, the sum is over all kinemat-
ically allowed Standard Model annihilation final states
f , each with a branching ratio Bf and an e± distri-
bution dN f
e /dEe, and Npairs(~x) is the number density
i.e.
DM(~x)/(2m2
the number
of WIMP pairs at a given point ~x,
of WIMP particle pairs per volume element squared:
Npairs(~x) = ρ2
DM), where ρDM stands for the
DM density. The particle physics framework sets the
quantity hσvi0, the list of Bf and the mass of the WIMP,
mDM. The latter also determines the energy scale of the
pair annihilation event, and, together with the specific
final state f , the dN f
e /dEe spectral functions, which we
numerically compute with the Monte Carlo code Pythia
[18]. In addition, mDM enters in the determination of the
local number density of WIMP pairs.
Once the source function Qe(Ee, ~x) is determined, the
e± spectrum and density are affected by spatial diffusion
and energy loss processes, usually described -- under the
assumptions of negligible convection and re-acceleration
effects -- by a diffusion-loss equation of the form
]
1
-
s
2
-
m
c
s
n
o
t
o
h
p
[
x
u
l
F
10-1
10-2
10-3
10-4
10-5
10-6
10-7
10-8
10-9
10-10
101
2
]
1
-
s
2
-
10-3
m
c
s
n
o
t
o
h
p
[
x
u
l
F
10-4
10-5
20
30
70
Photon Energy [keV]
50
100
b b, mχ=46 GeV
W+W-, mχ=82 GeV
τ+τ−
, mχ=10 GeV
102
103
104
Photon Energy [keV]
105
EGRET limit
106
∂
∂t
dne
dEe
(Ee, ~x) = ~∇ ·hD(Ee, ~x)~∇
∂Eehb(Ee, ~x)
dne(Ee, ~x)
dEe
∂
dEe
dne(Ee, ~x)
i
i + Qe(Ee, ~x),
+
FIG. 1: The hard X-ray and γ-ray spectrum for three DM par-
ticle models, plus a single-temperature MEKAL model [19] for
the thermal X-ray emission, compared with the INTEGRAL
data [6] and with the EGRET upper limit [9].
(2)
where dne/dEe is the number density of electrons per
unit energy, D is the diffusion coefficient, and
b(Ee, ~x) = bIC + bsyn + bCoul + bbrem
(3)
encodes the various energy loss mechanisms [1].
Knowledge of the distribution of the DM-induced
e± population dne/dEe, of the magnetic field struc-
ture and strength, as well as of the electron, gas and
starlight densities allows one to compute the WIMP-
induced secondary emissions. Specifically, at radio fre-
quencies the DM-induced emission is dominated by the
synchrotron radiation of the relativistic secondary elec-
trons and positrons.
IC scattering of the non-thermal
e± on target CMB and starlight photons gives rise to a
spectrum of photons stretching from below the extreme
ultra-violet up to the soft γ-ray band, peaking in the
X-ray energy band. Non-thermal bremsstrahlung, i.e.
the emission of γ-ray photons in the deflection of the
charged particles by the electrostatic potential of ion-
ized gas, contributes in the soft γ-ray band. Finally, a
hard γ-ray component arises from prompt emission in
WIMP pair annihilations, mostly originating from the
two photon decay of neutral pions, and, at the high en-
ergy end of the spectrum, from internal bremsstrahlung
from charged particle final states. The γ-ray spectrum
extends up to energies equal to the kinematic limit set by
the WIMP mass, Eγ ≤ mDM, and might feature one or
more monochromatic lines associated to two-body anni-
hilation final states where one (or both) of the particles
is a photon. We refer the reader to Ref. [1] for details on
the computation of the multi-wavelength emission from
DM annihilation.
We show in Fig. 1 the photon flux in the hard X-ray
and γ-ray bands for three benchmark DM particle model
accounting for the INTEGRAL data. We also include
a thermal component from the bremsstrahlung emis-
sion of the intra-cluster medium, obtained with a single-
temperature MEKAL model [19] with the abundance
fixed to 0.49 compared to the solar value and kT = 8.5
keV [6]. The thermal component was normalized to pro-
duce the best fit for the INTEGRAL data below 20 keV.
The DM models were instead normalized to obtain the
best global fit to the data above 20 keV. We chose DM
models with Bf = 1 for f = b¯b, W +W −, τ +τ −, i.e. each
model pair annihilating into a single Standard Model fi-
nal state. The three particular final states were selected
for two reasons: (1) the resulting e± spectra dN f
e /dEe
range from the softest (b¯b) to the hardest (τ +τ −) possible
case [1]; (2) the three final states correspond to common
well-defined cases found in supersymmetric DM models
[2]. For instance, in the minimal supergravity scenario
[20] f ≃ b¯b corresponds to the so-called bulk and funnel
regions where the neutralino has a relic abundance com-
patible with the cold DM density, τ +τ − is found in the
coannihilation region (where neutralino pair-annihilation
proceeds predominantly through scalar tau exchange)
and W +W − in the focus point region [21]. This choice of
benchmark models follows here and generalizes the ap-
proach of [1] -- linear combinations of the considered mod-
els produce almost any WIMP multi-wavelength emission
spectrum.
For each final state, we selected the DM particle mass
giving the lowest χ2 in the fit to the INTEGRAL data:
mDM (f = b¯b) = 46 GeV, mDM (f = W +W −) = 82
GeV and mDM (f = τ +τ −) = 10 GeV. Following [22] we
assumed, for the diffusion coefficient, the form
D(Ee) = D0
d2/3
B
B1/3
µ (cid:18) Ee
1 GeV(cid:19) , D0 = 3.1 × 1028 cm2s−1,
where dB ≃ 20 is the minimum scale of uniformity of
the magnetic field in kpc and Bµ is the average mag-
netic field in µG. Notice that, while we take into account
spatial diffusion in the numerical computation, the effect
of changing parameters such as D0 or dB is minimal on
the resulting WIMP multi-wavelength spectrum. In the
computation of the photon flux we neglected the IC from
starlight and assumed an average thermal gas density
nth = 10−3 cm3, relevant for the computation of bCoul
and bbrem in Eq. (3).
In the portion of the spectrum
shown in Fig. 1 the value of the magnetic field is not cru-
cial, and was fixed here, for reference, to B = 0.1 µG.
As shown in the inset, the fit to the INTEGRAL data
improves dramatically with the contribution from DM
annihilation. Also, the emission in the γ-ray band is
compatible with the EGRET limit [9], shown for refer-
ence with a horizontal orange line.
What is the DM pair annihilation cross section re-
quired to reproduce the spectra shown in Fig. 1? To
answer this question we need to integrate the number
density of DM pairs over the line of sight. In turn, this re-
quires knowledge of the DM density profile for the Ophi-
uchus cluster. We follow here the analysis of Ref. [1],
and assume, for reference, the DM density profile ob-
tained in the numerical simulations of Ref. [25] (namely
ρ(r) = ρ0g(r/a), with g(x) = exp[−2(xα − 1)/α] and
α ≃ 0.17) and the DM substructure setup outlined in
[26]. We verified that using other DM profiles changes
our predictions by less than one order of magnitude. We
derived from Ref. [27] a virial mass of ∼ 1.5×1015M⊙h−1
and a virial concentration of ∼ 10. We parametrize the
contribution of substructures via the fraction fs of total
mass in subhalos, and assume a ratio Rs = 5 between
the concentration parameter in subhalos and that in iso-
lated halos with equal mass, Rs ≡ hcsi/hcviri. In Fig. 2
(a) we compute the cross section as a function of the
DM particle mass, for the three benchmark final states,
giving the best fit to the INTEGRAL data. We assume
fs = 0.5 for the lower lines, as suggested by numerical
simulations [28], while we neglect the contribution from
substructures for the upper lines.
While we find rather large values for the pair annihi-
lation cross section compared to the naive expectation
hσvi0 ≃ 3 × 10−26 cm3s−1 motivated by requiring a ther-
mal WIMP relic abundance compatible with the CDM
density through simple scaling arguments, the range we
get is consistent with several examples of supersymmet-
ric DM models (see e.g. Fig. 15 in [1]). Also, the values
we obtain for both hσvi0 and mDM are consistent with
all available particle physics constraints on DM, and are
compatible with WIMPs being in the right density to-
100
1000
3
]
1
-
s
3
m
c
[
0
>
v
σ
<
10-7
10-8
)
γ
γ
>
-
M
D
M
D
(
R
B
10
10-19
10-20
10-21
10-22
10-23
10-24
10-25
(a)
(b)
]
1
-
s
2
-
m
c
.
h
p
[
10-2
10-3
10-4
10
Cross Section
Integrated γ ray flux
(Eγ>100 MeV)
Best Fit Masses
(c)
Monoch. γ ray line
b b, mχ=46 GeV
W+W-, mχ=82 GeV
τ+τ−
, mχ=10 GeV
100
mDM [GeV]
1000
FIG. 2: The preferred pair annihilation cross section (a), the
integrated γ-ray flux above 100 MeV (b) and the minimal
branching ratio for the detection of the monochromatic γ-
ray line at Eγ = mDM (c), as a function of the DM particle
mass.
In panel (a) the upper lines refer to the case of no
substructures, the lower lines refer to the substructure setup
described in the text, and the gray shaded region is ruled out
by EGRET [23] and H.E.S.S. [24] data on the gamma-ray flux
from the galactic center region.
day provided, for instance, non-thermal production or a
modified cosmological expansion rate is assumed at the
time of WIMP freeze-out [29]. In addition, uncertainties
on (1) the dark matter density distribution and (2) the
galactic and extra-galactic gamma-ray background un-
dermine the possibility of ruling out WIMP models with
large pair-annihilation cross sections via gamma-ray data
from the center of the Galaxy [23, 24, 30, 31], nearby
galaxies [32, 33, 34], and from the galactic halo [35]. For
instance, in the upper panel of fig. 2 we shade in gray
the region ruled out by EGRET and H.E.S.S. data from
the galactic center region, assuming a cored dark mat-
ter profile following the analysis of Ref. [30, 31]. Clearly,
gamma-ray data from the galactic center do not rule out
the range of cross sections we find.
Similarly, antimatter constraints not only depend on
the dark matter density across the hole galactic halo, but
are also affected by sizable uncertainties in the diffusion
and propagation of cosmic rays in the Galaxy [35]. Even
for cross sections as large as to account for the EGRET
log10 ( ν [Hz] )
15
10
20
25
b b, mχ=46 GeV
W+W-, mχ=82 GeV
τ+τ−
, mχ=10 GeV
INTEGRAL data
Radio data
T
S
A
L
G
T
h
e
r
m
a
l
B
r
e
m
s
.
10-10 10-8
10-6
10-2
100
10-4
102
Photon Energy [keV]
104
106
108
]
1
-
s
2
-
m
c
g
r
e
[
)
ν
(
S
ν
10-10
10-11
10-12
10-13
10-14
10-15
FIG. 3: The spectral energy distribution for the multi-
wavelength emission of three DM particle models and for the
thermal X-ray emission. We also show the INTEGRAL [6]
and radio [8] data.
data on the galactic gamma-ray emission [36] (hence at
the level of the cross sections in the shaded gray region in
fig. 2), antimatter fluxes from WIMP annihilations are in
general compatible with available data [37]. Finally, both
the search for energetic neutrinos from WIMP annihila-
tion in the Sun or the Earth, and direct WIMP detection,
depend on the WIMP-nucleon scattering cross section, a
quantity which is unrelated to the WIMP pair annihila-
tion. Even in special models, such as supersymmetry, the
range of predictions is so wide to make it impossible to
constrain hσvi0 with neutrino fluxes or direct detection
searches [38, 39].
In conclusion, the WIMP pair anni-
hilation rates we consider here, while larger than what
naively expected, are compatible with WIMP cosmolog-
ical DM production and DM searches.
Panel (b) in Fig. 2 shows the integrated γ-ray flux
above 0.1 GeV for the best fit models as a function of
mDM. In all cases we find that the expected γ-ray flux is
well above the anticipated GLAST LAT integral flux sen-
sitivity, estimated to be around a few ×10−10 cm−2s−1
[40]. Fig. 2 (c) shows the branching ratio hσvitot/hσviγγ
for the monochromatic DM DM → γγ channel needed to
obtain, for the best fit models, the detection of at least
10 photons1 with Eγ = mDM. Notice that the values
shown are independent of the assumed DM profile. While
hσvitot/hσviγγ is entirely model dependent, the range we
obtain is generically consistent with what is expected e.g.
in supersymmetry [41], and especially in next-to-minimal
1 The branching ratio corresponding to a larger photon flux can
be obtained by linearly rescaling the lines in Fig. 2 (c).
10
0.25
0.2
0.15
0.1
0.05
]
G
µ
[
B
]
G
µ
[
B
0.25
0.2
0.15
0.1
0.05
10
4
100
1000
disfavored by radio data
b b
best fit to radio data
0.25
0.2
]
0.15
G
µ
[
B
0.1
0.05
W+ W-
τ+τ-
100
mDM [GeV]
1000
FIG. 4: Preferred values of the cluster magnetic field as a
function of the DM particle mass, for three representative
pair annihilation final states.
supersymmetric extensions of the Standard Model [42].
GLAST can therefore easily detect a sizable number of
monochromatic energetic γ rays, depending on the spe-
cific DM particle model.
We illustrate the whole broad-band spectrum for the
DM annihilation interpretation of the INTEGRAL hard
X-ray emission in Fig. 3. Particularly crucial here is the
question of whether the radio data on the Ophiuchus clus-
ter are compatible with our predictions. This question
depends on the assumed value of the average cluster mag-
netic field B. In the Figure we use values of B giving the
best fit to the radio data, namely B(f = b¯b) = 0.15 µG,
B(f = W +W −) = 0.1 µG and B(f = τ +τ −) = 0.18 µG.
Clearly, a linear superposition of the various final states
can yield a remarkably good fit to the available radio
data. We also indicate the anticipated GLAST sensitiv-
ity, which shows that a sizable flux of γ-rays is expected
if the DM annihilation scenario is indeed correct.
As pointed out in the analysis of Ref. [6], simple IC in-
terpretations of the INTEGRAL data force the estimate
for the magnetic field in the range B ∼ 0.1 − 0.2 µG,
which agrees with the present analysis. While this range
is compatible with the value obtained from the hard X-
ray emission in Coma [13], it is below the estimates ob-
tained, for other clusters, through Faraday rotation mea-
sures [43] (see however the discussion in [12]). We show
in Fig. 4 with black and red lines the values of B respec-
tively giving the best fit to radio data and exceeding by
5-σ the measured radio flux in at least one bin. While
values of B above the red lines are disfavored by radio
data, they are not strictly ruled out if one accounts, e.g.,
for a radial dependence of the magnetic field [1].
The detection of γ-rays from cosmic rays in clusters
of galaxies is generically expected to be possible with
GLAST [12, 44]. Specifically, the analysis of Ref. [45] in-
dicates that the Ophiuchus cluster will be detectable by
GLAST provided the ratio of the energy density of cos-
mic rays to that of the thermal gas is larger than 0.3%
to 8.8%, depending on assumptions on the cosmic ray
spectral index and radial distribution. This warrants a
systematic comparison between the expected γ-ray flux
from cosmic rays and from DM in clusters, which is cur-
rently under way [46].
In general, though, the γ-ray
spectra expected from cosmic rays (a power law above
∼ 1 GeV, with a spectral index depending on the pri-
mary cosmic-ray protons spectral index) are significantly
different from what expected from DM annihilation, see
Fig. 1 and 3. If DM annihilation is responsible for most
of the hard X-ray emission in the Ophiuchus cluster,
GLAST will collect such large statistics that discrimina-
tion from a cosmic ray γ-ray emission appears reasonably
feasible.
In summary, we showed that the origin of the non-
thermal particles plausibly responsibly for the recently
firmly discovered non-thermal hard X-ray emission from
the Ophiuchus cluster might be generated by electrons
and positrons produced in WIMP pair annihilations, pro-
vided the rate for the latter is large enough. This sce-
nario is compatible with all observational information on
the cluster, with particle DM production and searches,
and, more importantly, will be thoroughly tested by
GLAST. The future γ-ray telescope might even detect
the monochromatic two-photon emission provided the
particle model has a large enough branching ratio in that
channel. Other galaxy clusters exhibiting non-thermal
activity [17] will also be outstanding sites for GLAST to
look for signatures of WIMP DM pair annihilation [46].
Finally, radio data from the Ophiuchus galaxy cluster can
also be accounted for in the DM annihilation scenario, as
long as the average magnetic field in the Ophiuchus clus-
ter is below ≈ 0.2 µG.
The author wishes to thank Tesla Jeltema for several
insightful comments and for help with the computation
of the thermal X-ray emission.
∗ Electronic address: [email protected]
[1] S. Colafrancesco, S. Profumo and P. Ullio, Astron. As-
trophys. 455 (2006) 21 [arXiv:astro-ph/0507575].
[2] G. Jungman, M. Kamionkowski, and K. Griest, Phys.
Rept. 267, 195 (1996); G. Bertone, D. Hooper, and
J. Silk, Phys. Rept. 405, 279 (2005).
[3] D. Hooper and S. Profumo, Phys. Rept. 453 (2007) 29
[arXiv:hep-ph/0701197].
5
[4] S. Colafrancesco, S. Profumo and P. Ullio, Phys. Rev. D
75 (2007) 023513 [arXiv:astro-ph/0607073].
[5] T. E. Jeltema and S. Profumo, "Searching for Dark Mat-
ter with X-ray Observations of Local Dwarf Galaxies"
arXiv:0805.1054 [astro-ph].
[6] D. Eckert, N. Produit, S. Paltani, A. Neronov and
T. L. Courvoisier, "INTEGRAL discovery of non-
thermal hard X-ray emission from the Ophiuchus cluster"
arXiv:0712.2326 [astro-ph].
[7] M. D. Johnston et al., Ap.J 245 (1981) 799.
[8] B. Y. Mills, O. B. Slee, E. R. Hill, Australian J. Phys., 13,
676 (1960); O. B. Slee, Australian J. Phys., Ap. Suppl.,
43, 1 (1977); O. B. Slee and C. S. Higgins, Australian
J. Phys., Ap. Suppl., 36, 60 (1975); J. R. Ehman, J. R.
Dixon and J. D. Kraus, Ap. J. 75 (1970) 351; B. B. Jones
and E. A. Finley, Australian J. Phys., 27, 687 (1974);
[9] O. Reimer, M. Pohl, P. Sreekumar and J. R. Mattox,
Astrophys. J. 588 (2003) 155 [arXiv:astro-ph/0301362].
[10] L. Feretti and G. Giovannini, arXiv:astro-ph/0703494.
[11] M. Rossetti and S. Molendi, Astron. Astrophys. 414, L41
(2004) [arXiv:astro-ph/0312447].
[12] C. Pfrommer, arXiv:0707.1693 [astro-ph].
[13] R. Fusco-Femiano et al., Astrophys. J. 513, L21 (1999)
[arXiv:astro-ph/9901018].
[14] M. Renaud, G. Belanger, J. Paul, F. Lebrun and R. Ter-
rier, arXiv:astro-ph/0606114.
[15] R. Fusco-Femiano, R. Landi and M. Orlandini, Astro-
phys. J. 624, L69 (2005) [arXiv:astro-ph/0504147].
[16] J. S. Sanders, A. C. Fabian, S. W. Allen and
R. W. Schmidt, Mon. Not. Roy. Astron. Soc. 349 (2004)
952 [arXiv:astro-ph/0311502]; J. S. Sanders, A. C. Fabian
and R. J. H. Dunn, Mon. Not. Roy. Astron. Soc. 360
(2005) 133 [arXiv:astro-ph/0503318].
[17] J. Nevalainen, T. Oosterbroek, M. Bonamente and
S. Colafrancesco, Astrophys. J. 608 (2004) 166
[arXiv:astro-ph/0311142].
[18] T. Sjostrand, S. Mrenna and P. Skands, arXiv:0710.3820
[hep-ph].
[19] J. S. Kaastra and R. Mewe, in Atomic Data Needs for
X-ray Astronomy, p. 161, (2000), ed. by M. A. Bautista,
T. R. Kallman, and A. K. Pradhan.
[20] See e.g. H. P. Nilles, Phys. Rept. 110 (1984) 1.
[21] H. Baer
020
JHEP 0510,
et
al.,
(2005)
[arXiv:hep-ph/0507282].
[22] S. Colafrancesco and P. Blasi, Astropart. Phys. 9 (1998)
227 [arXiv:astro-ph/9804262].
[23] H. A. Mayer-Hasselwander et al., Astron. Astrophys. 335
(1998) 161.
[24] F. Aharonian et al.
[H.E.S.S. Collaboration], Phys.
Rev. Lett. 97, 221102 (2006) [Erratum-ibid. 97, 249901
(2006)] [arXiv:astro-ph/0610509].
[25] J. F. Navarro et al., Mon. Not. Roy. Astron. Soc. 349
(2004) 1039 [arXiv:astro-ph/0311231].
[26] J. S. Bullock et al., Mon. Not. Roy. Astron. Soc. 321,
559 (2001) [arXiv:astro-ph/9908159].
[27] T. H. Reiprich and H. Boringer, Ap.J. 567 (2002) 716.
[28] J. Diemand, M. Zemp, B. Moore, J. Stadel and M. Car-
ollo, Mon. Not. Roy. Astron. Soc. 364 (2005) 665
[arXiv:astro-ph/0504215].
[29] See e.g. S. Profumo and P. Ullio, JCAP 0311 (2003)
006 [arXiv:hep-ph/0309220] and R. Catena et al.,
arXiv:0712.3173 [hep-ph].
[30] A. Cesarini, F. Fucito, A. Lionetto, A. Morselli
267 (2004)
and P. Ullio, Astropart. Phys. 21,
6
[arXiv:astro-ph/0305075].
[31] S. Profumo, Phys. Rev. D 72,
103521
(2005)
[arXiv:astro-ph/0508628].
B. Sadoulet, Phys. Rev. Lett. 74 (1995)
[arXiv:hep-ph/9412213].
5174
[39] S. Profumo and C. E. Yaguna, Phys. Rev. D 70, 095004
[32] N. W. Evans, F. Ferrer and S. Sarkar, Phys. Rev. D 69,
(2004) [arXiv:hep-ph/0407036].
123501 (2004) [arXiv:astro-ph/0311145].
[33] L. Bergstrom and D. Hooper, Phys. Rev. D 73, 063510
[40] http://www-glast.slac.stanford.edu/
[41] L. Bergstrom and P. Ullio, Nucl. Phys. B 504 (1997) 27
(2006) [arXiv:hep-ph/0512317].
[arXiv:hep-ph/9706232].
[34] S. Profumo and M. Kamionkowski, JCAP 0603, 003
[42] F. Ferrer, L. M. Krauss and S. Profumo, Phys. Rev. D
(2006) [arXiv:astro-ph/0601249].
[35] I. V. Moskalenko, S. W. Digel, T. A. Porter, O. Reimer
and A. W. Strong, Nucl. Phys. Proc. Suppl. 173 (2007)
44 [arXiv:astro-ph/0609768].
[36] W. de Boer, C. Sander, V. Zhukov, A. V. Gladyshev
and D. I. Kazakov, Astron. Astrophys. 444 (2005) 51
[arXiv:astro-ph/0508617].
[37] I. Gebauer, arXiv:0710.4966 [astro-ph].
[38] M. Kamionkowski, K. Griest, G. Jungman and
74 (2006) 115007 [arXiv:hep-ph/0609257].
[43] K.-T. Kim, P. P. Kronberg and P. C. Tribble, Ap.J. 379
(1991) 80.
[44] P. Blasi, S. Gabici and G. Brunetti, Int. J. Mod. Phys.
A 22 (2007) 681 [arXiv:astro-ph/0701545].
[45] S. Ando and D. Nagai, arXiv:0705.2588 [astro-ph].
[46] J. Kehayias, T. Jeltema and S. Profumo, in preparation.
|
astro-ph/0311239 | 1 | 0311 | 2003-11-11T01:01:38 | Rest-frame optical continua of L ~ L*, z>3 quasars: probing the faint end of the high z quasar luminosity function | [
"astro-ph"
] | Near-IR photometry for 20 radio-loud z>3 quasars, 16 of which are radio- selected, are presented. These data sample the rest-frame optical/UV continuum, which is commonly interpreted as emission from an accretion disk. In a previous study, we compared the rest-frame optical/UV continuum shapes of 15 optically bright (V<17.5) z>3 quasars with those of 27 low redshift (z~0.1) ones that were matched to the high redshift sample in evolved luminosity (i.e. having luminosities ranging from 1-7 times the characteristic luminosity, L*, where L*~(1+z)^{~3}) to look for signs of evolution in the central engines. We found the continuum shapes at z~0.1 and z>3 similar, consistent with no significant change in the ratio mdot/M, where mdot is the accretion rate with respect to the Eddington rate and M is the black hole mass. This study expands our earlier high redshift sample to lower luminosity, away from extreme objects and towards a luminosity overlap with lower redshift samples. The distribution of rest-frame optical/UV continuum shapes for this fainter sample is broader, extending further to the red than that of the brighter z>3 one. Three quasars from this fainter sample, two radio-selected and one optically-selected, have optical continuum slopes alpha<-1 (F_{nu}~nu^{alpha}). The optically-selected one, LBQS0056+0125, appears to be reddened by dust along the line of sight or in the host galaxy, whereas the radio-selected ones, PKS2215+02 and TXS2358+189, could derive their red continua from the contribution of a relatively strong synchrotron component to the rest-frame optical. These objects may represent a bridge to a population of very red high redshift quasars to which ongoing or future near-IR, optical and deep X-ray surveys will be sensitive. | astro-ph | astro-ph |
Mon. Not. R. Astron. Soc. 000, 000 -- 000 (0000)
Printed 8 March 2018
(MN LATEX style file v2.2)
Rest-frame optical continua of L ∼ L∗, z > 3 quasars:
probing the faint end of the high z quasar luminosity
function
Olga P. Kuhn
Joint Astronomy Centre, 660 N. A'ohoku Place, Hilo, HI, 96720
email: [email protected]
8 March 2018
ABSTRACT
Near-IR photometry for 20 radio-loud z > 3 quasars, 16 of which are radio-selected,
are presented. These data sample the rest-frame optical/UV continuum, which is com-
monly interpreted as emission from an accretion disk. In a previous study, we compared
the rest-frame optical/UV continuum shapes of 15 optically bright (V< 17.5) z > 3
quasars with those of 27 low redshift (z ∼ 0.1) ones that were matched to the high
redshift sample in evolved luminosity (i.e. having luminosities ranging from 1-7 times
∼ (1 + z)∼3) to look for signs of evolution
the characteristic luminosity, L∗, where L∗
in the central engines. We found the continuum shapes at z ∼ 0.1 and z > 3 similar,
consistent with no significant change in the ratio m/M, where m is the accretion rate
with respect to the Eddington rate and M is the black hole mass. This study expands
our earlier high redshift sample to lower luminosity, away from extreme objects and to-
wards a luminosity overlap with lower redshift samples. The distribution of rest-frame
optical/UV continuum shapes for this fainter sample is broader, extending further to
the red than that of the brighter z > 3 one. Three quasars from this fainter sample,
two radio-selected and one optically-selected, have optical continuum slopes α < −1
(Fν ∼ ν α). The optically-selected one, LBQS0056+0125, appears to be reddened by
dust along the line of sight or in the host galaxy, whereas the radio-selected ones,
PKS2215+02 and TXS2358+189, could derive their red continua from the contribu-
tion of a relatively strong synchrotron component to the rest-frame optical. These
objects may represent a bridge to a population of very red high redshift quasars to
which ongoing or future near-IR, optical and deep X-ray surveys will be sensitive.
Key words: quasars: general -- galaxies: evolution -- galaxies: high-redshift
1
INTRODUCTION
Within the past decade, there has been significant progress
towards characterizing the evolution of the quasar luminos-
ity function. Up to redshifts z < 2, it is well described as
luminosity evolution, with the characteristic luminosity L∗
increasing by a factor 40 − 50 from z = 0 to z = 2.3 (Boyle
et al. 2000). Quasar activity peaks around z∼ 2.5 − 3 (e.g.
Schmidt, Schneider & Gunn 1991) but from there to z > 4,
the space density drops by a factor ∼ 2 − 20 (e.g. Jarvis &
Rawlings 2000, Fan et al. 2001a, Vigotti et al. 2003). This
high redshift turnover is identified with the epoch of quasar
formation (Warren, Hewett & Osmer 1994), but the cumu-
lative effects of dust along the line of sight may also play a
role, the significance of which is not yet well determined.
Despite our improved knowledge of the statistical evo-
lution of quasars, a fundamental question, 'how do individ-
ual quasars form and evolve?', remains unanswered. Models
which combine theories of structure formation with those of
energy generation in quasars can reproduce the evolution of
the quasar luminosity function (e.g. Siemiginowska & Elvis
1997; Haehnelt, Natarajan & Rees 1998; Haiman & Menou
2000; Kauffman & Haehnelt 2000; Hatziminaoglou, Siemigi-
nowska & Elvis 2001), but these differ in details so do not
constrain the evolution of the central engine.
The optical/UV continua and emission lines originate
within the central parsec and reflect conditions in the cen-
tral engine. At low redshift, the emission line correlations
involved in Boroson & Green's (1992) eigenvectors 1 and 2
have been well studied, and within the past couple of years
in particular, the physical drivers of these have become bet-
ter understood, enabling them to be used as tracers of the
central mass and accretion rate (Marziani et al. 2001; Boro-
son 2002). The optical/UV continuum is often attributed to
emission from an accretion disk. Model fits can be used to
2 O. P. Kuhn
estimate the mass and accretion rate, though factors such
as intrinsic reddening or additional components, such as
synchrotron emission which may be important in the flat-
spectrum radio quasars (Serjeant & Rawlings 1997; Fran-
cis, Whiting & Webster 2000; Whiting, Webster & Francis
2001), complicate matters. Use of multiple datasets: emis-
sion line, continuum and polarization; should yield the best
estimates of central engine parameters, however, these are
feasible only for the brighter quasars.
In a previous study, we compared the rest-frame opti-
cal/UV continua of 15 z > 3 quasars with those of 27 z ∼ 0.1
quasars matched to the high redshift ones in evolved lumi-
nosity (Kuhn et al. 2001, hereafter Paper I). We found no
evidence for significant evolution of the continuum shape.
The high redshift sample in our previous study was se-
lected nearly a decade ago and was limited. We had aimed
to gather spectroscopic as well as photometric data. This
meant that the objects had to be bright and relatively few.
The resulting set which consisted of the optically brightest
z > 3 quasars was nearly evenly split between radio-loud
and radio-quiet quasars, and several of these were consid-
ered the most luminous objects in the universe when they
were discovered (e.g. HS 1946+7658; Hagen et al. 1992). Fi-
nally, our low-z/high-z sample matching criteria imposed a
strong redshift-luminosity degeneracy.
To move the high-redshift sample away from the most
extremely luminous objects, towards some overlap with the
low-redshift one (and other low or intermediate redshift sam-
ples drawn from the literature), and increase its size, a set
of 20 fainter z > 3 quasars was selected. For the previous
'bright z > 3' sample, a comparison of the optical/UV spec-
tral indices measured using both photometric and spectro-
scopic data with those measured from the photometry alone
shows that photometry alone is adequate at least for sin-
gle power law fits to continuum shape. This paper describes
the observations and distribution of optical/UV continua for
this new 'faint z > 3' sample. Section 2 discusses the sample
selection and section 3, the data acquisition and reduction.
In section 4, the rest-frame optical/UV continua are pre-
sented and, in section 5, compared with those of the 'bright
z > 3' sample. Section 6 discusses the reddest quasars with
α < −1, and conclusions are drawn in section 7. Except
in Figure 1, a Friedmann cosmology with Ho = 75 km s−1
Mpc−1, qo = 0.1 and Λ = 0 is used.
2 SAMPLE
2.1 Selection
To minimize optical selection effects, the 'faint z > 3' sample
was limited to radio-loud quasars. The targets were drawn
from a listing of all objects with z > 3 that were cat-
alogued in NED1 as both quasars and radio sources. To
be accessible from UKIRT and from La Palma (where we
sought optical data), a declination limit: −20◦ < δ < 60◦;
was imposed. A cross-check with the latest edition of the
V´eron-Cetty & V´eron catalogue (2000; VCV9) revealed a
few new quasars; and vice-versa, a few NED quasars were
1 NED, the NASA/IPAC Extragalactic Database
not in VCV9. Quasars appearing in one or the other cata-
logue were kept. Those noted as lensed were removed, and
a final list of 20 radio-loud quasars accessible during the
allocated nights, i.e. with right ascension between 22h and
08h30m, constituted the target sample, hereafter referred to
as z3f (for z > 3 faint; Table 1).
Of the 20 z3f quasars, 16 were radio-selected and 4 were
originally discovered in optical -- objective prism or grism
-- surveys. The radio-selected quasars tend to have high
radio-to-optical flux ratios (Table 1) and flat radio spec-
tra (Table 4). This is probably a consequence of the sur-
vey frequency (most are from 5GHz and 1415MHz surveys,
so sample still higher rest-frame frequencies) and sensitiv-
ity. Deeper and longer wavelength surveys would be more
effective in selecting steep-spectrum high redshift quasars.
This radio-selected subsample avoids the bias towards blue
optical/UV continua that may result from some optical sur-
veys, but on the other hand may include blazars in which
a synchrotron component contributes to and reddens the
optical/UV continua (Whiting et al. 2001). The 4 optically
selected radio-loud quasars were kept so that their colors
could be compared with those of the radio-selected ones.
2.2 Radio-loudnesses
The sample quasars are classified as radio-loud no matter
which criterion -- rest-frame radio-luminosity or radio-to-
optical flux ratio -- is used. Their rest-frame 1.4GHz lu-
minosities are computed using equation 1, from Stern et
al. (2000), where I have assumed the radio spectral index,
αr = −0.5 (Fν ∼ ν αr ), to be consistent with the assump-
tions made by Gregg et al. (1996) who classify as radio-loud
those quasars with log(L1.4GHz)> 32.5 [erg s−1 Hz−1] (for
Ho = 50, qo = 0.5).
L1.4GHz = 4πd2
L
F1.4GHz,obs
(1 + z)(1+αr )
(1)
In the equation, dL is
the luminosity distance and
F1.4GHz,obs(≡ FN V SS) is the observed flux at 1.4 GHz from
the NRAO-VLA Sky Survey (NVSS; Condon et al. 1998).
The values of F1.4GHz,obs and L1.4GHz are listed in the fourth
and fifth columns of Table 1. To compare the luminosities
L1.4GHz with the radio-quiet/radio-loud threshold of Gregg
et al. (i.e. to convert them to the cosmological parameters
that Gregg et al. adopted), a constant which depends on
redshift but ranges from 0.02 to 0.1 for this sample should
be subtracted. All of the z3f quasars have rest-frame 1.4
GHz luminosities above the radio-loud threshold of Gregg
et al (1996). The radio-to-optical flux ratio, RL, is calcu-
lated as in Bechtold et al. (1994), although the rest-frame
5GHz and 1450A luminosities are computed directly from
the radio data available from NED and the NVSS and from
the near-IR data published here:
RL = log10
L5GHz
L
1450A
(2)
Objects with RL > 1 are considered radio-loud. For the
three objects with only one radio measurement, a radio spec-
tral index, αr = −0.18 is assumed. This is the median αr for
the 16 radio-selected quasars. Values of RL for the sample
quasars are listed in column 6 of Table 1.
Rest-frame optical continua of L ∼ L∗ z > 3 quasars
3
2.3 Radio spectral indices
Most of the quasars were classified as flat-spectrum radio-
quasars. The spectra of a couple which had more exten-
sive radio data were determined to be peaked at Giga-Hz or
higher frequencies (as noted in Table 1). One, B30749+420,
was classified as a compact symmetric object (CSO). To
measure their radio spectral indices, I fit a power law to the
radio data taken from NED. The results confirm that none of
the quasars with greater than one radio-flux measurement,
i.e. the 16 radio-selected z3f quasars plus Q2311 − 036, is a
steep-spectrum quasar; all have αr > −0.5 (column 6 of Ta-
ble 4). Radio data at a greater range of wavelengths would
be needed to determine whether any more of these are Giga-
Hz peaked or CSOs. The tendency to find GPS quasars at
high redshift and their interpretation as young radio sources
(O'Dea 1998) suggests that there may be more within this
high redshift sample.
2.4 z−L coverage
A consideration in selecting the sample was to increase the
luminosity range at high redshift and reduce the redshift-
luminosity degeneracy that was inherent in our earlier study.
Figure 1 plots the redshift-absolute magnitude distribution
of the 'faint z > 3' sample, together with the earlier 'bright
z > 3' (hereafter z3b) and low redshift ones.
Cross-correlating the z > 3 quasars in the VCV9 catalog
with the positions of 2MASS sources in the 2nd incremental
data release2 yields 56 with detections at J, H and Ks. The
redshifts and absolute magnitudes of these 2MASS-detected
quasars are also indicated Figure 1. While not homogeneous,
the sample is large and as such provides a reference for the
other z > 3 samples.
3 OBSERVATIONS & DATA REDUCTION
3.1 Near-IR photometry
Near-infrared photometry was obtained for 18 of the z3f tar-
gets in the fall of 2000 using UFTI (UKIRT Fast-Track Im-
ager, Roche et al. 2002) at the United Kingdom Infrared
Telescope (UKIRT). Data for the remaining 2 were obtained
in the spring of 2001 using IRCAM at UKIRT. Deeper im-
ages of the field of PC0027+0525 were obtained in Au-
gust 2003, and magnitudes measured from these rather than
from the November 2000 ones were used in the anaylsis.
Table 2 lists the dates of the observations and results. All
20 quasars were observed through the J, H and K Mauna
Kea Observatory NIR filters (MKO-NIR; Tokunaga, Simons
& Vacca 2002), and 8 were also observed through the Z
(λcen ∼ 0.95µm) filter. Most of the observations were made
on 17 and 18 November (UT), though a few objects were ob-
served in September and October to be quasi-simultaneous
with optical data which had been obtained in late August
(Appendix A). All 20 were re-observed approximately one
year later (October 2001 - July 2002) at K to provide a rough
indication of their variability (see section 3.2).
All of the near-IR observations were carried out in a
2 http://www.ipac.caltech.edu/2mass/releases/second
-32
-30
-28
-26
-24
-22
0
1
2
3
4
5
z
Figure 1. Redshift-absolute magnitude distribution of three
z > 3 datasets considered here: (1) 17 of the 20 target (z3f)
quasars that are listed in, so have absolute magnitudes from,
V´eron-Cetty & V´eron (2001, VCV10; filled boxes); (2) the 15
previously studied bright z > 3 quasars (open triangles); and (3)
56 quasars with JHKs from 2MASS (crosses). For reference, the
redshift-absolute magnitude distribution of the 27 'Atlas' (Elvis
et al. 1994a) quasars used as a comparison sample in Paper I
are also plotted (filled triangles). All the absolute magnitudes are
from VCV10, which assume a k-correction based on a Fν ∼ ν−0.3
approximation to the quasar continuum. The solid curve traces
characteristic luminosity as a function of redshift: L∗ ∼ (1+z)3.34
up to z ∼ 2 (Boyle, Shanks & Peterson 1988); the dashed line
shows a later formulation which takes into account the slowing
of evolution at z ∼ 2 (note that in both cases, the curve is only
an extrapolation for z >∼ 2). Dotted lines mark tracks of con-
stant apparent magnitude. A Friedmann cosmology with Ho=50
km s−1 Mpc−1 and qo = 0 is used for this plot, since that is
what VCV10 adopt when computing absolute magnitudes and
it is one of the cosmologies for which Boyle et al. (1988, 1991)
parameterize luminosity evolution.
grid (or jitter) pattern. At J, H and K, this was either a 5 or
9-point jitter, depending on source brightness, of individual
60-second observations. For the faintest sources, the 9-point
jitter was repeated 2 or 3 times. At Z, longer integrations
were needed to be background limited, and most of the Z
band observations were carried out as 5 or 3-point jitters of
120 or 250-second integrations.
Standard stars from the set of UKIRT faint standards
(Hawarden et al. 2001) were observed approximately every
hour for flux calibration. While nights in September and
October were photometric, the two in November had some
cloud, during the second part of Nov 17 and first part of
Nov 18. Frequent observations of standards enabled the data
from the beginning of Nov 17 to be salvaged, and the two
quasars observed through cirrus at the end of that night
were re-observed with IRCAM in April 2001. Since there
was cirrus at the start of Nov 18, it was treated as non-
photometric and repeat observations to build up signal-to-
noise on several faint objects were made. Later on, skies
cleared and standard star observations looked reliable, so Z
band observations were made of several of the targets.
Reduction of the near-IR data was straightforward up
to the point of doing aperture photometry. The ORAC-
4 O. P. Kuhn
Table 1. Faint z > 3 radio-loud sample
Quasara
PKS 2215+02
MG3 J222537+2040
MG3 J225155+2217
Q2311-036
TXS 2342+342
TXS 2358+189
MG1 J000655+1416
PC0027+0525
LBQS0056+0125
UM672
MG3 J015105+2516
PKS 0201+113
MG3 J023222+2318
MG1J024614+1823
PKS 0335-122
PKS 0336-017
MG2 J062425+3855
Q0642+449
B3 0749+426
PMNJ0833+0959
18.4
20.5
−
mb
21.5h
R18.6
20.2
za
3.55
3.56
3.668
3.034 R18.8
3.053
3.10
3.20
4.099 R21.49
3.149
3.119
3.10
3.61
3.42
3.59
3.442 R20.2
3.197 R18.8
3.469 R18.7
18.49
3.396
R18.3
3.59
3.75
−
18.6
18
R19.8
R19.5
R19.9
−
Fc
N V SS
781
221
190
79
155
266
184
5
6
46
200
781
461
217
476
593
809
453
711
122
1.4GHz
Ld
35.6
35.1
35.1
34.5
34.8
35.0
34.9
33.6
33.4
34.3
34.9
35.7
35.4
35.1
35.4
35.4
35.6
35.4
35.6
34.9
RLe
4.6
3.0
3.9
2.9
2.9
4.5
2.9
3.1
1.9
2.5
3.9
3.9
4.1
3.5
4.5
3.7
4.2
3.8
3.2
3.1
notesg
FSRQ
FSRQ
FSRQ
FSRQ,ALS(1)
FSRQ,ALS(2)
FSRQ,ALS(2)
FSRQ
FSRQ
GPS(5),FSRQ(6),ALS(2)
FSRQ
FSRQ
FSRQ, ALS(4)
radio surveysf
87GB, TXS, WB92, PHJFS
87GB, WB92
87GB, WB92
PMN, opt(UM 659)
87GB, WB92
87GB, WB92
87GB, WB92, TXS
opt(SSG)
opt(SSG,LBQS)
opt(UM)
87GB, WB92
87GB, WB92, TXS
87GB, WB92, TXS
87GB, WB92
PMN, TXS
PMN, 87GB, WB92, TXS, PHJFS GPS(7)
B3, 87GB, WB92, S4, TXS
B3, 87GB, WB92, S4, TXS
87GB, WB92, TXS
87GB, WB92
FSRQ
HFP(8)
CSO(9)
FSRQ
ALS(3)
a Name as it appears in NED: coordinates used were from NED and not repeated here. Redshift z from
NED.
b optical magnitude, at V if not indicated as R, from VCV10, except when otherwise noted.
c FN V SS is the 21-cm flux in mJy from the NRAO-VLA Sky Survey (NVSS; Condon et al. 1998).
d Column lists log10(L1.4GHz ): L1.4GHz is the rest-frame luminosity at 1.4GHz in erg s−1 Hz−1.
e RL is the rest-frame radio(5GHz) to optical(1450A) flux ratio, as defined in Bechtold et al. (1994); RL> 1
is radio-loud. The radio and optical rest-frame fluxes are computed directly for each object from power law
fits to the radio and to the near-IR data.
f Radio and/or optical surveys which detected the quasar (indicated in this column and by the prefix
in column 1): PKS: 2.7 GHz, Parkes radio survey (e.g. Shimmins et al. 1966); PHJFS: PKS Half-Jansky
Flat-Spectrum Sample (Drinkwater et al. 1997); MG: 5 GHz MIT Green Bank 5GHz survey (e.g. Bennett
et al. 1986); 87GB: 5 GHz (Gregory & Condon 1991; Becker, White & Edwards 1991); TXS: 365 MHz
(Douglas et al. 1996); WB92: 1.4 GHz (White & Becker 1992); B3: 408 MHz, 3rd Bologna Catalog of Radio
Sources (Ficarra, Grueff & Tomassetti 1985); S4: 5 GHz, Fourth 'Strong' radio source survey (Pauliny-Toth
et al. 1978); PMN: 5 GHz, Parkes-MIT-NRAO radio survey (Griffith et al. 1994).
opt =discovered in an optical survey: LBQS -- objective-prism-selected (LBQS0056+0125 reported by
Chaffee et al. 1991); UM -- Univ. of Michigan, obj-prism selection (McAlpine & Feldman 1982); SSG --
grism survey (Schmidt, Schneider & Gunn 1987, Schneider, Schmidt & Gunn 1991).
g Notes:
ALS = line-of-sight absorption line system: (1) Lyα+metal system at z ∼ 2.7 (Bechtold 1994); (2) Lyα
absorption systems at high redshift (z >∼ 3) reported by White, Kinney & Becker (1993). The one toward
TXS 2358+189 may not be damped, but those toward TXS 2342+342 and PKS 0201+113 are; (3) z ∼ 2.7
damped Lyα system (Pettini et al. 1997); (4) z ∼ 3.178 damped Lyα system (Ellison et al. 2001).
FSRQ = flat spectrum radio quasar;
GPS = Giga-Hz Peaked: (5) O'Dea, Baum & Stanghellini (1991); (6) Stanghellini et al. (1998); (7) Savage
et al (1990); (8) HFP = High Frequency Peaker (Dallacasa et al. 2000); (9) CSO = Compact Symmetric
Object (Peck & Taylor 2000).
h Optical magnitude from NED
DR pipeline 3 (Currie et al. 1999) was used to process the
images: basically it (1) subtracts a dark, (2) applies a bad-
pixel mask, (3) combines the 3, 5 or 9 frames from each group
to make a flat field, masking the central region where the
object should fall, (4) divides this flat into each observation
of the group and finally (5) shifts and averages these to
produce the final mosaic.
Aperture photometry was done using the IRAF 4 task
digiphot.apphot.phot. The difficulties in aperture pho-
tometry arose primarily because of greater-than-average and
variable seeing. For the standard stars, a 45-pixel radius
aperture was adopted (diameter 8.′′2). The curves of growth
for many of the quasars became noisy at radii larger than
10 or 15 pixels, however, so an aperture correction was nec-
essary. For consistency, all of the quasar data were cor-
3 ORAC-DR is an online Data Reduction Pipeline developed at
the Joint Astronomy Centre by Frossie Economou and Tim Jen-
ness in collaboration with the UK Astronomy Technology Centre
as part of the ORAC project.
4 IRAF is distributed by the National Optical Astronomy Obser-
vatories, which are operated by the Association of Universities for
Research in Astronomy, Inc., under cooperative agreement with
the National Science Foundation.
Rest-frame optical continua of L ∼ L∗ z > 3 quasars
5
Table 2. Near-IR observations
Quasar
PKS2215+02
Za
19.63 ± 0.05
...
Ja
18.64 ± 0.04
...
Ha
18.05 ± 0.03
...
MG3J222537+2040
17.69 ± 0.04
...
17.08 ± 0.03
...
16.70 ± 0.04
...
MG3J225155+2217
Q2311-036
...
...
...
18.02 ± 0.03
...
Ka
17.06 ± 0.02
17.10 ± 0.02
16.13 ± 0.02
16.07 ± 0.03
18.16 ± 0.05
17.98 ± 0.04
16.18 ± 0.04
...
16.23 ± 0.11b
16.34 ± 0.01
16.34 ± 0.02
17.89 ± 0.04
...
18.44 ± 0.06
16.49 ± 0.04
16.48 ± 0.03
19.10 ± 0.10
19.05 ± 0.04
15.66 ± 0.04
15.66 ± 0.06
15.75 ± 0.04
15.74 ± 0.11b
17.78 ± 0.04
17.78 ± 0.04
16.80 ± 0.03
15.65 ± 0.03
17.26 ± 0.03
17.40 ± 0.03
16.30 ± 0.05
16.24 ± 0.06
17.56 ± 0.03
17.75 ± 0.04
16.65 ± 0.03
...
16.52 ± 0.05
16.38 ± 0.03
...
16.81 ± 0.03
15.70 ± 0.03
...
15.75 ± 0.06
...
15.53 ± 0.01d
15.47 ± 0.03
...
16.62 ± 0.02d
16.61 ± 0.02
UT date
2000 Sep 17
2001 Oct 6
2000 Sep 17
2001 Oct 6
2000 Nov 18
2001 Oct 6
2000 Nov 17
2000 Nov 18
2002 Jul 17
2000 Oct 13
2001 Oct 6
2000 Nov 18
2000 Oct 13
2001 Oct 6
2000 Nov 17
2001 Oct 8
2000 Nov 18
2003 Aug 08
2000 Nov 17
2001 Oct 8
2000 Nov 17
2002 Jul 17
2000 Nov 17
2001 Oct 6
2000 Nov 17
2001 Oct 6
2000 Sep 17
2002 Feb 16
2000 Nov 17
2001 Oct 6
2000 Nov 18
2002 Feb 16
2000 Nov 18
2000 Oct 12
2001 Oct 5
2000 Nov 17
2000 Nov 18
2001 Oct 5
2000 Nov 17
2000 Nov 18
2001 Oct 5
2000 Nov 18
2001 Apr 15
2002 Feb 16
2000 Nov 18
2001 Apr 15
2002 Feb 16
comments
Z-band data non-photometric
∆K = +0.04
∆K = −0.06
∆K = −0.18
∆K = +0.05
∆K = 0
scaled to Nov 17
∆K = +0.55
∆K ∼ 0
scaled to Nov 17
c
∆K = 0
∆K ∼ 0
∆K = 0
∆K = −1.15
∆K = +0.14
∆K = −0.06
JHK scaled to Sep 17 & Nov 17
∆K = +0.19
∆K = −0.13
∆K = +0.43
∆K = +0.05
IRCAM
∆K = −0.06
IRCAM
∆K ∼ 0
19.34 ± 0.07
...
18.67 ± 0.06
...
17.48 ± 0.04
...
16.76 ± 0.03
...
...
...
17.26 ± 0.02
...
16.76 ± 0.02
...
19.70 ± 0.06
19.69 ± 0.06
...
18.76 ± 0.04
...
...
17.16 ± 0.03
...
16.98 ± 0.04
...
20.55 ± 0.09
20.46 ± 0.04
17.24 ± 0.05
...
19.81 ± 0.08
...
16.54 ± 0.05
...
16.99 ± 0.04
...
16.41 ± 0.04
...
19.03 ± 0.04
...
18.72 ± 0.05
...
17.95 ± 0.03
...
17.54 ± 0.03
...
18.47 ± 0.03
...
17.94 ± 0.04
...
17.86 ± 0.03
...
17.33 ± 0.03
...
19.50 ± 0.03
...
19.04 ± 0.05
...
18.50 ± 0.03
...
18.01 ± 0.04
...
...
...
18.20 ± 0.03
...
...
17.73 ± 0.06
...
17.16 ± 0.05
...
...
18.24 ± 0.03
...
17.54 ± 0.04
17.56 ± 0.04
...
17.21 ± 0.03
...
...
17.73 ± 0.03
...
17.19 ± 0.02
...
...
...
17.18 ± 0.04
...
16.56 ± 0.05
...
...
...
...
...
16.48 ± 0.02d
16.04 ± 0.01d
...
...
...
...
17.69 ± 0.02d
17.08 ± 0.02d
...
...
...
...
...
...
...
...
...
...
...
...
...
...
...
...
...
...
...
...
...
...
...
...
TXS2342+342
TXS2358+189
MG1J000655+1416
PC0027+0525
LBQS0056+0125
UM672
MG3J015105+2516
PKS0201+113
MG3J023222+2318
MG1J024614+1823
PKS0335-122
PKS0336-017
MG2J062425+3855
Q0642+449
B30749+426
PMNJ0833+0959
a All magnitudes except those obtained with IRCAM and those obtained 2002 July 17 were measured within an aperture of
radius 10 pixels (1.′′8 diameter) and an aperture correction applied to determine the magnitude within the 45 pixel radius
(8.′′19 diameter) aperture used for the standard stars. The quasar magnitudes obtained using UFTI on 2002 July 17 were
measured within apertures of radius, R=20(3.′′6 diameter), and corrected to the R=30(5.′′ 4 diameter) aperture used in
determining the zeropoint. Magnitudes are determined assuming only an extinction correction; a synthetic color correction
to account for the different spectral shapes of Vega and the quasar is incorporated in the zero-magnitude flux.
b 2002 Jul 17 was not strictly photometric and there were not enough stars for relative photometry, but the data appeared
to be of sufficient quality to show that neither of these quasars varied appreciably.
c The quasar magnitudes obtained for PC0027+0525 on 2003 Aug 08 (in sub-half-arcsecond seeing) were measured within
apertures of radius, R=10(1.′′8 diameter), and corrected to the R=30(5.′′4 diameter) aperture used in determining the
zeropoint. Bad pixels were interpolated over before magnitudes were computed; this affected the J-band magnitude. The
quasar was barely detected on the 2000 Nov 17/18 data, though it was clearly seen on the 2003 Aug 08 images. The Nov
17/18 magnitudes are only listed to show their rough consistency with the 2003 magnitudes.
d JHK magnitudes for these two objects were measured with IRCAM. For these observations, aperture corrections were
made from an R=20(3.′′ 25 diameter) to R=40(6.′′ 5 diameter) aperture.
rected from a small aperture of radius 10 pixels (diameter
1.′′8) to the 45 pixels standard star one. The variable seeing
meant that the aperture correction had to be determined
individually for each mosaic. Many fields had only one star
bright enough to use in determining an aperture correction.
In this case, the difference between its instrumental magni-
tudes measured within the R=10 pixel and R=45 pixel aper-
tures was taken to be the aperture correction, AC, and the
quadrature sum of the errors in these instrumental magni-
tudes was adopted as the error in it. A few fields had several
stars bright enough for aperture corrections. For these, the
aperture correction was determined from the star with the
smoothest curve of growth and the error was taken to be the
standard deviation, σN −1, in the measurements. The aper-
ture corrections determined in this way, for each frame, were
reasonable -- not too scattered and generally following the
seeing variations. On a single night with good seeing (∼ 0.′′5,
2000 October 13), average aperture corrections ranged from
−0.18 ± 0.06 at J to −0.12 ± 0.01 at K, while on a night with
poor seeing (0.′′8 to 2′′, 2000 November 17) they ranged from
−0.52±0.17 at J to −0.41±0.11 at K. The error here reflects
the seeing variations and is smaller for an individual frame.
Zero-points, ZP , and extinction coefficients, kext, were
computed for each night at each filter. The quasar magni-
6 O. P. Kuhn
tudes, mQ, were then determined from their instrumental
magnitudes within the R=10 aperture, m10, as:
mQ = m10 − AC − kext(χ − 1) + ZP
(3)
where χ is the airmass. Errors are quadrature sums of the
errors in the instrumental magnitude, aperture correction
and zero point.
3.2 Variability
Prior to these observations, most of the target quasars had
not been extensively observed; in particular, there is little
information on their variability properties. Section 2.3 con-
firms that most of the radio-selected z3f quasars have flat
radio spectra. The set of flat-spectrum radio-loud quasars
includes blazars, which contain a beamed synchrotron com-
ponent that may contribute significantly to the rest-frame
optical (e.g. Serjeant & Rawlings 1997), thus complicating
the use of the rest-frame optical/UV continuum shape in
probing the central mass and accretion rate. Signatures of
beamed emission include rapid, high-amplitude variability
(e.g. Marcha et al. 1996). Through service observations at
UKIRT, all 20 targets were re-observed at K, approximately
one year after the initial run. Conditions were not always
photometric, and differential photometry was done for most
of the fields by comparison with the earlier images. The re-
sulting magnitudes are listed in Table 2. Three quasars var-
ied by at least 0.4 mag: TXS2358+189, MG2 J062425+3855
and PKS0201+113, which brightened by 1.2 mag at K in 11
months. It is likely that beamed synchrotron emission con-
tributes to the rest-frame optical continua of these three,
however, the data do not rule out the possibility of such a
component in the other quasars, even though they varied
less. It should be possible to select radio-quiet samples from
optical multi-color or grism surveys or deep near-IR sur-
veys that are not biased to blue optical/UV continua and
avoid the complications of radio-optical synchrotron contri-
butions.
4 SPECTRAL ENERGY DISTRIBUTIONS
4.1 Magnitude-to-flux conversion
No color-correction was made to the IR magnitudes listed in
Table 2. However, a set of zero-magnitude fluxes which in-
corporate a synthetic color correction were computed. These
were determined to be the fluxes, at the central wavelengths
of each of the IR bands, of a power law: Fν,P L = n ν −0.46
(representative of the shape of the composite spectrum of
657 quasars from the FBQS, Brotherton et al. 2001); which
has been normalized to have the same broad band magni-
tude as Vega, i.e.
n =
R φλFλ,V egadλ
R φλFλ,P Ldλ
,
(4)
where φλ represents the transmission through the filter, mul-
tiplied by the atmospheric transmission and detector re-
sponse when available.
Color-corrected zero-magnitude fluxes were computed
for each of the broadband filters used in UFTI. For these,
φλ incorporates both the filter and atmospheric transmis-
sion functions. Data from two other instruments (Paper I):
the MMT-IR photometer (Rieke 1984) and OSIRIS (De-
Poy et al. 1993) at CTIO; are used to re-compute rest-
frame continuum shapes for the z3b quasars (section 5.2).
For these, φλ was simply assumed to have a tophat shape
starting and ending at the 50% cut-on and cut-off wave-
lengths of the MKO-NIR filters. For each filter and trans-
mission function (UFTI or 'tophat'), the central wavelengths
and color-corrected zero-magnitude fluxes are listed in the
second and fourth columns of Table 3. For comparison, the
zero-magnitude fluxes for Vega (i.e. the flux of Vega at each
of the central wavelengths), are also included in the Table.
The zero-magnitude fluxes were found to depend little on
the shape of φλ, so for simplicity the color-corrected zero-
magnitude fluxes computed for UFTI were used to convert
all the IR magnitudes to fluxes.
4.2 Galactic extinction
The fluxes were corrected for Galactic extinction, assum-
ing the values of E(B-V) from Schlegel, Finkbeiner & Davis
(1998; as reported in NED), and the near-IR and optical
Galactic reddening law of Rieke, Lebofsky & Low (1985)
and Savage & Mathis (1979).
4.3 Spectral energy distributions
For the 20 z3f quasars, rest-frame radio-to-UV spectral en-
ergy distributions (SEDs) were constructed from the data
presented here and from radio data from the literature
(listed in NED). These are plotted in Figure 2.
5 SPECTRAL INDICES
5.1 α for the faint z > 3 sample
Rest-frame optical/UV spectral indices were determined by
fitting a power law through the rest-frame SEDs between
1500 and 5990 A (log ν=14.7-15.3). This wavelength range
was chosen since it includes J, H and K for z = 3 − 4.1, the
range of redshifts in the sample. The power law fits were
made with and without the Z-band point, which was not
obtained for all the objects, but the difference is minimal.
For the analysis, the fits to just J, H and K were used.
The 1500-5990A spectral indices determined from these fits
(with and without the emission line correction made to the
photometry; see Appendix B) are listed in Table 4 and their
distribution is plotted in Figure 3.
5.2 Bright z > 3 sample
5.2.1 α for the bright z > 3 sample
The original aim of this project was to expand, in number
and luminosity range, the set of 15 bright z > 3 quasars for
which Kuhn et al. (2001) constructed rest-frame optical/UV
SEDs (Paper I, table 1). In this section, the SEDs of the new
z3f quasars are compared with those of the bright z3b ones,
to determine whether increasing the number has reduced the
scatter, and whether there is any trend with luminosity. The
Rest-frame optical continua of L ∼ L∗ z > 3 quasars
7
48
46
44
48
42
46
44
42
48
46
44
48
42
46
44
42
48
46
44
42
10
12
14
10
12
14
Figure 2. Rest-frame radio/optical/UV spectral energy distributions (SEDs) of the 20 'faint z > 3' sample quasars. Radio data (open
boxes) are taken from NED, and the optical (stars; see Appendix A) and IR data (filled boxes for the 'original' JHK data; open triangles
for K band 'variability check' measurements and a few J datapoints not simultaneous with those used in the power law fits; and X's for
the Z band points) are from this paper. The mean SED of ∼ 18 low redshift (z ∼ 0.1) radio-loud quasars (dotted line; Elvis et al. 1994a),
normalized to the quasar's rest-frame 4400A luminosity, is overplotted for comparison. Power law fits to the radio continua and to the
rest-frame optical/UV continua (from JHK only) are plotted (solid line segments -- these extend beyond the frequency ranges of the fits
so that they would be easy to see).
8 O. P. Kuhn
48
46
44
48
42
46
44
42
48
46
44
48
42
46
44
42
48
46
44
42
10
12
14
10
12
14
Figure 2 -- Continued
Rest-frame optical continua of L ∼ L∗ z > 3 quasars
9
Table 3. Zero-magnitude fluxes
Filtera
UFTI Z
UFTI MKO-NIR J
'tophat' J (1.17 − 1.33µm)
UFTI MKO-NIR H
'tophat' H (1.49 − 1.78µm)
UFTI MKO-NIR K
'tophat' K (2.03 − 2.37µm)
λb
0
0.9525
1.25
1.25
1.635
1.635
2.20
2.20
Fλ(Vega)c
5.239
2.982
2.982
1.192
1.192
0.4151
0.4151
Fλ(PL)c
6.921
3.040
3.021
1.208
1.203
0.4192
0.4189
a Name of the filter. I did not have filter curves for the MMT IR photometer or OSIRIS so approximate
the filter by a tophat and give the adopted cut-on and cut-off values in the table.
b The central wavelength (A) of the filters used in UFTI. Since these filter curves are close to a tophat
shape and the cut-on and cut-off values are set mainly by the atmosphere, these central wavelengths are
used for the MMT and OSIRIS data as well.
c Zero-magnitude flux [10−10 erg s−1 cm−2 A−1]: the flux, Fλ(λ0), of Vega or of a ν−0.46 power law
normalized to have the same broad band magnitude as Vega.
Table 4. Rest-frame optical and radio spectral indices
Quasar
αa,b
opt
PKS 2215+02
−1.30 ± 0.05
MG3 J222537+2040 −0.21 ± 0.04
MG3 J225155+2217 −0.31 ± 0.13
−0.65 ± 0.05
Q2311-036
−0.08 ± 0.04
TXS 2342+342
TXS 2358+189
−1.37 ± 0.09
MG1 J000655+1416 +0.52 ± 0.08
−0.77 ± 0.09
PC 0027+0525
−1.05 ± 0.10
LBQS 0056+0125
UM 672
−0.50 ± 0.09
MG3 J015105+2516 −0.44 ± 0.09
PKS 0201+113
−0.23 ± 0.07
MG3 J023222+2318 −0.35 ± 0.07
MG1 J024614+1823 −0.82 ± 0.09
−0.67 ± 0.05
PKS 0335-122
PKS 0336-017
+0.08 ± 0.05
MG2 J062425+3855 −0.40 ± 0.04
−0.79 ± 0.06
Q0642+449
−0.11 ± 0.03
B3 0749+426
PMN J0833+0959
−0.38 ± 0.04
La,b
4400A
46.22
46.62
45.85
46.43
46.42
45.70
46.41
45.70
46.62
46.62
45.79
46.36
46.13
46.49
45.96
46.34
46.48
46.72
46.88
46.50
αa,c
opt(corr)
−1.24 ± 0.05
−0.16 ± 0.04
−0.34 ± 0.13
−0.75 ± 0.05
−0.10 ± 0.04
−1.51 ± 0.09
+0.36 ± 0.08
−0.94 ± 0.09
−1.19 ± 0.10
−0.66 ± 0.09
−0.61 ± 0.09
−0.33 ± 0.07
−0.43 ± 0.07
−0.98 ± 0.09
−0.57 ± 0.05
+0.06 ± 0.05
−0.34 ± 0.04
−0.77 ± 0.06
−0.13 ± 0.03
−0.33 ± 0.04
La,c
4400A
(corr)
46.13
46.55
45.77
46.40
46.38
45.66
46.36
45.58
46.57
46.58
45.75
46.29
46.06
46.43
45.90
46.28
46.41
46.65
46.81
46.44
αd
r
−0.45 ± 0.06
−0.35 ± 0.02
−0.02 ± 0.05
−0.30 ± 0.00
−0.20 ± 0.09
−0.05 ± 0.07
−0.44 ± 0.03
−0.18e
−0.18e
−0.18e
+0.22 ± 0.08
+0.25 ± 0.13
−0.08 ± 0.03
+0.05 ± 0.06
−0.38 ± 0.04
−0.39 ± 0.05
−0.31 ± 0.08
+0.46 ± 0.05
−0.22 ± 0.08
−0.16 ± 0.11
Ld
5GHz
45.15
44.52
44.45
43.92
44.27
44.39
44.43
43.22e
42.92e
43.79e
44.32
45.02
44.89
44.48
44.87
44.94
45.30
44.96
44.93
44.34
a Power law fit through log νrest = 14.7 − 15.3 (λrest ∼ 1500 − 5990A) and monochromatic rest-frame luminosity [erg
s−1] of this power law fit at λrest = 4400A.
b No correction has been made for emission lines.
c Emission line contribution has been estimated and removed (see Appendix B).
d Power law fit through log νrest = 9 − 11.5 and luminosity [erg s−1] of this power law fit at νrest = 5 GHz.
e Only one radio data point was found, from NVSS and at 1415MHz. To estimate the rest-frame luminosity at 5GHz,
the median spectral index determined for the 16 radio-selected quasars, αr = −0.18, was assumed.
first step in making the comparison is to re-determine the
spectral indices of the z3b quasars following the same meth-
ods as described in sections 4 and 5.1 for the z3f sample,
namely from the broad band magnitudes alone. Dropping
the requirement for IR spectroscopy adds 2 quasars to the
z3b sample, SP82 1 and Q1442+101, for which we had ob-
tained near-IR magnitudes (table 6 of Paper I) but failed to
obtain spectra. One quasar which was missing photometry
at J (Q0114-089) had to be dropped. This yields a net total
of 16 quasars in the revised z3b sample. Finally, I observed
the lensed quasar, Q1208+101, on 21 May 2001 at UKIRT
and obtained the following magnitudes (for both compo-
nents): Z=16.60 ± 0.01, J=16.09 ± 0.01, H=15.63 ± 0.01 and
K=15.16 ± 0.01. At J and K, these are 0.15 and 0.26 mag-
nitudes brighter than what Kuhn et al. (2001) measured in
1993; the differences are within range of what is observed
for quasar variability (Giveon et al. 1999; Hook et al. 1994
remark on the anomalously high variability of Q0055 − 269,
another quasar from the z3b sample that was discovered in
a similar manner as Q1208 + 101). In this paper, only these
new data are used for Q1208+101.
For the quasars that had been observed with OSIRIS,
one photometric point was sufficient to normalize the entire
1 − 2.5µm cross-dispersed spectrum, so we had not obtained
photometry at all three bands: J, H and K. To derive broad
band magnitudes from the spectra, I integrated over the
product of these with the curves of filter-plus-atmospheric
transmission (tophat with cut-on and cut-off values as listed
in Table 3), first, to determine the scaling factor needed to
match the integral with the broad band photometry (typ-
10 O. P. Kuhn
6
4
2
0
0.5
0
-0.5
-1
-1
-0.5
0
0.5
Figure 4. Comparison of the optical spectral indices measured
from fits to broad band photometry, with (open triangles and dot-
ted error bars) and without (filled squares and solid error bars)
a correction for emission line contributions, against those deter-
mined from narrow band monochromatic luminosities (Paper I,
Table 15). The line, α(photometry) = α(narrowband), is drawn
for reference.
Figure 4, making the emission line correction (Appendix B)
improves the agreement for some objects but worsens it for
others.
5.3 Comparing the bright and faint
As the histograms in Figure 3 show, the z3f sample includes
redder quasars than the z3b one. Three of the z3f quasars
have spectral indices α < −1 and are "red quasars" accord-
ing to the definition of Gregg et al. (2002). A Kolmogorov-
Smirnov (K-S) test gives a probability of 5% that the dis-
tributions for the z3b and z3f samples are drawn from the
same parent population. A Student's t-test shows that the
mean spectral indices for each sample differ significantly --
the probability of randomly obtaining a separation as large
as measured, P (t), is 1.4%.
A number of explanations can be put forth:
First, the z3f sample contains only radio-loud quasars
whereas the z3b sample is evenly divided between radio-quiet
and radio-loud. Several of the z3f quasars varied, which sug-
gests that they may be blazars in which a non-thermal con-
tribution to the rest-frame optical could redden their rest-
frame optical/UV continua (Whiting et al. 2001). If the two
radio-selected quasars in the red tail are removed from the
z3f sample, the K-S probability that the bright and faint
samples are drawn from the same parent population in-
creases to 13% -- there are no longer sufficient objects in
the tail to render the distributions of spectral indices signif-
icantly different. However the mean spectral indices for each
sample are still significantly different (P (t) = 4%).
Second, dust reddens and extincts the emitted contin-
uum, so in the optical/UV, redder objects would be expected
to be fainter. Richards et al. (2001) note this trend in the
sample of ∼ 2600 SDSS quasars for which they present op-
tical photometry, and it is evident in Figure 5 which plots
spectral index against optical luminosity for the faint and
bright z > 3 samples as well as the 56 z > 3 quasars with
Figure 3. Distributions of rest-frame optical spectral indices for
the 20 z3f quasars (top) and 16 z3b quasars (bottom). Superposed
on these are the distributions for the 16 radio-selected z3f quasars
(top; right slanted thick hash marks) and the 8 radio-loud z3b
ones (bottom; left slanted thick hash marks). Note the red tail
to the z3f sample, with 3 of the quasars, 2 radio-selected and 1
optically-selected, having α < −1.
ically at H), and second, to determine magnitudes in the
bands for which we had not obtained photometry.
The near-IR magnitudes, determined from the OSIRIS
spectra or directly, for all 16 z3b quasars were processed in
the same way as described above for the z3f quasars: they
were converted to fluxes (section 4.1), dereddened by the
Galactic value (section 4.2), corrected for emission line con-
tributions (Appendix B) and blueshifted to the rest-frame
(section 4.3).
5.2.2 Broad-band photometry vs. Narrow band
luminosities
To confirm the validity of using broad band magnitudes to
measure rest-frame optical spectral indices, for 14 of the 15
bright z > 3 quasars studied in Paper I (all but Q0114−089),
the optical spectral indices computed from the photometry
alone (this time using the zero-magnitude fluxes listed in
Table 9 of Paper I) were compared with those determined
by fitting a line through a set of narrow band monochro-
matic luminosities (Paper I, Table 15) which covered ap-
proximately the same rest-frame region as did J, H and K:
i.e. 3023A, 4200A, 4750A and 5100A for z = 3 − 3.5; 2660A,
3023A, 4200A and 4750A for z = 3.6 − 3.8; and 2500A,
2660A, 3023A and 4200A for z > 4. Figure 4 compares the
optical spectral indices determined by these two methods
and shows that, while not in perfect agreement, the dis-
tributions of spectral indices computed by fitting through
narrow-band line-free windows and through only broad band
photometry do not differ significantly. As can be seen in
Rest-frame optical continua of L ∼ L∗ z > 3 quasars
11
Figure 5. Rest-frame optical spectral index, α, vs. luminosity for several samples: the 20 z3f sample quasars (filled boxes); the 16 quasars
in the z3b sample (filled triangles; the radio-loud ones are circled); and 56 z > 3 quasars, 17 radio-loud (open boxes) and 39 radio-quiet
(crosses) from VCV9 for which J, H and Ks magnitudes were published in the 2nd incremental data release of the 2MASS PSC. Starting
at α = −0.46, reddening vectors are plotted for E(B-V) = 0 to 0.2 in steps of 0.05 and assuming both Galactic and Small Magellanic
Cloud extinction laws.
J, H and Ks magnitudes from the 2MASS 2nd incremental
data release.
Finally, the redder colors of the z3f with respect to
the z3b quasars could represent an intrinsic difference in
their central engines. The spectrum emitted by a thin ac-
cretion disk is expected to peak at a frequency deter-
mined by the apparent disk temperature, which for op-
tically thick/geometrically thin disks follows the relation:
log T ∼ 1
− 2.4(cos θ − 1) where m is the accretion
rate relative to the Eddington limit, M is the black hole
mass, and θ is the inclination of the disk axis to the line
of sight (Sun & Malkan 1989, McDowell et al. 1991). The
thermal disk spectra emitted by the fainter set would be sys-
tematically redder if, for example, both sets of quasars had
similar central masses, but the fainter ones were accreting
at lower rates.
4 log m
M
The importance of dust at high redshift (e.g. Andreani,
Franceschini & Granato 1999; Warren, Hewett & Foltz
2000), together with the possibility of a significant non-
thermal component in these radio-loud quasars, leads to
the suspicion that extrinsic factors are responsible for the
on-average redder colors of these z3f quasars. These com-
plicate the use of optical/UV continua to probe the central
engines and point to the need for a similar study of radio-
quiet quasars and for spectroscopic data.
6 RED QUASARS
Three of the sample quasars have α < −1; two of these,
PKS2215+02 and TXS2358+189, are radio-selected, and
one, LBQS 0056+0125, was discovered in a grism survey
(PC0056+0125, Schmidt, Schneider & Gunn 1987).
6.1 LBQS 0056+0125
LBQS0056+0125 has an optical spectral index α = −1.05
(Table 4). In the discovery paper, Schmidt, Schneider &
Gunn (1987) remarked on the presence of narrow absorp-
tion lines redwards of Lyα. Since then several line-of-sight
Lyα absorption line systems have been detected; one at
z = 2.7771 (Schneider, Schmidt & Gunn 1991; Pettini et
al. 1997) is damped (DLA). To redden an average quasar
continuum (the FBQS composite) emitted at the quasar
12 O. P. Kuhn
redshift, z = 3.149, to α = −1.05, requires an extinction
E(B-V) ∼ 0.12 in the DLA, assuming the Small Magellanic
Cloud (SMC) extinction law (Pr´evot et al. 1984, Bouchet et
al. 1985). Pettini et al. (1997) measure the neutral hydro-
gen column for this DLA to be log(NHI ) = 21.11 ± 0.07.
The above estimate for E(B-V) is consistent with this if
the gas-to-dust ratio in the DLA is approximately 2 times
that in the Milky Way(MW; 4.8 × 1021 cm−2 s−1, Savage &
Mathis 1979) -- a factor ∼ 5 − 15 times lower (dustier) than
existing data imply for DLAs (Pei, Fall & Bechtold 1991,
Pettini et al. 1997). The above assumes no intrinsic extinc-
tion. If the SMC law is used also to describe the extinction
in the host galaxy, and the extinction in the DLA is fixed at
E(B-V)=0.027, what is implied by log(NHI ) = 21.11 and a
gas-to-dust ratio 10 times that in the MW, then an emitted
continuum (FBQS composite) would need to undergo an in-
trinsic extinction, E(B-V)∼ 0.1 mag, which corresponds to
an intrinsic column density NHI ∼ 5 × 1020f cm−2 where
f is the ratio of the gas-to-dust ratio in the quasar host to
that in the MW. This is larger than the column densities
measured for low redshift optically selected quasars (Laor
et al. 1997), but consistent with the values inferred for sets
of X-ray selected (Puchnarewicz et al. 1996) and high red-
shift radio-loud quasars (e.g. Elvis et al. 1994b). So the red
color of LBQS0056+0125 can be explained by intrinsic and
line-of-sight extinction, but with gas-to-dust ratios that are
on the low side for the intervening absorbers and quasar host
galaxies. The emitted continuum may itself be redder than
the FBQS composite. Also, the gas-to-dust ratios in DLAs
do show a lot of scatter (Pettini et al. 1997); this might also
explain why other quasars from the z3f sample with high
redshift DLAs (e.g. TXS 2342+342 and PKS0201+113) do
not also appear very red. Warren et al. (2001) have recently
imaged the field of LBQS0056+0125 with NICMOS(NIC2)
on HST and detect two possible candidates for the galaxy
counterpart to the DLA.
PKS2215+02 was detected in a 14.7 ksec observation with
the ROSAT HRI, with an unabsorbed 0.1 − 2.4 keV flux
(observed frame) equal to 3.2×10−13 erg s−1 cm−2 (Siebert
& Brinkmann 1998). Siebert & Brinkmann noted that this
quasar is optically very faint and has the highest X-ray-
to-optical and radio-to-optical flux ratios of the radio-loud
z > 3 quasars they studied. They suggest that it is heav-
ily absorbed although note that in the optical/UV, PKS
2215+02 does not stand out as redder than the other Parkes
Half Jansky quasars that Francis et al. (2001) studied. The
rest-frame optical continuum presented here is slightly red-
der than but not inconsistent with that of Francis et al.
(2001). Ellison et al. (2001) find no evidence for a damped
Lyα system along the line-of-sight to PKS 2215+02, so if
the red color derives from extinction, this must occur at the
quasar. The data presented here, which show the quasar
as one of the strongest radio emitters of the z3f sample
and as having one of the highest radio-to-optical flux ratios
(RL = 4.6; Table 1), together with evidence for variability
at 5GHz (Siebert & Brinkmann 1998) and in the rest-frame
optical, favor a non-thermal contribution rather than ex-
tinction as the cause of the red color. Further monitoring,
or other tests for blazar-like activity, are needed to confirm
this hypothesis.
TXS2358+189 (z = 3.10), like PKS2215+02, has a very
high radio-to-optical flux ratio, RL = 4.5. It has a flat ra-
dio spectrum, but insufficient data to measure radio vari-
ability. At K, TXS2358+189 dimmed by 0.55 magnitudes
from September 2000 to October 2001; this was the second-
largest one-year variation seen in the sample and suggests
the presence of beamed synchrotron emission. White, Kin-
ney & Becker (1993) find evidence in its spectrum for a Lyα
absorption system, though the data do not conclusively show
whether it is damped. In sum, it is probable that the red
optical continua of both TXS2358+189 and PKS2215+02
derive from the blending of a synchrotron component.
6.2 PKS2215+02 and TXS2358+189
6.3 Characteristics of the α < −1 quasars
The other two quasars with α < −1 are radio-selected. They
have some of the the highest radio-to-optical flux ratios of
the z3f sample (see Table 1). PKS2215+02 belongs to the
Parkes Half Jansky sample (Drinkwater et al. 1997), which
was found to contain sufficient numbers of red quasars to
lead to the suggestion, later proved well-founded at least
for low redshift AGN (Cutri et al. 2000), that optical sur-
veys were missing a large fraction of quasars (Webster et
al. 1995). The Parkes Half Jansky sample has been well
studied in the optical and IR (Francis et al. 2000; Whiting
et al. 2001) and X-ray (Siebert et al. 1998). Francis et al.
(2000) measured the following magnitudes for PKS2215+02
in early September 1997: B=21.84 ± 0.31; V=20.42 ± 0.10;
R=20.14 ± 0.12; I=20.00 ± 0.20; J=19.20 ± 0.50; H=18.21 ±
0.29; Kn=19.34 ± 1.78. These are consistent with the mea-
surements presented here, which were made 3 years later,
except at J and K5, although they quote a large uncertainty
in the Kn magnitude. Between September 2000 and Octo-
ber 2001 the quasar did not vary at K. In 1995 May/June,
5 Though Francis et al. used the Kn filter, they normalized their
magnitudes to the K band zero point.
It is not too surprising to find several red quasars among the
z3f sample. First, if the z = 2.5− ∼ 4 decline in quasar space
density signals the formation epoch of quasars, then some of
the z3f quasars may be intrinsically young. If a ULIRG (e.g.
Sanders et al. 1989) or a dust-enshrouded phase (e.g. Egami
et al. 1996; Fabian 1999) is the first stage in the evolution
of quasars, then young quasars might be expected to be red
(but see Yu & Tremaine 2002). Second, the largely radio-
selected sample included objects with high radio-to-optical
flux ratios and with flat radio spectra, several of which, from
their variability, appear to be blazars and thus may be red-
dened by a non-thermal contribution to the rest-frame opti-
cal (Whiting et al. 2001). And third, the probability of DLAs
and Lyα systems along the line of sight increases with red-
shift.
The 3 α < −1 z3f quasars can be interpreted as red-
dened by line-of-sight dust or as containing a strong syn-
chrotron component. Ongoing and planned optical, near-IR
and deep X-ray surveys will be more sensitive to reddened
quasars than previous surveys; here we compare the prop-
erties of these 3 α < −1 quasars with the limits of such
surveys as the SDSS, WFCAM/UKIDSS and Chandra deep
fields.
Rest-frame optical continua of L ∼ L∗ z > 3 quasars
13
The 3 α < −1 z3f quasars are relatively faint:
LBQS0056+0125 has r=18.9 (Schneider et al. 1991) and
K=15.7; PKS2215+02 has V=20.6 and K=17.1 and
TXS2358+189 has V∼ 21.3 and K=18.4.
From the SDSS commissioning data, Fan et al. (2001b)
selected a uniform sample of high redshift quasars, setting
a flux limit at i∗ ∼ 20; this would have missed one or two of
the α < −1 z3f quasars. In the near-IR, the LAS (large area
survey), which should complement the SDSS, will reach to
K=18.4 (http://www.ukidss.org/surveys/surveys.html) so
would have picked up all (TXS2358+189 lies at its limit).
The Chandra deep (∼ 0.1 − 1 Ms) fields (Barger et al. 2001;
Hornschemeier et al. 2001; Giacconi et al. 2002) reach down
to flux levels of order 6 × 10−17 to 4 × 10−15 erg s−1 cm−2
(2 − 10 keV). Few of the z3f quasars have X-ray data, but
the two detected by ROSAT-HRI have Fx ∼ 3 − 6 × 10−13
erg s−1 cm−2 and the one upper limit was consistent with a
flux, Fx ∼ 10−14 erg s−1 cm−2 (Siebert & Brinkmann 1998).
The unabsorbed 0.5 − 2 keV (observed frame) X-ray fluxes
for a set of z > 4 quasars range from 1 − 30 × 10−15 erg
s−1 cm−2 (Vignali et al. 2001), so future deep X-ray surveys
should turn up more quasars similar to those in the z3f
sample.
7 CONCLUSIONS
This paper presents near-IR photometry for 20 radio-loud
(16 radio-selected) z > 3 quasars.
Some of the targets had not previously been extensively
observed -- NED lists references only to the radio surveys
in which they were detected and the source of the optical
identification -- so the photometry presented here are the
first for these.
While the underlying aim is to look for evolution in the
central engines, uncertainties due to extinction and the pos-
sible contribution of a non-thermal component complicate
efforts to estimate key central engine parameters from these
data alone.
The main conclusions that can be drawn from this
dataset are:
(i) First, the continuum shapes of faint radio-loud z > 3
quasars show more scatter than those of the bright z > 3
quasars from Paper I. Choosing objects with luminosities
close to L∗, rather than the extremely high luminosity
quasars of the bright z > 3 sample, has not led to a conver-
gence in the optical/UV continuum shapes of z > 3 quasars.
(ii) And second, the distribution of continuum shapes of
these faint radio-loud z > 3 quasars is slightly redder than
that of the bright z > 3 ones, which were a mix of radio-loud
and radio-quiet quasars, with 3 having α < −1. These 3 are
not all radio-selected, so while a synchrotron contribution
cannot be ruled out as a cause for the red colors of the radio-
selected quasars, the red optically-selected quasar probably
suffers intrinsic or line-of-sight reddening. Study of a care-
fully selected set of high redshift radio-quiet quasars would
eliminate the possibly of synchrotron reddening. Near-IR
and optical spectroscopy would complement the more easily
obtained photometry, enabling measures of the central mass
via the Hβ (e.g. Laor 1998), MgII (McLure & Jarvis 2002)
or CIV (Vestergaard 2002) emission line widths, and an es-
timate of intrinsic reddening via the Lyα/Hβ ratio (Netzer
et al. 1995, Bechtold et al. 1997).
The 3 α < −1 quasars may represent a bridge to a larger
population of high redshift quasars that are either intrinsi-
cally reddened or reddened by line-of-sight systems (Warren
et al. 2000), the discovery of which is anticipated from on-
going and future optical, near-IR, and deep X-ray surveys.
ACKNOWLEDGMENTS
The bulk of the data presented here were obtained at
UKIRT, the United Kingdom Infrared Telescope, which is
operated by the Joint Astronomy Centre on behalf on the
U.K. Particle Physics and Astronomy Research Council.
Some of the data were obtained as part of the UKIRT Service
Programme. I thank the staff at UKIRT for their help with
obtaining and reducing the data, in particular Andy Adam-
son, Malcolm Currie, Chris Davis, Tom Kerr, Sandy Leggett
and Thor Wold and John Davies for obtaining the optical
photometry presented here. Also I thank the referee, Dr.
Stephen Warren, for comments which helped me to improve
the paper. Construction of the optical/UV continua for the
bright z > 3 dataset used a large suite of instruments and
telescopes, as described in Paper I. One of the instruments
mentioned in this paper, OSIRIS, is the Ohio State Infrared
Imager/Spectrometer, and was built using funds from NSF
award AST 90-161112 and AST 92-18449 by the OSU As-
tronomical Instrumentation Facility. This publication makes
use of data products from the Two Micron All Sky Survey,
which is a joint project of the University of Massachusetts
and the Infrared Processing and Analysis Center/California
Institute of Technology, funded by the National Aeronautics
and Space Administration and the National Science Foun-
dation. This research has also made use of the NASA/IPAC
Extragalactic Database (NED) which is operated by the Jet
Propulsion Laboratory, California Institute of Technology,
under contract with the National Aeronautics and Space
Administration.
REFERENCES
Allen's Astrophysical Quantities, 4th Edition, 2000, ed. A.
N. Cox, AIP Press, New York.
Andreani, P., Franceschini, A., Granato, G. 1999, MNRAS,
306, 161.
Barger, A. J., Cowie, L. L., Steffen, A. T., Hornschemeier,
A. E., Brandt, W. N., Garmire, G. P.. 2001, ApJ, 560, 23.
Bechtold, J. 1994, ApJS, 91, 1.
Bechtold, J., et al. 1994, AJ, 108, 759.
Bechtold, J., Shields, J., Rieke, M., Ji. P., Scott, J., Kuhn,
O., Elvis, M., Elston, R. 1997, in Emission Lines in Active
Galaxies: New Methods and Techniques, ASP Conference
Series, Vol 113, eds. B. M. Peterson, F.-Z. Cheng and A.
S. Wilson, Astron Soc. Pac., San Francisco, p. 122.
Becker, R. H., White, R. L., Edwards, A. L. 1991, ApJS,
75, 1.
Bennett, C.L., Lawrence, C.R., Burke, B.F., Hewitt J.N.,
Mahoney, J. 1986 ApJS, 61, 1.
Boroson, T. A., Green, R. F. 1992, ApJS, 80, 190.
Boroson, T. A. 2002, ApJ, 565, 78.
14 O. P. Kuhn
Bouchet, P., Lequeux, J., Maurice, E., Pr´evot, L., Pr´evot-
Burnichon, M . L. 1985, A&A, 149, 330.
Boyle, B. J., Shanks, T., Peterson, B. A. 1988, MNRAS,
235, 935.
Boyle B. J., Jones L.R., Shanks T., Marano B., Zitelli V.,
Zamorani G., 1991, in The Space Distribution of Quasars,
ASP Conference Series, Vol 21, ed. D. Crampton, Astron
Soc. Pac., San Francisco, p. 191.
Boyle, B. J., Shanks, T., Croom, S. M., Smith, R. J., Miller,
L., Loaring, N., Heymans, C. 2000, MNRAS, 317, 1014.
Brotherton, M. S., Tran, H. D., Becker, R. H., Gregg, M.
D., Laurent-Muehleisen, S. A., White, R. L. 2001, ApJ,
546, 775.
Chaffee, F. H., et al. 1991, AJ, 102, 461.
Condon, J. J., Cotton, W. D., Greisen, E. W., Yin, Q. F.,
Perley, R. A., Taylor, G. B., Broderick, J. J. 1998, AJ, 115,
1693.
Currie, M. J., Wright, G. S., Bridger, A., Economou, F.
1999, in Astronomical Data Analysis Software and Sys-
tems VIII, ASP Conference Series, Vol 172, eds. D. M.
Mehringer, R. L. Plante, D. A. Roberts, Astron Soc. Pac.,
San Francisco, p. 175.
Cutri, R. M., Nelson, B. O., Kirkpatrick, J. D., Huchra, J.
P., Smith, P. S. 2000, in The New Era of Wide-Field As-
tronomy, ASP Conference Series, Vol 232, eds. R. Clowes,
A. Adamson, G. Bromage, Astron Soc. Pac., San Francisco,
p. 78.
Dallacasa, D. Stanghellini, C., Centonza, M., Fanti, R.
2000, A&A, 363, 887.
DePoy, D. L., Atwood, B., Byard, P., Frogel, J., O'Brien,
T. 1993, Proc. SPIE 1946, 667.
Douglas, J. N., Bash, F. N., Bozyan, F. A., Torrence, G.
W., Wolfe, C. 1996, AJ, 111, 1945.
Drinkwater, M. J., et al. 1997, MNRAS, 284, 85.
Egami, E., Iwamuro, F., Maihara, T., Oya, S., Cowie, L.
L. 1996, AJ, 112, 73.
Ellison, S. L., Yan, L., Hook, I. M., Pettini, M., Wall, J.
V., Shaver, P. 2001, A&A, 379, 393.
Elvis, M., et al. 1994a, ApJS, 95, 1.
Elvis, M., Fiore, F., Wilkes, B. J., McDowell, J. C. 1994b,
ApJ, 422, 60.
Fabian, A. C. 1999, MNRAS, 308, L39.
Fan, X. et al. 2001a, AJ, 121, 54.
Fan, X. et al. 2001b, AJ, 121, 31.
Ficarra, A., Grueff, G., Tomassetti, G. 1985, A&AS, 59,
255.
Forster, K., Green, P. J., Aldcroft, T. L., Vestergaard, M.,
Foltz, C. B., Hewett, P. C. 2001, ApJS, 134, 35.
Francis, P. J., Whiting, M. T., Webster, R. L., 2000, PASA,
53, 56.
Giacconi, R., et al. 2002, ApJS, 139, 369.
Giveon, U., Maoz, D., Kaspi, S., Netzer, H., Smith, P. S.
1999, MNRAS, 306, 637.
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., Lacy, M., White, R. L., Glikman, E.,
Helfand, D., Becker, R. H., Brotherton, M. S. 2002, AJ,
564, 133.
Gregory, P. C., Condon, J. J. 1991, ApJS, 75, 1011.
Griffith, M. R., Wright, A. E., Burke, B. F., Ekers, R. D.
1994, ApJS, 90, 179.
Haehnelt, M. G. , Natarajan, P., Rees, M. J. 1998, MNRAS,
T. G.,Leggett,
300, 817.
Hagen, H.-J. et al. 1992, A&A, 253, L5.
Haiman, Z.,Menou, K. 2000, ApJ, 531, 42.
Hatziminaoglou, E.,Siemiginowska, A.,Elvis, M. 2001, ApJ,
547, 90.
Hawarden,
S. K.,Letawsky, M.
B.,Ballantyne, D. R.,Casali, M. M. 2001, MNRAS,
325, 563.
Hook,I.M., McMahon,R.G., Boyle,B.J., Irwin,M.J. 1994,
MNRAS, 268, 305.
Hornschemeier, A. E., et al. 2001, ApJ, 554, 742.
Jarvis, M. J., Rawlings, S. 2000, MNRAS, 319, 121.
Kauffmann, G., Haehnelt, M. 2000, MNRAS, 311, 576.
Kuhn, O., Bechtold, J., Elvis, M., Elston, R. 2001, ApJS,
136, 225. (Paper I)
Landolt, A. 1992, AJ, 104, 340.
Laor, A., Fiore, F., Elvis, M., Wilkes, B. J., McDowell, J.
C. 1997, ApJ, 477, 93.
Laor, A. 1998, ApJ, 505, L83.
McAlpine, G. M., Feldman, F. R. 1982, ApJ, 261, 412.
McDowell, J., Kuhn, O., Elvis, M., Wilkes, B. 1991, in
"Structure and Emission Properties of Accretion Disks",
Proc. of the Sixth IAP Astrophysics Meeting/IAU Collo-
quium No. 129, eds. C. Bertout, S. Collin-Souffrin, J. P.
Lasota, J. Tran Thanh Van, Gif sur Yvette Cedex, France:
Editions Fronti`eres, 473.
McIntosh, D. H., Rieke, M. J., Rix, H.-W., Foltz, C. B.,
Weymann, R. J. 1999, ApJ, 514, 40.
McLure, R. J., Jarvis, M. J. 2002, MNRAS, 337, 109.
Marcha, M. J. M., Browne, I. W. A., Impey, C. D., Smith,
P. S. 1996, MNRAS, 281, 425.
Marziani, P., Sulentic, J. W., Zwitter, T, Dultzin-Hacyan,
D., Calvani, M. 2001, ApJ, 558, 553.
Netzer, H., Brotherton, M. S., Wills, B. J., Han, M., Wills,
D., Baldwin, J. A., Ferland, G. J., Browne, I. W. A. 1995,
ApJ, 448, 27.
O'Dea, C. P., Baum, S. A., Stanghellini, C. 1991, ApJ, 380,
660.
O'Dea, C. P. 1998, PASP, 110, 493.
Pauliny-Toth, I. I. K., Witzel, A., Preuss, E., Kuhr, H.,
Kellerman, K. I., Fomalont, E. B. 1978, AJ, 451.
Peck, A. B., Taylor, G. B. 2000, ApJ, 534, 90.
Pei, Y. C., Fall, S. M., Bechtold, J. 1991, ApJ, 378, 6.
Pettini, M.. King, D. L., Smith, L. J., Hunstead, R. W.
1997, ApJ, 478, 536.
Pr´evot, M. L., Lequeux, J., Maurice, E., Pr´evot, L., Rocca-
Volmerange, B. 1984, A&A, 132, 389.
Puchnarewicz, E. M., et al. 1996, MNRAS, 315, 1.
Richards, G. T., et al. 2001, AJ, 121, 2308.
Rieke, G. H. 1984, in MMTO Visiting Astronomer Infor-
mation, Multiple Mirror Telescope Observatory Technical
Report No. 13.
Rieke, G. H., Lebofsky, M. J., Low, F. J. 1985, AJ 90, 900.
Roche P.F. et al. 2002, Proc SPIE 4841, Instrument Design
and Performance for Optical/IR Ground-Based Telescopes,
eds. M Iye and A.F Moorwood.
Sanders, D. B., Phinney, E. S., Neugebauer, G., Soifer, B.
T., Matthews, K. 1989, ApJ, 347, 29.
Savage, A., Jauncey, D. L., White, G. L., Peterson, B. A.,
Peters, W. L., Gulkis, S., Condon, J. J. 1990, AuJPh, 43,
241.
Savage, B. D., Mathis, J. S. 1979, ARA&A, 17, 73.
Rest-frame optical continua of L ∼ L∗ z > 3 quasars
15
Schlegel, D. J., Finkbeiner, D. P., Davis, M. 1998, ApJ,
500, 525.
Schmidt, M. A., Schneider, D. P., Gunn, J. E. 1987, ApJ,
316, L1.
Schmidt, M. A., Schneider, D. P., Gunn, J. E. 1991, in The
Space Distribution of Quasars, ASP Conf Series, Vol 21,
ed. D. Crampton, Astron Soc. Pac., San Francisco, p. 110.
Schneider, D. P., Schmidt, M. A., Gunn, J. E. 1991, AJ,
101, 2004.
Schneider, D. P., Schmidt, M. A., Gunn, J. E. 1997, AJ,
114, 36.
Serjeant, S., Rawlings, S. 1997, Nature, 379, 304.
Shimmins, A. J., Day. G. A., Ekers, R. D., Cole, D. J. 1966,
AuJPh, 19, 837.
Siebert, J., Brinkmann, W. 1998, A&A, 333, 63.
Siebert, J., Brinkmann, W., Drinkwater, M. J., Yuan, W.,
Francis, P. J., Peterson, B. A., Webster, R. L. 1998, MN-
RAS, 301, 261.
Siemiginowska, A., Elvis, M. 1997, ApJ, 482, L9.
Stanghellini, C., O'Dea, C. P., Dallacasa, D., Baum, S. A.,
Fanti, R., Fanti, C. 1998, A&AS, 131, 303.
Stern, D.,Djorgovski, S. G., Perley, R. A., de Carvalho, R.
R., Wall, J. V. 2000, AJ, 119, 1526.
Sun, W.-H., Malkan, M. A. 1989, ApJ, 346, 68.
Tokunaga, A. T., Simons, D. A., Vacca, W. D. 2002, PASP,
114, 180.
V´eron-Cetty, M.-P., V´eron, P. 2000, ESO Sci Rep, 19, 1
(VCV9).
V´eron-Cetty, M.-P., V´eron, P. 2001, A&A, 374, 92
(VCV10).
Vestergaard, M. 2002, ApJ, 571, 733.
Vignali, C. Brandt, W. N., Fan, X., Gunn, J. E., Kaspi, S.,
Schneider, D. P., Strauss, Michael A. 2001, AJ, 122, 2143.
Vigotti, M., Carballo, R., Benn, C. R., DeZotti, G., Fanti,
R., Gonzalez Serrano, J. I., Mack, K.-H., Holt, J. 2003,
ApJ, 591, 43.
Warren, S. J., Hewett, P. C., Osmer, P. J. 1994, ApJ, 421,
412.
Warren, S. J., Hewett, P. C., Foltz, C. B. 2000, MNRAS,
312, 827.
Warren, S. J., Møller, P., Fall, S. M., Jakobsen, P. 2001,
MNRAS, 326, 759.
Webster, R. L., Francis, P. J., Peterson, B. A., Drinkwater,
M. J., Masci, F. J. 1995, Nature, 375, 469.
White, R. L., Becker, R. H. 1992, ApJS, 79, 331.
White, R. L., Kinney, A. L., Becker, R. H. 1993, ApJ, 407,
456.
Whiting, M. T., Webster, R. L., Francis, P. J. 2001, MN-
RAS, 323, 718.
Yu, Q., Tremaine, S. 2002, MNRAS, 335, 965.
APPENDIX A: OPTICAL PHOTOMETRY
Optical, V and R-band, photometry for several of the target
quasars was obtained by J. K. Davies over the course of
his 5 night run (2000 August 24-28 UT) at the Jacobus
Kapteyn Telescope in La Palma. Since these data were not
obtained for all of the target quasars, they were not used in
the analysis. However they are plotted in Figure 2, which
shows that, except for the reddest quasars with α < −1, the
optical data generally trace the UV turnover that is seen in
the low redshift mean SED (Elvis et al. 1994a).
The optical images were reduced using the ccdproc
and apphot packages in IRAF. All frames were bias sub-
tracted, and master sky and dome flats were produced for
each filter. The bias-subtracted images were first divided by
the sky or dome flats to remove small scale features, and
then, to reduce the residual gradient, they were divided by
a heavily smoothed median of the sky- or dome-flattened
science frames. Standard stars from Landolt (1992) were
observed for flux calibration. Zero points, extinction and
color coefficients were determined from fits to standard star
instrumental magnitudes measured through an aperture of
radius 20 pixels (diameter 13.′′2). As with the near-IR pho-
tometry, aperture corrections were needed. In this case, the
radius of the small aperture was 6 pixels (diameter 4′′) ex-
cept for one instance when it had to be decreased to 4 (di-
ameter 2.′′6), as noted in Table A1. The aperture corrections
were determined for each image. The telescope jumped dur-
ing some of the integrations, causing E-W elongated images
with multiple peaks. These are noted in Table A1. The dis-
tance between the peaks was comparable to or smaller than
the 20-pixel radius aperture, and the magnitudes were mea-
sured in the same way as from the round images, however the
results should be regarded with more uncertainty than the
error in table A1 indicates. Since some objects were observed
either at V or R only, two sets of transformation equations,
with and without the color terms, were fit. The appropriate
set was used to determine the magnitudes listed in Table
A1. For the two objects observed only at V or R, a synthetic
color correction was incorporated in the magnitude-to-flux
conversion, by using the zero magnitude fluxes for a ν −0.46
power-law rather than for Vega (as discussed in Section 4.1).
Effective wavelengths and zero-magnitude fluxes for Vega
and for the ν −0.46 power-law were computed for both the
Harris V and R filters used at the JKT. For V, these are
0.545µm, 36.46 × 10−10 and 36.11 × 10−10 erg s−1 cm−2A−1,
respectively. For R, these are 0.639µm, 22.44 × 10−10 and
22.11 × 10−10 erg s−1 cm−2A−1, respectively.
APPENDIX B: ESTIMATED EMISSION LINE
CONTRIBUTIONS TO THE BROAD BAND
FLUXES
For z = 3−4, the emission lines of MgII, Hβ, [OIII]4959,5007
as well as the blended FeII lines fall within the Z, J, H and
K bands. I used the FBQS composite spectrum of Brother-
ton et al. (2001) to compute, as a function of redshift, the
relative contribution of emission lines to the continuum flux
within the band. The bands were defined as discussed in sec-
tion 4.1 and listed in Table 3. Following Brotherton et al.
(2001), the continuum level was taken to be a power law fit
through the composite at 1285, 2200, 4200 and 5770A, which
had a spectral index, α = −0.46. The ratios of continuum
to total flux within a band range from ∼ 0.7 − 0.96 and are
plotted as a function of redshift in Figure B1. For z > 3, the
blended FeII and Balmer continuum falls within J and its
contribution to the broad band flux is greater than the emis-
sion line contribution at K, which explains why the spectral
indices computed from line-corrected fluxes are redder than
those computed from uncorrected fluxes (Table 4). Differ-
16 O. P. Kuhn
Table A1. Optical observations
Quasar
PKS2215+02
MG3J222537+2040
TXS2342+342
TXS2358+189
MG1J000655+1416
MG3J015105+2516
PKS0201+113
MG3J023222+2318
PKS0335-122
PKS0336-017
Va
20.58 ± 0.14b
18.78 ± 0.03
18.89 ± 0.04
21.45 ± 0.19
21.21 ± 0.16
18.49 ± 0.04
20.46 ± 0.09
19.90 ± 0.05
20.12 ± 0.07
20.15 ± 0.10
20.30 ± 0.08
...
...
Ra
20.19 ± 0.05
18.42 ± 0.02
18.52 ± 0.03
...
...
18.12 ± 0.03
20.45 ± 0.09
19.23 ± 0.03
19.81 ± 0.06
20.13 ± 0.08
18.53 ± 0.08
18.50 ± 0.07
UT date
2000 Aug 28
2000 Aug 25
2000 Aug 24
2000 Aug 28
2000 Aug 28
2000 Aug 26
2000 Aug 25
2000 Aug 27
2000 Aug 25
2000 Aug 28
2000 Aug 28
2000 Aug 28
2000 Aug 28
comments
tel jumped 1x during V image
tel jumped 1x, not color corrected
tel jumped 2x, not color corrected
tel jumped 1x during R image
not color corrected
not color corrected
a All magnitudes were measured within an aperture of radius 6 pixels (4′′diameter) except PKS2215+02 at R, and an
aperture correction was applied to determine the magnitude within the 20 pixel radius (13.′′2 diameter) aperture used for
the standard stars. Magnitudes are determined assuming an extinction and color correction except when the quasar was
observed in only one band, in which case a synthetic color correction is incorporated into the magnitude to flux conversion.
b On the R-band image, the aperture correction had to be made from an aperture of radius 4, rather than 6, pixels.
1.1
1
0.9
0.8
3
3.5
4
4.5
Figure B1. Ratio of continuum to total (continuum plus emis-
sion line) flux within the Z, J, H and K bands as a function of
redshift. The thick lines are used for the UFTI Z, J, H and K
filters and thin dotted lines for the 'tophat' approximations to
J, H, K. Inset is the mean FBQS composite of Brotherton et al.
(2001; dotted line) and continuum fit (a Fν ∼ ν−0.46 power law;
solid line) that were used in this calculation.
ences in line strengths from object to object (e.g. McIntosh
et al. 1999, Forster et al. 2001), however, are a significant
source of uncertainty. In the analysis, I have opted to use
uncorrected spectral indices since these involve one fewer
assumption and more easily compared to results from other
studies.
This paper has been typeset from a TEX/ LATEX file prepared
by the author.
|
0808.2582 | 1 | 0808 | 2008-08-19T12:24:17 | Muon Charge Ratio of Ultrahigh Energy Cosmic Rays | [
"astro-ph",
"hep-ex",
"hep-ph",
"nucl-ex"
] | The muon charge ratio of ultrahigh energy cosmic rays may provide information to detect the composition of the primary cosmic rays. We propose to extract the charge information of high energy muons in very inclined extensive air showers by analyzing their relative lateral positions in the shower transverse plane. | astro-ph | astro-ph | November 16, 2018 12:26 WSPC/INSTRUCTION FILE
mabq
Modern Physics Letters A
c(cid:13) World Scientific Publishing Company
8
0
0
2
g
u
A
9
1
]
h
p
-
o
r
t
s
a
[
1
v
2
8
5
2
.
8
0
8
0
:
v
i
X
r
a
Muon Charge Ratio of Ultrahigh Energy Cosmic Rays∗
BO-QIANG MA
School of Physics and State Key Laboratory of Nuclear Physics and Technology,
Peking University, Beijing 100871, China
[email protected]
The muon charge ratio of ultrahigh energy cosmic rays may provide information to
detect the composition of the primary cosmic rays. We propose to extract the charge
information of high energy muons in very inclined extensive air showers by analyzing
their relative lateral positions in the shower transverse plane.
The most high energy particles can be observed by human being are from cosmic
rays. The study of them belongs to frontiers of human knowledge in combination of
cosmology, astrophysics, and particle physics, and can provide better understanding
of the universe from most small to most big, i.e., connecting quarks to the cosmos.
The universe is not empty, but full of background relic particles from the big bang.
It has long been anticipated that the highest energy cosmic rays would be protons
from outside the galaxy, and there is an upper limit of the highest energy in the
observed proton spectrum, commonly referred to as the GZK cutoff 1, as the protons
traveling from intergalactic distances should experience energy losses owing to pion
productions by the photons in the cosmic background radiation. Although there
have been attentions for the cosmic ray events above the GZK cutoff, it is natural
to expect that these ultrahigh energy cosmic rays come from sources within the GZK
zone 2, i.e., not far from us in more than tens of Mpc. Recently there are also reports
on the observation of the GZK cut-off by new experiments 3. However, questions
about the composition of such ultrahigh energy cosmic ray particles, e.g., whether
they are protons, neutrons, or anti-nucleons 4, are still open to investigations.
Muons in the air showers are mainly from decays of pions and kaons produced
in the interactions of the primary cosmic rays with the atmosphere. The very high
energy secondary pion and kaon cosmic rays can be considered as from the current
fragmentation of partons in deep inelastic scattering of the primary cosmic rays with
the nucleon targets of the atmosphere in a first approximation 5. We also consider
only the favored fragmentation processes, i.e., the π+, which is composed of valence
u and ¯d quarks, is from the fragmentation of u and ¯d quarks in the nucleon beam,
∗Invited talk at 2007 International Symposium on Cosmology and Particle Astrophysics
(CosPA2007), November 13-15, 2007, Taipei.
1
November 16, 2018 12:26 WSPC/INSTRUCTION FILE
mabq
2 BO-QIANG MA
and the π−, which is composed of valence ¯u and d quarks, is from the fragmentation
of ¯u and d quarks 6. Similarly, the K +, which is composed of valence u and ¯s, is from
the fragmentation of u and ¯s quarks, and the K −, which is composed of valence
¯u and s, is from the fragmentation of ¯u and s quarks. The µ+ is from the decay
of a π+ or a K + and the µ− is from the decay of a π− or a K −. We can roughly
estimate the muon charge ratio by
µ+
µ− =
R 1
0 dx (cid:8)(cid:2)u(x) + ¯d(x)(cid:3) + κ [u(x) + ¯s(x)](cid:9)
R 1
0 dx {[d(x) + ¯u(x)] + κ [¯u(x) + s(x)]}
,
(1)
where q(x) is the quark distribution with flavor q for the incident hadron beam
and κ ∼ 0.1 → 0.3 is a factor reflecting the relative muon flux and fragmentation
behavior of K/π. Secondary collisions do not influence the above estimation, since
the current parton beams still keep their flavor content and act as the current
partons after the strong interactions with the partons in the atmosphere targets.
Adopting a simple model estimation of the parton flavor content in the nucleon
without any parameter 7, we find that µ+/µ− ∼ 1.7 for proton and µ+/µ− ∼ 0.7
for neutron. This simple evaluation is in agreement with the empirical expectation
of µ+/µ− ≈ 1.66 for proton and µ+/µ− ≈ 0.695 for neutron 8 as well as that in an
extensive Monte Carlo calculation 9, thus it provides a clear picture to understand
the dominant features for the muon charge ratio by the primary hadronic cosmic
rays. For the µ+/µ− ratio for antiproton, it is equivalent to the µ−/µ+ ratio for
proton by using Eq. (1), thus we find µ+/µ− ∼ 0.6 for antiproton, which is close to
that for neutron. The µ+/µ− ratio for antineutron is also equivalent to the µ−/µ+
ratio for neutron, and it is µ+/µ− ∼ 1.4, which is close to that for proton. It
is hard to distinguish between the primary neutrons and antiprotons (or protons
and antineutrons) by the µ+/µ− ratio of the air shower, unless very high precision
measurement is performed and also our knowledge of the muon charge ratio for each
nucleon species is well established.
The study of cosmic rays with primary energies above 105 GeV are typically
based on the measurements of extensive air showers (EAS) that they initiate in the
atmosphere. The ground detector array records the secondary particles produced in
shower cascades, including photons, electrons (positrons), muons, and some hadrons.
Then their arrival times and density profiles are used to infer the primary energy
and composition of the incident cosmic ray particle, usually through comparison
with simulated results. Photons, electrons and positrons are the most numerous
secondary particles in an EAS event. However, for very inclined showers, these
electromagnetic components would travel a long slant distance and are almost com-
pletely absorbed before they reach the ground. On the other hand, muons are decay
products of charged mesons in shower hadronic cascades. Most high energy muons
survive their propagation through the slant atmospheric depth, during which they
lose typically a few tens of GeV's energy. These high energy muons carry impor-
tant information about the nature of the primary cosmic ray hadron, which will be
extracted from their energy spectrum and lateral distribution.
November 16, 2018 12:26 WSPC/INSTRUCTION FILE
mabq
Muon Charge Ratio of Ultrahigh Energy Cosmic Rays
3
As discussed by Hwang and I 4, the ratio of positive versus negative muons
µ+/µ− is a significant quantity which can help to discern the primary composition,
and at high energies this charge ratio also reflects important features of hadronic
meson production in cosmic ray collisions. In order to obtain such muon charge
information, we would need a way to distinguish between positive and negative high
energy muons. Unfortunately, existing muon detectors available at shower arrays,
usually scintillators and water Cerenkov detectors, are not commonly equipped with
magnetized steel to differentiate the muon charges. Even if they were, the limited
region of the magnetic field prevents definite determination of high energy muons'
track curvature.
This invites us to think of the geomagnetic field as a huge natural detector for
muon charge information. Apparently, after being produced high in the atmosphere,
a positively charged muon would bend east on its way down while a negatively
charged muon would bend west, introducing an asymmetry into the density profile
of the shower front. If their separation is large enough as compared with other
circularly symmetric "background" deviations, it will be possible to distinguish the
positive muons from the negative ones.
To see such an effect, Xue and I 10 analyzed the possibility of obtaining the
charge information of high energy muons in very inclined extensive air showers.
We have demonstrated that positive and negative high energy muons in sufficiently
inclined air showers can be distinguished from each other through their opposite ge-
omagnetic deviations in the transverse plane. We developed a revised Heitler model
to calculate this distinct double-lobed distribution, and studied the condition for
the two lobes of either positive or negative muons to be separable with confidence.
From our criterion of resolvability, we concluded that a zenith angle 75◦ ≤ θ ≤ 85◦
will be most suitable for our approach.
There are already some results from full air shower simulations that take into
account the geomagnetic effect on muon propagation 11,12,13,14,15,16,17,18. They
illustrated remarkable double-lobed muon lateral density profile in very inclined
air showers, which is in agreement with our expectation qualitatively. However, no
present study has fully considered the high energy part of muon content, which
can be used to compare with our results. Thus we would like to propose future
simulations of very inclined extensive air showers that focus on the behavior of high
energy muons. They also have to keep track of the muon charges and the relation
to their lateral positions. For more detailed analysis and discussion, please refer to
Ref.10.
In summary, we propose to extract the charge information of high energy muons
in very inclined extensive air showers by analyzing their relative lateral positions in
the shower transverse plane. This muon charge information is helpful to detect the
composition of cosmic rays, e.g., the neutron or antiproton content of the ultrahigh
energy cosmic rays.
November 16, 2018 12:26 WSPC/INSTRUCTION FILE
mabq
4 BO-QIANG MA
Acknowledgments
I am very grateful to Pauchy Hwang and the organizers for their invitation and
warm hospitality. I also thank Pauchy Hwang and BingKan Xue for the collab-
orated results in this talk. This work is partially supported by National Natural
Science Foundation of China (Nos. 10721063, 10575003, 10528510), by the Key
Grant Project of Chinese Ministry of Education (No. 305001), and by the Research
Fund for the Doctoral Program of Higher Education (China).
References
1. K. Greisen, Phys. Rev. Lett. 16, 748 (1966); G.T. Zatsepin and V.A. Kuzmin, Pis'ma
Zh. Eksp. Tero. Fiz. 4, 114 (1966) [JETP Lett. 4, 78 (1966)].
2. F.W. Stecker, Phys. Rev. Lett. 21, 1016 (1968). For an extensive review, see,
F.W. Stecker, astro-ph/0101072.
3. R. Abbasi et al., HiRes Collaboration, Phys. Rev. Lett. 100, 101101 (2008). Also Pierre
Auger Observatory reported the observation of the GZK cut-off.
4. W-Y.P. Hwang and B.-Q. Ma, Eur. Phys. J. A 25, 467 (2005).
5. See, e.g., B.-Q. Ma, I. Schmidt, J. Soffer, and J.-J. Yang, Nucl. Phys. A 703, 346 (2002).
6. See, e.g., B.-Q. Ma, I. Schmidt, and J.-J. Yang, Phys. Rev. D 65, 034010 (2002).
7. Y.-J. Zhang, B. Zhang, and B.-Q. Ma, Phys. Lett. B 523, 260 (2001).
8. R.K. Adair, Phys. Rev. Lett. 33, 115 (1974); R.K. Adair et al., Phys. Rev. Lett. 39,
112 (1977); O.C. Allkofer et al., Phys. Rev. Lett. 41, 832 (1978).
9. J.N. Capdevielle and Y. Muraki, Astropart. Phys. 11, 335 (1999). See, e.g., Fig. 7.
10. B. Xue and B.-Q. Ma, Astropart. Phys. 27, 286 (2007).
11. M. Ave, R. A. Vazquez and E. Zas, Astropart. Phys. 14, 91 (2000).
12. H. H. Aly, M. F. Kaplon, and M. L. Shen, Nuovo. Cim. 31, 905 (1964).
13. J. R. Horandel, J. Phys. G 29, 2439 (2003).
14. C. A. Ayre et al., J. Phys. A 5, L102 (1972).
15. T. Hebbeker and C. Timmermans, Astropart. Phys. 18, 107 (2002).
16. S. Ostapchenko, Nucl. Phys. Proc. Suppl. 151, 147 (2006).
17. See, e.g., W. D. Apel et al. [KASCADE Collaboration], arXiv:astro-ph/0510810.
18. P. Hansen, T. K. Gaisser, T. Stanev and S. J. Sciutto, Phys. Rev. D 71, 083012 (2005);
A. Cillis and S. J. Sciutto, arXiv:astro-ph/9908002.
|
astro-ph/0504322 | 1 | 0504 | 2005-04-14T11:58:51 | Optimal column density measurements from multiband near-infrared observations | [
"astro-ph"
] | We consider from a general point of view the problem of determining the extinction in dense molecular clouds. We use a rigorous statistical approach to characterize the properties of the most widely used optical and infrared techniques, namely the star count and the color excess methods. We propose a new maximum-likelihood method that takes advantage of both star counts and star colors to provide an optimal estimate of the extinction. Detailed numerical simulations show that our method performs optimally under a wide range of conditions and, in particular, is significantly superior to the standard techniques for clouds with high column-densities and affected by contamination by foreground stars. | astro-ph | astro-ph | Astronomy & Astrophysics manuscript no. 2267ms
(DOI: will be inserted by hand later)
July 2, 2018
5
0
0
2
r
p
A
4
1
1
v
2
2
3
4
0
5
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
Optimal column density measurements from multiband
near-infrared observations
M. Lombardi1,2
1 European Southern Observatory, Karl-Schwarzschild-Strasse 2, D-85748 Garching bei Munchen, Germany
2 University of Milan, Department of Physics, via Celoria 16, I-20133 Milan, Italy
Received ***date***; Accepted ***date***
Abstract. We consider from a general point of view the problem of determining the extinction in dense molecular
clouds. We use a rigorous statistical approach to characterize the properties of the most widely used optical and
infrared techniques, namely the star count and the color excess methods. We propose a new maximum-likelihood
method that takes advantage of both star counts and star colors to provide an optimal estimate of the extinction.
Detailed numerical simulations show that our method performs optimally under a wide range of conditions and, in
particular, is significantly superior to the standard techniques for clouds with high column-densities and affected
by contamination by foreground stars.
Key words. dust, extinction -- methods: statistical -- ISM: clouds -- infrared: ISM -- stars: formation
1. Introduction
A detailed comprehension of the structure and physical
properties of dark molecular clouds is a critical step to
understand fundamental processes such as star and planet
formation. Still, after several decades of investigations,
very little is known about dark clouds, about their re-
lationship to the diffuse interstellar medium, and about
the composition and evolution of the dust grains present
in the clouds.
Molecular clouds are composed of approximately 99%
molecular hydrogen and helium, but due to the absence
of a dipole moment these molecules are virtually unde-
tectable at the low temperature (∼ 10 K) that charac-
terize these objects. As a result, astronomers have been
using different tracers to study the density distribution
of molecular clouds. Historically, the first technique was
based on a statistical analysis of the angular density of
stars observed through a cloud (Wolf 1923; Bok 1937).
This method, known as star counts, uses the dust (which
is responsible for the extinction of light) as a tracer of H2
and He.
More recently, radio observations of carbon monox-
ide (CO) and its isotopes have been used to study the
spatial distribution and the physical conditions of molec-
ular clouds, under the assumption that CO is a reli-
able tracer of the matter in the cloud, i.e. that the ra-
tio N (CO)/(cid:0)N (H2) + N (He)(cid:1) is approximately constant.
Send offprint requests to: M. Lombardi
Correspondence to: [email protected]
Radio spectroscopy techniques have provided extremely
effective in studying the most dense regions (above
1022protons cm−2) and, more importantly, give dynamical
information on the cloud structure. However, with the ad-
vent of various dust detectors in the near-infrared (NIR),
far infrared, millimeter, and sub-millimeter bands, it has
become clear that several poorly constrained processes
(e.g., deviations from thermodynamic equilibrium, opac-
ity variations, chemical evolution, and more importantly
depletion of molecules) can significantly affect the results
of analyses based on radio observations (e.g. Lada et al.
1994; Alves et al. 1999; Harjunpaa et al. 2004).
The reddening of background stars offers a natural
method to study the distribution of dust in molecu-
lar clouds, and thus the hydrogen column density. This
technique can be better applied to the infrared bands,
which compared to optical bands are less affected by
extinction and are less sensitive to the physical prop-
erties of the dust grains (Mathis 1990). Before the ad-
vent of large format array cameras, the lack of instru-
mental sensitivity clogged infrared observations to small,
dense clouds (e.g. Jones et al. 1980; Frerking et al. 1982;
Jones et al. 1984; Casali 1986). More recently, infrared ar-
rays have made it possible to measure thousands of stars
and to extend the original technique to entire molecular
cloud complexes (Lada et al. 1994, 1999; Alves et al. 2001;
Lombardi & Alves 2001). Such measurements are free of
the complications that plague molecular-line data and per-
mit a detailed analysis of the cloud density distribution.
2
M. Lombardi: Optimal column density measurements from multiband near-infrared observations
Although the success of the NIR color excess method
is evident, there is still a need for a deeper understanding
of its limitations, statistical biases, and uncertainties. The
aim of this paper is twofold: (i) to study in detail the sta-
tistical properties of the star count and color excess meth-
ods and (ii) to describe a new, optimal method based on
a maximum-likelihood analysis. The method is described
here for NIR observations, but could equally well be ap-
plied to other infrared bands for which the extinction law
and the intrinsic color distribution are known with good
accuracy. The paper is organized as follows. In Sect. 2
we introduce our formalism and consider from a statisti-
cal point of view the extinction of stars by a foreground
cloud. The two main techniques used to obtain extinction
measurements, namely the star count and the NIR color
excess methods, are discussed in Sect. 3. In Sect. 4 we
describe the new maximum-likelihood method and show
by means of numerical simulations that it performs excel-
lently even in the presence of a significant contamination
by foreground stars. Our conclusions are briefly reported
in Sect. 5. Finally, in Appendix A we consider the sta-
tistical properties of the median and of related estimators
used often in infrared studies of dark clouds to remove the
effects of foreground stars.
2. Basic relations
Suppose that N stars with magnitudes (on a given band)
{ mn} are observed in a given region of the sky (here-
after we will use the hat to denote measured quanti-
ties). These observed magnitudes depend on the original,
unreddened, absolute star magnitudes {Mn}, on the star
distances {Dn}, on the individual extinctions {An}, and
on the photometric errors {ǫn}. In particular, we have
mn = Mn + 5 log10 Dn − 5 + An + ǫn = mn + ǫn , (1)
where the distances {Dn} are taken to be expressed in
parsecs. The quantity mn that appears in the r.h.s. of this
equation is the reddened magnitude of the n-th star free
from measurement errors (i.e., the magnitude that would
be observed in the limit of an extremely long exposure
time).
2.1. Single-band probability distributions
In order to consider the most general situation in a proper
statistical way, we introduce several probability distribu-
tions:
ρ(M, D): the probability distribution (density) of stars
with absolute magnitude M at distance D;
pA(AD): the probability distribution of extinction for ob-
jects located at distance D, i.e. the probability of an
extinction of A magnitudes on a star given that its
distance is D;
pǫ( mm): the probability distribution of photometric
measurement errors, i.e. the probability of measuring
a magnitude m for a star given that its true reddened
magnitude is m.
These probability distributions are now defined in the
most general way. Later, however, we will consider spe-
cial forms of them that allow us to simplify the relevant
equations. Note also that we take the extinction A at dis-
tance D as a random variable. Hence, we have introduced
above the probability distribution pA(AD) rather than
a (deterministic) function A(D). This general approach
can be used to describe molecular clouds with "patchy"
column densities, which seems to be quite common (see,
e.g., Lada et al. 1999; Cambr´esy 1999). Finally, we use a
general form for the photometric error that includes the
common case where the error depends on the magnitude
of the star. Moreover, we also consider through pǫ( mm)
the case of undetected objects, i.e. the case where a star
of true magnitude m close to the detection limit produces
no detectable flux; in this case we write m = null. Hence,
the quantity
c(m) = 1 − pǫ(nullm)
(2)
represents the completeness of our observations for stars
with true magnitude m.
We assume that both pA(AD) and pǫ( mm) are nor-
malized to unity, i.e.
Z ∞
pǫ(nullm) +Z ∞
−∞
0
pA(AD) dA = 1
pǫ( mm) d m = 1
∀D ,
∀m .
(3)
(4)
Moreover, we take ρ(M, D) to be normalized so that the
quantity
dN = D2ρ(M, D) dD dΩ dM
(5)
represents the expected number of stars located at a dis-
tance in the range [D, D + dD] in a patch of the sky of
solid angle dΩ, and with absolute magnitudes between M
and M + dM .
Equation (1) can be used to obtain the probability dis-
tribution of reddened magnitudes pm(m). Using the pre-
vious definitions we find
pm(m) =Z ∞
0
dD D2Z ∞
× ρ(m − 5 log10 D + 5 − A, D) .
dA pA(AD)
0
(6)
Finally, pm(m) can be converted into the distribution of
observed magnitudes p m( m) by convolving it with the
measurement error probability distribution pǫ( mm):
p m( m) =Z ∞
Because of the way pǫ( mm) is defined, Eq. (7) includes
also the case m = null. Note also that p m( m) is normalized
so that p m( m) d m represents the expected angular density
of stars with observed magnitudes on the range [ m, m +
d m].
pǫ( mm)pm(m) dm .
(7)
∞
Equations (6) and (7) are basic equations of this paper.
However, in the form they are written, they are by far too
general to be useful in practical cases.
M. Lombardi: Optimal column density measurements from multiband near-infrared observations
3
2.2. Simplifications
In order to make use of the general relations written above
we introduce a number of useful simplifications.
First, we note that so far we have considered obser-
vations carried out on a single band. In order to threat
multi-band data, we introduce a reddening law as follows:
for any band λ, the extinction Aλ is given by
Aλ = kλAV ,
(8)
where kλ is a constant and AV is the extinction in the V
band. In other words, Eq. (8) states that the ratio Aλ/Aλ′
between the extinction in two bands λ and λ′ is a constant
for a given cloud. In reality, this ratio is known to vary
with the environment for short wavelengths, but is almost
universal in the NIR (and, according to a recent work of
Indebetouw et al. 2004, for wavelengths in the range 3 --
10 µm as well). In the rest of this paper we will follow the
usual notation and express all extinctions in terms of AV .
A second important assumption that we will use re-
gards the functional form of ρ(M, D). First, we will as-
sume that this function can be separated in M and D,
i.e. can be written as the product of two functions, one
involving M only, and one involving D only: ρ(M, D) =
pM (M )ρ(D). In practice, this assumption is equivalent to
saying that we observe a single "population" of stars at
all distances. Since observations show that star number
counts are well approximated by an exponential law of
the form pm(m) ∝ 10αm (where the exponent α ≃ 0.34 is
approximately independent of the band considered in the
NIR), we also assume that pM (M ) follows a similar law.
We then write
ρ(M, D) = νλ10αM ρ(D) ,
(9)
where νλ is a normalization factor. In the following we will
assume that ρ(D), similarly to α, is independent of the
band considered, while νλ is not. Note also that, because
of the exponential form of this equation in M , a change
of the normalization factor νλ is in practice equivalent to
a change of the zero-point used for the magnitude M .
Using Eqs. (9) and (8) we can slightly simplify Eq. (6)
by performing the integration over D:
pm(m) = Z ∞
0
dD D2−5αρ(D)Z ∞
× νλ105α10αm10−αkλAV
0
dAV pAV (AV D)
≡ aνλ10αmZ ∞
0
dAV pAV (AV )10−αkλAV , (10)
where we have defined a distance-weighted probability dis-
tribution for AV as
105α
dD D2−5αρ(D)pAV (AV D) , (11)
and where a is a numerical factor introduced in Eq. (10)
to ensure that pAV (AV ) is normalized to unity:
a Z ∞
pAV (AV ) =
0
a = 105αZ ∞
0
dD D2−5αρ(D) .
(12)
V ) = δ(A′
Note that pAV (AV ) has a simple interpretation, being
the probability distribution for stars to have undergone
a given extinction regardless of their distance. For exam-
ple, if all stars were subject to the same extinction AV , we
would have pAV (A′
V − AV ). Note also that, since
α ≃ 0.34 can be taken to be approximately independent of
the band considered, pAV (AV ) is also band-independent.
Equation (10) shows an interesting property of extinc-
tion studies: the probability distribution pm(m) depends
on pAV (AV D) only through pAV (AV ). This point is par-
ticularly important since it shows an intrinsic limitation
of extinction measurements. Indeed, all observables are
derived from p m( m), the distribution of observed magni-
tudes, and this function depends only on pm(m) (other
than observational "limiting factors" such as pǫ( mm) or
c( m)). Since, as noted above, pm(m) does not depend di-
rectly on pAV (AV D), any estimator based on observed
magnitudes only1 will not provide any information on
pAV (AV D) directly but (in the best case scenario) only
on pAV (AV ). Since pAV (AV ) is a sort of convolution of
pAV (AV D), in general it is not possible to have a complete
knowledge on pAV (AV D), not even if we know the star
density distribution ρ(D); this limitation, among other
things, prevents us from gathering information on the
three-dimensional structure of a molecular cloud.
A case where, instead, the function pAV (AV D) can be
recovered is a deterministic model for the extinction, i.e.
a cloud complex whose extinction AV depends directly on
the distance D (and, thus, is not a random variable). In
this case we have
pAV (AV D) = δ(cid:0)AV − AV (D)(cid:1) .
The extinction AV (D) has a simple relationship with the
integrated gas column density ρgas:
(13)
AV (D) =
1
β Z D
0
dD′ ρgas(D′) ,
(14)
where β ≃ 2 × 1021 cm−2 mag−1 is the ratio between gas
density and V -band extinction (Lilley 1955; Bohlin et al.
1978). Suppose now that we carry out observations on the
cloud and measure pAV (AV ). In order to show that we can
recover the structure of the cloud, let us call D(AV ) the
inverse of the function AV (D) defined in Eq. (14), i.e. the
distance at which the integrated extinction is AV . Then
we have
105α
pAV (AV ) =
=
105α
a (cid:2)D(AV )(cid:3)2−5α
ρ(cid:0)D(AV )(cid:1)
3 − 5α
a
d
ρ(cid:0)D(AV )(cid:1)D′(AV )
dAV h(cid:2)D(AV )(cid:3)3−5αi .
(15)
1 Through this paper we assume that the distance of individ-
ual stars is not an observable; if, instead, the distance of each
star can be estimated, then in principle one can also directly
measure pAV (AV D).
4
M. Lombardi: Optimal column density measurements from multiband near-infrared observations
(cid:2)D(AV )(cid:3)3−5α
In other words, we have obtained a differential equation in
. If ρ(D) is known, using the boundary con-
dition D(0) = 0 we can in principle solve this differential
equation and obtain D(AV ) which, in turn, can be in-
verted into AV (D) and ρgas(D). It is superfluous to stress
that this process in reality is far from being trivial.
Finally, we consider a very simple deterministic model
for pAλ(AλD) that describes a thin cloud at a distance
D0 with uniform column density AV :
pAV (A′
V D) =(δ(A′
δ(A′
V )
if D < D0 ,
V − AV ) otherwise .
(16)
The related distribution pAV (AV ) takes then the simple
form
pAV (A′
V ) = f δ(A′
V ) + (1 − f )δ(A′
V − AV ) ,
(17)
where f is a real number in the range [0, 1] given by
f =(cid:20)Z D0
0
dD D2−5αρ(D)(cid:21)(cid:30)(cid:20)Z ∞
0
dD D2−5αρ(D)(cid:21) .
(18)
Hence, f represents the fraction of foreground stars (i.e.
stars at distance D < D0) present in any apparent mag-
nitude bin in regions with negligible extinction. In the
following, we will refer to the simple case described in
Eqs. (16) and (17) as "thin cloud approximation" (the
term is borrowed from gravitational lensing theory).
2.3. The completeness function
Above, in Eq. (2), we introduced the completeness func-
tion c(m), which gives the probability to detect a star of
magnitude m. Typically, c(m) is close to unity for bright
stars, and vanishes for very faint stars; note however that
c(m) might be smaller than unity at low m in crowded
fields or close to bright objects. The completeness func-
tion c(m) enters naturally in the definition of the error
probability distribution pǫ( mm), which can be written as
pǫ( mm) =(c(m)pǫ( mm)
if m 6= null ,
if m = null .
(19)
1 − c(m)
This representation reflects the measurement process: for
any star, there is first a random process that "decides"
whether the object is detected or not (with probability
fixed by c(m)); then, if the star is detected, there is a
second random process that generates the observed mag-
nitude m according to pǫ( mm). Note also that this latter
probability distribution is normalized [cf. Eq. (4)].
In order to carry out some simplifications (see below
Sect. 4.2), we also consider a different representation of
the completeness function in which the order of the two
random processes described above is swapped: this leads
to a completeness function c( m) defined in terms of the
observed magnitude. In other words, we first generate for
every star the "observed" magnitude m according to a
probability distribution pǫ( mm), and then decide whether
the star is really observed using a second random process
controlled by the completeness function c( m). This im-
plies, among other things, a modification of Eq. (7), that
now becomes
−∞
pǫ( mm)pm(m) dm .
p m( m) = c( m)Z ∞
This equation is valid for detected stars, i.e. if m 6= null.
Since c( m) represents the probability that a star with mea-
sured magnitude m be detected, we evaluate the proba-
bility that the star is not detected as
(20)
p m(null) =Z ∞
−∞
d m(cid:2)1− c( m)(cid:3)Z ∞
−∞
dm pǫ( mm)pm(m) .
(21)
An important point to observe here is the fact that we have
modified Eq. (7) into Eqs. (20) and (21) by defining two
functions, the new error distribution pǫ( mm) and the new
completeness function c( m), which replace, respectively,
pǫ( mm) and c(m). Indeed, both these couples of functions
are intimately related, and this was implicitly taken into
account above by using a single distribution pǫ( mm) to
describe both the photometric errors and the completeness
of the observations. Interestingly, it is possible to find a
relationship between the couple pǫ( mm), c( m) and the
couple pǫ( mm), c(m) by requiring that, for any reddened
magnitude probability distribution pm(m), the observed
magnitude probability distribution p m( m) evaluated using
Eq. (7), or using Eqs. (20) and (21), agrees. We find then
c(m)pǫ( mm) = c( m)pǫ( mm) ,
1 − c(m) =Z ∞
−∞
pǫ( mm)(cid:2)1 − c( m)(cid:3) d m .
if we integrate
Since pǫ( mm) is normalized to unity,
Eq. (22) over m and substitute the result into Eq. (23),
we find
Z ∞
−∞
pǫ( mm) d m = 1 ,
(24)
i.e. pǫ is also normalized to unity. In summary we have:
(22)
(23)
(25)
(26)
−∞
c(m) =Z ∞
pǫ( mm) =
R ∞
−∞ c( m)pǫ( mm) d m
c( m)pǫ( mm) d m ,
c( m)pǫ( mm)
.
Since Eq. (25) involves a convolution, it is in general not
possible to invert it and express c( m) in terms of c(m).
Note that Eq. (26) is essentially Bayes' theorem applied
to pǫ( mm).
Although Eqs. (25) and (26) show that the description
of the completeness c(m) in terms of the true reddened
magnitude m is more general than the description c( m)
in terms of the observed magnitude m, we argue that the
latter is more practical to use:
M. Lombardi: Optimal column density measurements from multiband near-infrared observations
5
-- Operationally, the completeness function is usually
evaluated from the observed magnitudes by compar-
ing the expected number of stars in a given magnitude
bin with the number of stars detected in the same bin.
-- Often, it can be sensible to use some a posteriori cuts
in the star catalog. For example, if we know that ob-
servations of faint stars are particularly unreliable, we
can just discard all objects with observed magnitude
m larger than a given threshold mlim. This cut could
be easily described in terms of c( m):
c( m) = H(mlim − m) =(1 if m ≤ mlim ,
0 otherwise .
(27)
-- The use of c( m) allows us to evaluate analytically
p m( m) for some special simple probability distribu-
tions. This point is particularly valuable for the practi-
cal applications that we will describe below in Sect. 4.
2.4. Multi-band probability distributions
So far we have focused our analysis on single-band mea-
surements. In order to extend the discussion to observa-
tions carried out in different bands, we need to generalize
the relevant probability distributions.
We denote M = {M1, M2, . . . , MΛ} the magnitudes
of a star in Λ different bands; in general, we use bold
symbols such as k to indicate quantities that have different
values in the various bands. We will threat these as vector
quantities; in this way, e.g., the generalization of Eq. (1)
can written as
mn = M n + 5 log10 Dn− 5 + kAV + ǫn = mn + ǫn . (28)
Since we are now working with observations in differ-
ent bands, we need also to generalize ρ(M, D). We define
thus ρ(M , D), the probability to have a star with absolute
magnitudes M at distance D from us. Using this distri-
bution we can rewrite Eq. (6) as
pm(m) =Z ∞
0
dD D2Z ∞
× ρ(m − 5 log10 D + 5 − kAV , D) .
dAV pAV (AV D)
0
(29)
The generalization of Eq. (7) is also simple. Since
photometric measurements in different bands are inde-
pendent, we have to perform Λ different convolutions
with the measurement error probability distributions
(cid:8)pǫλ( mλmλ)(cid:9) corresponding to the various bands:
p m( m) =Z dΛm pm(m)
pǫλ( mλmλ) .
Λ
Yλ=1
(30)
Note that we consider a star detected if it is detected in
at least one band.
The integral of p m( m) over all admissible values of m
gives the expected local density of stars σ:
Since σ gives the normalization of p m( m), the conditional
probability that a star with magnitudes m be observed
given the fact that the star is detected is p m( m)/σ.
When using the alternate completeness functions
(cid:8)cλ( mλ)(cid:9) expressed in terms of the observed magnitudes
{ mλ}, Eq. (30) can be rewritten as
Yλ=1hδ( mλ − m′
p m( m) = Z dΛm′
λ)cλ( m′
λ)
Λ
λ)(cid:1)i
+ δ( mλ − null)(cid:0)1 − cλ( m′
Λ
(32)
pǫλ( m′
λmλ) .
×Z dΛm pm(m)
Yλ=1
The combination of delta distributions inside the brackets
in this equation ensures that m′
λ = mλ if mλ 6= null, and
b)(cid:1) (the probability of not
detecting the star) if mλ = null; the last integration is the
usual convolution with the measurement errors. Similarly,
Eq. (31) can also be written as
that we integrate over(cid:0)1− cλ( m′
σ = Z dΛm(cid:20)1 −
Λ
Yλ=1(cid:16)1 − cλ( mλ)(cid:17)(cid:21)
Λ
×Z dΛm pm(m)
pǫλ( mλmλ) .
Yλ=1
(33)
2.5. Further simplifications
As for the single-band case, we consider a number of re-
alistic and useful simplifications that will allow us to take
advantage of the formalism introduced above in practical
cases.
First, we still take ρ(M , D) to be separable into
ρ(M , D) = pM (M )ρ(D). Furthermore, we adopt for
pM (M ) a simple functional form, which is sufficiently ac-
curate for our purposes. Since we know that pm(m) ∝
10αm with α approximately independent of the band, and
since stars appear to have a limited scatter in their NIR
colors, we write
pM (M ) = ν10αM1
× exp(cid:20)−
(Ma − M1 − χa)C−1
ab (Mb − M1 − χb)
2
(cid:21) ,
(34)
where we have used Einstein's convection on repeated in-
dexes. In other words, we suppose that M1 is exponen-
tially distributed, and that star colors are distributed as
Gaussian random variables with averages hMa−M1i = χa;
C represents the covariance of these colors. This distribu-
tion can be rewritten in a more convenient way as (see
also below Sect. 4.2)
σ =Z(R∪{null})Λ\{null}
p m( m) dΛm .
(31)
pM (M ) = exp(cid:20)−
M TP M + 2QTM + R
2
(cid:21) .
(35)
6
M. Lombardi: Optimal column density measurements from multiband near-infrared observations
The form (34) for pM (M ) is particularly convenient for
several reasons. First, we can greatly simplify Eq. (29) and
obtain a result similar to Eq. (10):
pm(m) = aZ ∞
0
dAV pAV (AV )pM (m − kAV ) ,
(36)
where pAV (AV ) and a are still given by Eqs. (11) and
(12). Again, we observe that the fact that pm(m) de-
pends only on the distance-weighted pAV (AV ), implies
that pAV (AV D) is not an observable. Moreover, the forms
(34) and (35) are unmodified by a reddening M 7→ m =
M + 5 log10 D − 5 + kAV and by a convolution with
Gaussian photometric errors (see below Sect. 4.2).
3. Extinction measurements
We will describe in this section the star counts and color
excess methods in order to better describe the advantages
and limitations of them.
3.1. Star counts
In the 18th century the English astronomer William
Herschel noted that some regions presented few stars and,
following Newton's idea of a perfectly transparent space,
interpreted these "holes in the sky" as a real lack of stars.
This misconception survived the discovery of individual
dark clouds (Barnard 1919) and had serious consequences
on Shapley's calibration of the distance scale for Cepheids.
The "discovery" of dust in our Galaxy took place only in
1930 when Trumpler showed its importance in dimming
the light coming from distant open clusters. Finally, in
recent decades dust has no longer been seen only as an-
noying "fog" obscuring the light of background sources,
and has been shown to have a tremendous impact on the
evolution of galaxies and on the formation of stars and
stellar systems (see Li & Greenberg 2003 for a detailed
historical review).
It has long been recognized (Wolf 1923) that measure-
ments of the local density of stars in different regions of
the sky can be used to map the extinction. The original
technique consisted in comparing the number of stars in
magnitude bins in regions subject to extinction with the
number of stars in regions where the extinction is (sup-
posedly) negligible. This technique was then improved by
Bok (1956) which suggested to use counts up to a limiting
magnitude to reduce the error. In the past, the star count
technique was mainly applied to optical data (typically vi-
sually inspected Smith plates; e.g. Dickman 1978; Mattila
1986; Andreazza & Vilas-Boas 1996). In the last decade,
however, near-infrared digital data have been available,
and the star count technique has been finally applied
to NIR observations (e.g. Cambresy et al. 1997). In this
respect, a key role has been played by large NIR sur-
veys such as the Two Micron All Sky Survey (2MASS;
Kleinmann et al. 1994) and the Deep NIR southern Sky
Survey (DENIS; Epchtein et al. 1997).
λ
J
H
K
L
M
N
kλ
kλ
(Rieke & Lebofsky 1985)
(van de Hulst 1946)
0.282
0.175
0.112
0.058
0.023
0.052
0.246
0.155
0.089
0.045
0.033
0.013
Table 1. The extinction law in different infrared bands
(taken from Rieke & Lebofsky 1985 and van de Hulst
1946).
The star count method is easily described using our
notation. We first note that, as for multi-band probability
distributions, the integral of p m( m) over m gives the local
density of stars σ [cf. Eq. (31)]:
σ ≡ hσi = Z p m( m) d m
= Z d m c( m)Z dm pm(m)pǫ( mm) .
Inserting here Eq. (10), we immediately obtain
σ = σ(0)Z dAV pAV (AV )10−αkλAV ,
(37)
(38)
where σ(0) is the average density expected in absence of
extinction:
σ(0) = aνλZ d m c( m)Z dm 10αmpǫ( mm) .
(39)
Equation (38) shows that the ratio σ/σ(0) between the
average densities expected in presence and in absence of
extinction is simply related to pAV (AV ).
Clearly, a single measurement of σ/σ(0) cannot be used
to derive the distribution pAV (AV ). However, in the sim-
plest case where all stars are background to a thin cloud,
so that pAV (A′
V ), we find the classical re-
lation
σ
σ(0) = 10−αkλAV ,
V ) = δ(AV − A′
(40)
which can be immediately inverted to obtain AV (Bok
1956). Operationally, both densities σ and σ(0) are mea-
sured by dividing the number of stars observed against the
cloud and in a control field by the angular size of the re-
gions considered. Hence, the measurement errors on these
quantities are due to the randomness on the local number
of stars. If we ignore the correlation on the star positions,
and thus assume that these are a homogeneous Poisson
process (see, e.g., Cressie 1993), then the local number of
stars follows a simple Poisson distribution.
Although the estimator (40) can be applied to thin
clouds only, in principle more general situations can also
be handled by taking advantage of the particular form of
Eq. (38). We first observe that the integral appearing in
the r.h.s. of this equation is reminiscent of a Laplace's
M. Lombardi: Optimal column density measurements from multiband near-infrared observations
7
transform. Given a real function f (x) defined for non-
negative values x, its Laplace's transform is defined as
L[f ](k) =Z ∞
0
f (x) exp(−kx) .
(41)
An important property of Laplace's transform is that it
can be inverted: in other words, it is possible to obtain
f (x) provided one knows L[f ](k) for any non-negative
k. Using the notation just introduced, we can rewrite
Eq. (38) as
σ = σ(0)L[pAV ](αkλ ln 10) .
(42)
Hence, if we were able to measure the ratio σ/σ(0) for all
positive values of the product αkλ, we could in principle
derive pAV . As indicated above, the constant α ≃ 0.34 is
approximately independent of the band considered; how-
ever, kλ strongly depends on the band used to carry out
the observations (see Table 1). Hence, the product αkλ
that appears in Eq. (42) can be varied among a limited
set of values. In conclusion, although complete knowledge
of pAV is clearly impossible, we can still use multiband ob-
servations to investigate pAV in more complex situations
than the one considered in Eq. (40).
Let us consider, for example, the case of the thin cloud
approximation described in Eq. (17). Since, in general,
the fraction f of foreground stars is not known, we want
to estimate both AV and f in a given patch of the sky.
[Although f can be taken to be constant in many cases,
changes on this quantity have to be expected for different
regions of large cloud complexes because of the geometry
of the cloud and of changes of star densities due to the
Galactic structure.] For this, we insert Eq. (17) in Eq. (38),
thus obtaining a simple generalization of Eq. (40):
σ
σ(0) = f + (1 − f )10−αkλAV .
(43)
Hence, we can in principle use two different measurements
of ¯σ in two different bands to deduce both f and AV on
the region considered. In practice, it is preferable to follow
the approach described below in Sect. 4.
It is interesting to investigate in more detail the esti-
mator derived from Eq. (43), i.e.
AV = −
1
αkλ
log10(cid:20) N − Aσ(0)f
Aσ(0)(1 − f )(cid:21) ,
(44)
where N is the number of stars found in the region inves-
tigated, and A is its area. Assuming that σ(0) is measured
without significant errors (this is possible by using a large
control field), we can deduce the expected error on AV by
noting that N follows a Poisson distribution with average
[cf. Eq. (43)]
hNi = Aσ(0)(cid:2)f + (1 − f )10−αkλAV(cid:3) .
Using a first order approximation (valid for small relative
errors) we find
(45)
Var(cid:0) AV(cid:1) =(cid:18)
1
αkλ ln 10 ·
1
hNi − Af σ(0)(cid:19)2
hNi .
(46)
1
0.8
0.6
0.4
0.2
0
y
t
i
l
i
b
a
b
o
r
p
−5
0
5
10
15
20
AV
(from the
Fig. 1. The
cumulative probability distribution for
AV . The various curves are relative to a true ex-
tinction AV = [0, 2, 4, 6, 8, 10]
to
the right); the average values measured are instead
h AV i = [0.24, 2.33, 4.47, 6.64, 8.79, 10.76] (the scatters
around these values ranges from 1.9 to 4.3). In all cases we
assumed α = 0.34, kλ = 0.175, f = 0.1, and Aσ(0) = 20.
Along the curves the dots mark the actual possible val-
ues of AV that can be measured for different values of the
observed number of stars N .
left
4
2
0
)
V
A
(
r
r
E
,
V
A
−
i
V
A
h
−2
−4
−6
−8
Aσ(0) = 10
Aσ(0) = 20
Aσ(0) = 50
0
5
10
AV
15
20
Fig. 2. The bias (thick lines) and error (thin lines) on AV
evaluated using Eq. (44) for different values of Aσ(0). For
this figure we used the same parameters as in Fig.1, i.e.
α = 0.34, kλ = 0.175, and f = 0.1.
In case of vanishing f (i.e., if all stars are background to
the cloud), this expression takes a simpler form
Err(cid:0) AV(cid:1) =qVar(cid:0) AV(cid:1) =
1
1
αkλ ln 10
phNi
=
11.4
,
phNi
(47)
where the last expression is valid for the K band (cf.
Tab. 1).
Further statistical properties of the estimator (44) can
be better evaluated using numerical methods. Figure 1
shows, as an example, the expected cumulative probabil-
8
M. Lombardi: Optimal column density measurements from multiband near-infrared observations
ity distributions for AV for some typical cases. As ex-
pected, the cumulative distributions reach 0.5 around the
true value for AV , but the expected measurement errors
can be very large, especially at high column densities.
The numerical analysis of Eq. (44) also shows that the
estimator is significantly biased. In general, in the limit
σA ≫ 1, the estimator is biased toward large values of
AV , i.e. (cid:10) AV(cid:11) > AV ; the opposite happens for σA ∼ 1
(see Fig. 2). This change in behavior can be understood
by observing that, if AV is large, then Aσ is small and
thus we simply do not have enough background stars to
probe high column densities.
Interestingly, Eq. (46) is a reasonably good approxi-
mation of the true variance of AV . Typically, this approx-
imation slightly underestimates the true variance of AV ,
but again the opposite happens for low values of σA.
3.2. Near-infrared color excess (NICE and NICER)
The dust present in molecular clouds produces differ-
ent amounts of obscuration in different color bands (see
Table 1), so that background stars appear reddened.
Hence, the color excess in the red part of the spectrum
of background stars can be used to measure the extinc-
tion along the line of sight.
Historically, the color excess technique has been im-
peded by the limited instrumental sensitivity to small re-
gions (Jones et al. 1980; Frerking et al. 1982; Jones et al.
1984). However, in the early 1990s, with the advent of
near-infrared arrays, it became possible to accurately mea-
sure the NIR magnitudes of many stars from single point-
ing observations. This new technology was first exploited
by Lada et al. (1994), and since then has been successfully
applied to many molecular clouds (e.g. Horner et al. 1997;
Lada et al. 1999; Alves et al. 2001).
Lada's technique (called "near-infrared color excess"
or Nice) is based on measurements of a NIR color (e.g.,
H − K) of many stars. Since stars have relatively well
defined colors in the infrared, a significant intervening ex-
tinction can be detected as a reddening. Note that a key
point of the Nice method (and of similar methods based
on the reddening of stars; see, e.g., Schultheis et al. 1999)
is the assumption that all stars belong to a homogeneous
population.
The Nice method can be quantitatively described us-
ing the notation introduced so far. In particular, from
Eq. (29) we have
pm(m) = p(0)
m (m − kAV ) ,
(48)
where we have denoted p(0)
m (m) the l.h.s. of Eq. (29) for
AV = 0. Naively, Eq. (48) can be used to obtain a simple
estimate of the intervening infrared extinction AV . For
example, we find
hmi = hmi(0) + kAV ,
(49)
where again we used the superscript (0) to denote quan-
tities measured in a control region where AV = 0.
Equation (49) is better rewritten in terms of star colors.
Calling χλ = mλ − m1 and κλ = kλ − k1, we have
hχi = hχi(0) + κAV .
(50)
This equation, applied to a single color, is essentially the
Nice technique. More precisely, this technique uses the
simple average of a set of N angularly close stars to eval-
uate the column density:
AV =
1
N
N
Xn=1
χn − ¯χ(0)
κ
,
(51)
where ¯χ(0) is an estimate of hχi(0) (obtained, e.g., by mea-
suring the star colors on a control field where presumably
AV ≃ 0). As an example, consider the Nice method ap-
plied to the χ = H − K color. Since stars have a typical
scatter of 0.09 mag in this color, we expect an error on AV
of Err(cid:0) AV(cid:1) ≃ 1.4/√N . Hence, even in the presence of sig-
nificant photometric errors, the Nice method gives signif-
icantly more accurate results than the star count method
[cf. Eq. (47)].
As shown by Lombardi & Alves (2001), one can indeed
take full advantage of observations carried out in differ-
ent bands to obtain more accurate column density mea-
surements. The improved technique, called Nicer (Nice
Revised) optimally balances the information from different
bands and different stars. As a by-product of the analysis,
Nicer also allows us to evaluate the expected error on the
column density map, which is useful to estimate the sig-
nificance on the detection of substructures and cores. The
Nicer technique can be described using the following sim-
ple argument. Equation (50) written above can be taken
to be a system of (Λ − 1) equations to be approximately
solved for AV , the approximation being made necessary
because we can only measure hχi and hχi(0) with limited
accuracy. The "best" solution for AV can been obtained
by minimizing the chi-square quantity
χ2 =
N
Xn=1(cid:2) χn− ¯χ(0)−κAV(cid:3)T
(C +E)−1(cid:2) χn− ¯χ(0)−κAV(cid:3) ,
(52)
Consistently with the notation used above in Eq. (34), we
called C the covariance matrix of the star colors; moreover,
the symbol E was used to denote the covariance matrix of
measurement errors [the two covariance matrices have to
be added up in Eq. (52) in order to properly estimate the
expected scatter on star colors]. The best estimate of AV ,
obtained by minimizing the χ2 of Eq. (52), is precisely the
Nicer estimator.
Both the Nice and Nicer estimators appear to be
unbiased provided that: (i) there are no foreground stars
and (ii) the measured colors χn are unbiased estimates of
the true colors χn.
In reality, even if the two conditions considered above
are satisfied, both color excess methods can still be biased
because of selection effects introduced by the completeness
M. Lombardi: Optimal column density measurements from multiband near-infrared observations
9
0.15
0.1
i
V
A
h
−
V
A
0.05
0
0
30
20
i
V
A
h
10
0
f = 0.01
f = 0.03
f = 0.1
f = 0.3
10
AV
20
30
Fig. 3. The bias for the Nice method due to the different
completeness at low and high column densities. This plot
is evaluated by generating 100 000 stars for various values
of AV (marked as dots). Note that, as described in the
text, for AV ≃ 6.7 we observe a rapid increase in the bias
due to the change between a selection in the K band and
a selection in the H band (see Fig. 4). The small scale
oscillations on the plot are due to numeric effects.
0
10
20
AV
30
40
Fig. 5. The bias for the Nice method due to the contam-
ination by foreground stars. Solid lines represents, for dif-
ferent fractions of foreground stars, the median of the dis-
tribution of the measured column densities. Dashed lines,
instead show the expectation value for the median of AV
obtained from the measurements of 15 reddenings.
H
=
1
5
.
5
5
.
5
1
=
K
AV = 12
K
−
H
AV = 6
AV = 0
K
Fig. 4. A graphic explanation of the bias plotted in Fig. 3.
At low column densities, we select stars mainly on their
K magnitude, while at high column densities we mainly
select using their H magnitude. As a result, the average
intrinsic color (dotted line) of the observed stars (gray
bands) changes toward lower H − K values.
function. To understand this point let us make a simple
example. Suppose that we carry out our observations in
two bands, λ1 and λ2, and that both completeness func-
tions cλ(mλ) are not vanishing only on a narrow magni-
tude range. In this case we would always have χ ≃ χ(0)
(because a star is observed in both bands only if m1 and
m2 are inside the narrow detection window), and thus we
would always measure AV ≃ 0, independently of the real
column density. Although unrealistic, this example shows
that we must expect a bias for the Nice method; the ar-
gument for the Nicer method is similar and leads to the
same conclusion.
The bias of these two methods depends on the details
of the probability distribution pM (M ) and of the com-
pleteness functions cλ(mλ). However, as an example, we
evaluated the expected bias for various values of the col-
umn density and for the typical probability distributions
that we expect for the 2MASS catalog. As shown in Fig. 3,
the bias in the case considered appears to be limited, be-
low 0.2 in AV , and has a characteristic shape: it vanishes
for AV = 0, increases quickly for AV ∼ 7, and finally
saturates for AV > 11. This behavior has a simple expla-
nation:
-- Since the colors of unreddened stars are evaluated us-
ing a control field (where supposedly AV = 0), the bias
has to vanish for AV = 0.
-- The general trend of bias on AV can be understood
with the help of Fig. 4. At low AV , stars that are ob-
served in the K band are almost certainly also observed
on the H band, because of the values of the limiting
magnitudes (approximately 14.9 in H and 14.3 in K)
and of the average star colors (hH−Ki = 0.18). When,
instead, the reddening is large, we have the opposite
situation (because K is less affected by reddening than
H). This different selection at different column densi-
ties is the source of the observed bias.
-- Finally, at large column densities an asymptotic value
is reached because now only the H band is used to
select stars.
A more serious problem is related to the contamina-
tion by foreground stars, which can strongly bias our re-
sults in the direction of lower column densities. For regions
10
M. Lombardi: Optimal column density measurements from multiband near-infrared observations
0.23
0.96
y
t
i
l
i
b
a
b
o
r
p
e
v
i
t
a
l
e
r
9.31
0
10
20
30
AV
Fig. 6. The effect of a foreground star contamination
f = 0.1 on the distribution of column densities for differ-
ent values of AV . From the bottom to the top we show, for
AV ∈ {0, 10, 20, 30}, both the distribution of background
stars (solid plot, gray filled), and foreground stars (dashed
plot). Both distributions are taken to be Gaussian, the
first centered on AV , and the second on 0. As the col-
umn density increases, the relative contribution of fore-
ground stars also increases because of selection effects. At
AV ≃ 16, they become predominant and the median of the
whole distribution quickly moves toward the foreground
distribution.
with low extinction, where the expected number of fore-
ground stars is small compared to the expected number of
background stars, foreground star contamination is usu-
ally reduced by using a median estimator for AV , i.e. by
averaging the individual column densities measured in the
direction of each star with the median instead of the sim-
ple mean (e.g. Cambr´esy et al. 2002; Lombardi & Alves
2001). As shown in Fig. 5, this simple technique is very
effective and leads to almost unbiased results at low col-
umn densities (see Appendix A for a statistical discussion
on the median estimator). However, for a given fraction of
foreground stars f (which, as described above, is evaluated
in regions with negligible extinction) there is a threshold
for AV after which the cloud extinction makes the ex-
pected number of background stars smaller than the ex-
pected number of foreground stars. This threshold can be
evaluated using Eq. (43) and is
AV =
1
αkλ
log10
1 − f
f
.
(53)
Because of the way the median estimator is defined, we
observe in Fig. 5 an abrupt change on the measured col-
umn density AV close to this value (see also Fig. 6). Note
that the relatively smooth transition observed in Fig. 5
(solid lines) is due to the intrinsic scatter in the star col-
ors; indeed, in absence of any scatter, we would observe
an "instantaneous" change from h AV i = AV at low col-
umn densities to h AV i = 0 at high column densities. In
Fig. 5 we also plot the expectation value of a more inter-
esting quantity, the median over N = 15 measured col-
umn densities. This quantity differs from the median over
the whole distribution because of the statistical variations
on the local number of foreground stars. In other words,
since N = 15 is a relatively small number, in different
realizations of our simulations we can have a significantly
different number of foreground stars. For example, even at
low column density, we can have a large fraction of fore-
ground stars; similarly, even at very large column density,
there is a finite probability to have the majority of stars
in the background. As a result, the dashed line in Fig. 5
has a much smoother transition around the value given
by Eq. (53). This is at the same time a good and a bad
news. From one side, this means that we can be signifi-
cantly biased for values of AV smaller than the threshold
value of Eq. (53); from the other side, this also means that
we can still partially make use of the median estimator at
relatively large column densities.
For regions with very high column density, it is nor-
mally quite easy to identify and remove foreground stars
because of their anomalous colors with respect to the red-
dened background stars (see, e.g., Alves et al. 2001). Some
authors (e.g. Cambr´esy et al. 2002) make use of this infor-
mation to also remove stars in less dense regions using the
following strategy. The angular density σf of foreground
stars is determined using dense regions (where the fore-
ground/background identification is easy); then, for any
region of the cloud, the k bluer stars are thrown away,
where Nf = σfA is the expected number of foreground
stars in the analyzed region (deduced from the foreground
density measured in the dense regions). This technique is
quite simple and reasonably effective, but unfortunately
introduces a bias at low extinctions. Consider, indeed,
the limiting case where we have a negligible extinction
AV ≃ 0. The distinction between foreground and back-
ground stars is in this situation ambiguous, and thus the
Nf bluer objects will likely include also some background
stars. Hence, the results will be biased toward higher col-
umn density. The exact evaluation of the bias is non triv-
ial, but can be carried out using the techniques described
in App. A. Here we report only an approximated result
valid for AV ≃ 0 [see Eq. (A.15)],
(54)
Nf
2N
,
√2πErr(cid:0) AV(cid:1)
h AV i ≃
where Err(cid:0) AV(cid:1) is the average error of the measured ex-
tinction from a single star. Note that the result given in
Eq. (54) can be taken to be an upper limit for the bias,
since we expect this to decrease at high column densities,
where the identification of foreground stars is secure.
4. A maximum-likelihood approach
From the discussion above, it is clear that both the star
count and the Nice(r) methods can produce unsatisfac-
tory results. On one hand, the star count technique has a
low signal-to-noise ratio and produces significantly biased
results at high column densities (see Fig. 2). On the other
M. Lombardi: Optimal column density measurements from multiband near-infrared observations
11
hand, the color excess technique is more sensitive and has
a smaller bias, but can be severely affected by the con-
tamination of foreground stars, especially for large values
of AV .
As pointed out by Cambr´esy et al. (2002), we can take
advantage of the different behavior of the star count and
the color excess techniques in the various column density
regimes to partially solve the problem of the contamina-
tion by foreground stars. In particular, Cambr´esy et al.
note that it is better to use the color excess method at
low column densities because of its higher sensitivity, and
to switch to the star count method at very large column
densities where the Nice method is completely unreliable.
The optimal "turning point" where we need to change the
technique can be evaluated empirically by comparing the
individual results of the two methods at different column
densities.
The solution proposed by Cambr´esy et al. has the sig-
nificant advantage of being relatively simple to implement
and reasonably effective, but clearly is suboptimal in many
respects:
-- The choice of the "turning point" and of the matching
strategy is to some extent arbitrary.
-- The overall estimate of AV still remains significantly
biased at high column densities because at large AV
the star counts are used (and this method has a non-
negligible bias, see Fig. 2).
-- The density of foreground stars must still be deter-
mined separately and is taken to be constant on the
whole field.
Using the theoretical framework developed so far, it is
possible to design and implement an efficient maximum-
likelihood approach to the problem.
4.1. Likelihood
Suppose again that in a region of the sky of area A we
observe in various bands N stars with magnitudes { mn}.
Assuming that the area of the sky is small enough so that
there are no significant changes on the relevant parameters
of the problem (such as the unreddened density σ(0), local
expected density σ, or the reddening probability distribu-
tion pAV (AV )), then we can easily evaluate the joint prob-
ability distribution for such a star configuration. First, we
note that the number of stars inside the region follows a
Poisson distribution with average Aσ:
pN (N ) = e−Aσ (Aσ)N
N !
.
(55)
The joint star probability distribution, i.e. the likelihood,
is given by
L(cid:0){ mn}(cid:1) = pN (N )
=
e−AσAN
N !
p m( mn)
σ
p m( mn) .
N
Yn=1
Yn=1
N
Note that L depends on unknown quantities, such as
pAV (AV ), through pN (N ), p m( m), and σ: hence, assum-
ing that there is no prior for these unknown quantities,
we can obtain an estimate of them by maximizing L or,
equivalently, ln L. In the following subsections we will in-
vestigate in more detail this maximum-likelihood estima-
tor.
4.2. Implementation
We implemented the maximum-likelihood estimator using
the simplification described above. In particular, we used
the forms (34) and (35) for the source magnitude proba-
bility distribution pM (M ). A simple calculation gives the
following relationships between the coefficients of Eq. (34)
and Eq. (35):
Pab = C−1
ab − δ1aXa′
+ δ1aδ1bXa′,b′
C−1
a′b − δ1bXb′
C−1
a′b′ ,
C−1
ab′
(57)
(58)
C−1
a′bχb ,
ab χb + δ1aXa′
Qa = −δa1α ln 10 − C−1
R = −2 ln n + C−1
ab χaχb .
One of the advantages of the quadratic expression (35)
is that we can write the effects of reddening as a simple
transformation of the three quantities P , Q, and R. In
particular, the transformation M 7→ M + kAV can be
rewritten as
P 7→ P ,
R 7→ R − 2AV QTk + A2
Q 7→ Q − AV P k
V kT P k .
(60)
(59)
Hence, reddening does not change the functional form of
the probability distribution (35) but only the three pa-
rameters involved.
For the following discussion, it is also convenient to in-
troduce a new vector µ ≡ (M , 1) = (M1, M2, . . . , MΛ, 1),
and a (Λ + 1) × (Λ + 1) matrix S defined as
P11 ··· P1Λ Q1
...
...
PΛ1 ··· PΛΛ QΛ
Q1 . . . QΛ R
QT R(cid:19) =
S−1 =(cid:18) P Q
Then we can write the quadratic form appearing in
Eq. (35) simply as
(61)
. . .
...
.
pM (M ) = exp(cid:20)−
µTS−1µ
2
(cid:21) .
(62)
More importantly, the action of measurement errors can
be described as simple transformations of S, provided that
the measurement error on each band can be described as a
simple Gaussian with vanishing average and variance v. In
other words, assuming that the errors can be represented
as a convolution with the Gaussian kernel
(56)
pǫ(ǫ) =
1
(2π)Λ/2QΛ
λ=1 vλ
exp(cid:20)−
Λ
Xλ=1
ǫ2
λ
2v2
λ(cid:21) ,
(63)
12
M. Lombardi: Optimal column density measurements from multiband near-infrared observations
then the convolution pM ∗ Σ can be written as
(pM ∗ Σ)(µ) =r det S
(cid:21) ,
exp(cid:20)−
µTT −1µ
2
where T = S + diag(v2
det T
1, v2
2, . . . , v2
Λ, 0).
(64)
In conclusion, the simple implementation of the
maximum-likelihood method considered here can be sum-
marized in the following items:
1. The quantities P , Q, and R appearing in Eq. (35) are
determined in a control field free from any extinction.
Similarly, for each band the photometric errors, i.e. the
functions pǫλ( mλmλ), are calculated.
2. A cloud model (for example, the thin cloud approx-
imation) is chosen, and the relevant parameters (for
example, AV and f ) are taken to be unknown and
constant on the field.
3. For a given combination of the cloud parameters, the
likelihood function (56) is evaluated using the vari-
ous observed magnitudes. Note that the evaluation of
the expected star density σ and of the star probabil-
ity p m( m) when the star is undetected in one or more
bands is non-trivial and requires integrations over the
completeness functions [cf. Eq. (32)].
4. The last step is repeated for different values of the
cloud parameters and the ones corresponding to the
minimum of the likelihood function are taken as best
estimates.
5. The local curvature of L (i.e., the matrix of its second
derivatives) is used to estimate the errors on the cloud
parameters.
Clearly, one key point here is the speed of the likeli-
hood function, which needs to be evaluated several times
in our maximum-likelihood approach. In our implementa-
tion, the likelihood function has been optimized by per-
forming the relevant integrations (cf. point 3 above) using
appropriate bounds. In other words, when an integration
of p m( m) was requested, we estimated the area in the
magnitude space where this function was significantly dif-
ferent from zero, and performed the integral inside that
area (as opposed to performing the integral over the whole
parameter space). This optimization was found to have a
significant impact on the overall speed of our implemen-
tation.
4.3. Simulations
reliability and effectiveness of
The
the maximum-
likelihood approach were assessed through extensive nu-
merical simulations. The simulations were designed to
reproduce with reasonable accuracy the 2MASS near-
infrared data. We simulated star observations in three
bands, J, H, and K, and used the various parameters
described in Table 2.
We initially considered an area of the sky A and a thin
cloud characterized by a fraction of foreground stars f [cf.
Eq. (18)] and a reddening AV . We randomly generated
there stars inside this area using the following recipe:
2(cid:2)ln L − ln Lmin(cid:3)
150
100
50
0
6
8
AV
10
12
14
1.0
0.8
0.0
0.2
0.6
0.4
f
Fig. 7. Log-likelihood surface plot. The plot shows the
logarithmic of the likelihood ratio as a function of the two
unknown parameters AV and f (the real values used for
the simulation are 10 and 0.2 respectively). The simula-
tion has been carried out using σA = 25, but the actual
number of stars generated were N = 19 (because of the
Poisson noise on the number of stars). On the bottom
we also plot contours corresponding to 68.2%, 95.5%, and
99.7% confidence level regions. Note that the likelihood is
very smooth and has only a single well defined minimum.
1. We evaluated the expected local star density σ using
Eqs. (31) and (32). We found that the needed inte-
grations could be performed more efficiently using a
Monte-Carlo technique.
2. We calculated the effective number of stars N by gen-
erating a random integer distributed according to a
Poisson distribution with average Aσ.
cedure:
(a) We generated the unreddened magnitudes in the
3. For each of the N star we adopted the following pro-
three bands according to Eq. (34).
(b) We then uniformly generated a random number in
the range [0, 1], and considered the star to be in
Par.
α
χ1
χ2
C11
C12
Value Description
0.34
0.18
0.82
0.0078 Variance h(H − K − χ1)2i
0.0112 Covariance
Slope of number counts
Average color hH − Ki
Average color hJ − Ki
h(J − K − χ1)(H − K − χ2)i
C22
k1
k2
k3
mlim
mlim
mlim
3
v1
2
1
v2
v3
0.0375 Variance h(J − K − χ2)2i
0.112
Reddening law in K band
Reddening law in H band
0.175
Reddening law in J band
0.282
Limit magnitude in K band
14.3
Limit magnitude in H band
14.9
15.8
Limit magnitude in J band
Average photometric error
0.05
h( m1 − m1)2i1/2 on K band
Average photometric error
h( m2 − m2)2i1/2 on H band
Average photometric error
h( m3 − m3)2i1/2 on J band
0.05
0.05
ref. Eq.
(34)
(34)
(34)
(34)
(34)
(34)
(8)
(8)
(8)
(20) & (27)
(20) & (27)
(20) & (27)
(20)
(20)
(20)
Table 2. The common parameters used for all numerical
simulations.
M. Lombardi: Optimal column density measurements from multiband near-infrared observations
13
30
(cid:3)
20
n
i
m
L
n
l
−
L
n
10
l
(cid:2)
2
ML
Nice (mean)
Nicer (mean)
Nice (median)
Nicer (median)
0.2
0.1
0
V
A
−
i
V
A
h
−0.1
−0.2
0
6
8
10
AV
12
14
Fig. 8. The log-likelihood (ratio) as a function of AV for
different source densities, extremized for f . The largest
curve is obtained using σA = 10 (in this particular re-
alization we had N = 8); the mid curve using σA = 20
(N = 12), and the most peaked curve using σA = 40
(N = 35). In all cases the true value of the column den-
sity was AV = 10, and the fraction of foreground stars was
f = 0.2. Note that all curves are very well approximated
by parabolas. The intersections of the log-likelihood func-
tions with the three dashed lines mark the 68.2%, 95.5%,
and 99.7% confidence level regions.
the foreground if this number was smaller than f
[defined according to Eq. (18)].
(c) Background stars were reddened by adding kλAV
to each magnitude; the magnitudes of foreground
stars were left unchanged.
(d) We added photometric errors to all stars; these, for
simplicity, were taken to be Gaussian distributed
with standard deviation vλ = 0.05 independent of
the band and of the original magnitude.
(e) For each band, we uniformly generated a random
number in the range [0, 1], and took the star to be
detected in the band if this number was smaller
than the completeness function cλ( mλ). In the
simulations discussed here we used for simplicity
Heaviside functions for the completeness functions.
(f) We finally retained the star if it was detected in
at least one band; otherwise, we repeated all steps
above from point (a).
In summary, at the end of a single star generation we had
for each star the three magnitudes in the bands J, H, and
K (with possibly some magnitudes mλ = null) and the
associated measurement errors vλ.
We then used this dataset to test the reliability and ef-
ficiency of the maximum-likelihood estimator, and to com-
pare it with the other column density estimators consid-
ered in Sect. 3. The maximum-likelihood method was im-
plemented as described in Sect. 4.2, and was tested against
the data generated as described in the items above.
Figure 7 shows the log-likelihood surface plot obtained
in a typical simulation. The surface appears to be very
0
10
AV
20
30
Fig. 9. The bias h AV i − AV of various column density
measurement methods. The bias is evaluated from the av-
erages of 1 000 simulated fields with no foreground contri-
bution (f = 0) and with average number of stars σA = 25.
The maximum-likelihood method, marked in the legend as
ML for brevity, has negligible bias, especially for AV < 20.
smooth with a well defined minimum, an essential con-
dition for the reliability of the maximum-likelihood ap-
proach. Moreover, the log-likelihood function is found to
have approximately a parabolic shape, which further sim-
plifies the interpretation of the results obtained. For exam-
ple, this property allows us to use the likelihood ratio (see,
e.g., Eadie et al. 1971) to draw confidence level regions on
the parameter space (see the contours of Fig. 7).
Figure 8 represents the log-likelihood as a function of
AV for three different datasets. The figure was obtained
by minimizing the log-likelihood function with respect to
f for each value of AV in the range [5, 15], and by plotting
the value of this function.
In order to assess more quantitatively the merits of
the maximum-likelihood method, we compared the statis-
tical properties of various column density estimators. In
particular, we simulated a large number of "observations"
using the technique described above, and we studied the
distribution of the column densities estimated using the
maximum-likelihood method, the Nice, and the Nicer
methods. Simulations were carried out using a thin cloud
with true extinction in the range AV ∈ [0, 30] and with
different values of the foreground fraction f . For the Nice
and Nicer estimators we used both the simple mean and
the median of the individual extinction measured for each
star; moreover, assuming that the density of foreground
stars σf could be determined separately, we discarded in
each simulation the σfA bluer stars, and used only the re-
maining (redder) stars in the analysis. Note that in some
cases, for large values of AV and relatively large values of
f we had no usable star for the Nice and Nicer analysis;
in other words, all stars left after the foreground selec-
tion had only one band available (typically the K band).
In this case we just obtained a lower limit on AV by us-
ing the redder usable star (even if this star was originally
considered to be in the foreground).
14
M. Lombardi: Optimal column density measurements from multiband near-infrared observations
ML
Nice (mean)
Nicer (mean)
Nice (median)
Nicer (median)
0.9
0.8
0.7
0.6
0.5
0.4
0.3
)
V
A
(
r
r
E
0
10
20
30
AV
of various col-
umn density measurement methods, evaluated as in Fig. 9.
Fig. 10. The total error (cid:10)(cid:0) AV − AV(cid:1)2(cid:11)1/2
1
0
V
A
−
i
V
A
h
−1
−2
ML
Nice (mean)
Nicer (mean)
Nice (median)
Nicer (median)
ML
Nice (mean)
Nicer (mean)
Nice (median)
Nicer (median)
)
V
A
(
r
r
E
6
4
2
0
0
10
20
30
AV
Fig. 12. As for Fig. 10, but with f = 0.02.
1
0
V
A
−
i
V
A
h
−1
−2
ML
Nice (mean)
Nicer (mean)
Nice (median)
Nicer (median)
0
10
20
30
AV
Fig. 11. As for Fig. 9, but with f = 0.02.
0
10
20
30
AV
Fig. 13. As for Fig. 9, but with f = 0.05.
Figures 9 and 10 show, respectively, the bias and the
total error obtained from the three methods considered
here for AV ∈ [0, 30] and f = 0. From these plots it is
evident that the maximum-likelihood estimator does not
have any significant bias up to AV = 20 and a very small
one (of the order of 0.1) for larger column densities. Since
for f = 0 we never have foreground stars, the bias of the
Nice and Nicer techniques does not change if we use a
mean or a median estimator. Note also that the bias in
Fig. 9 for these two methods is the one discussed in detail
above (cf. Fig. 3). Regarding the total error, we observe a
steady increase of it for large values of AV . This can be
explained by noting that, although the average number
of stars σA = 25 is kept constant for all column densi-
ties, when AV is large most stars are only detected in the
K band and thus do not provide reddening information.
Figure 10 also shows that the maximum-likelihood method
is clearly superior, although Nicer (with the mean esti-
mator) also performs well. As expected, both median es-
timators are more noisy than the simple mean (which, for
f = 0, is optimal).
The situation changes quite dramatically when f > 0.
Figures 11 -- 16 show the bias and the total error of the var-
ML
Nice (mean)
Nicer (mean)
Nice (median)
Nicer (median)
)
V
A
(
r
r
E
8
6
4
2
0
0
10
20
30
AV
Fig. 14. As for Fig. 10, but with f = 0.05.
ious methods for increasing values of f . A careful study of
these results can reveal several interesting characteristics
of the Nice and Nicer techniques.
Let us initially focus on the bias plots, shown in
Figs. 11, 13, and 15. At low column densities, i.e. for
AV < 10, we find again the bias described in Sect. 3.2
M. Lombardi: Optimal column density measurements from multiband near-infrared observations
15
1
0
−1
−2
−3
−4
−5
V
A
−
i
V
A
h
ML
Nice (mean)
Nicer (mean)
Nice (median)
Nicer (median)
0
10
AV
20
30
Fig. 15. As for Fig. 9, but with f = 0.10.
ML
Nice (mean)
Nicer (mean)
Nice (median)
Nicer (median)
)
V
A
(
r
r
E
12
10
8
6
4
2
0
0
10
AV
20
30
Fig. 16. As for Fig. 10, but with f = 0.10.
and plotted in Fig. 3 (see also above Fig. 9 for the case
f = 0). At median column densities, i.e. for AV ∼ 20 (or
AV ∼ 30 for f = 0.02), the bias can be explained by the
selection effect due to the correction of foreground stars.
As explained above [see Eq. (54)], removing the Nf = σfA
bluer stars introduces a systematic error on the extinc-
tion estimate. This bias is positive (i.e. toward larger ex-
tinctions) and can be as large as 1 mag in AV . Finally,
at high column densities (AV ∼ 30) both methods sys-
tematically underestimate the extinction. This is due to
the heavy contamination by foreground stars present at
these large values of reddening. To better understand this
point, we note that the subtraction of the Nf bluer stars
is a simplistic approximation because the actual number
of foreground stars is not fixed (it is a Poisson random
variable with average Nf ). Depending on the number of
foreground stars with respect to Nf we can have three
different situations:
-- If the number of foreground stars is exactly Nf , then
at high column densities no bias is introduced and the
estimate of AV is correctly performed;
-- For a simulation with a number of foreground stars
smaller than Nf , a small bias toward higher column
densities is expected, because some of the bluer back-
ground stars are discarded;
-- Finally, if the number of foreground stars is underesti-
mated, a very large bias toward smaller column densi-
ties is introduced.
Clearly, for large values of AV the third effect is expected
to dominate. Indeed, Figs. 13 and 15 show that both the
Nice and Nicer methods significantly underestimate the
reddening for large values of AV . In theory, as mentioned
above, when applying the Nice or Nicer technique to
high extinction regions, it should be relatively straightfor-
ward to identify foreground stars by their color, and thus
the bias of these method could be smaller than suggested
by the plot shown here. Note that apparently Nicer
presents a larger bias compared to Nice. This is due to
the larger flexibility of the Nicer method, which is able
to obtain a column density estimate when any two of the
three bands are available. As a result, Nicer is more af-
fected by the contamination by foreground stars described
in the items above. Indeed, our simulations show that if
we force Nicer to use only stars with observed H and K
bands (i.e. essentially the same stars as the ones used by
Nice), its bias and its noise are drastically reduced and
become compatible with the ones of Nice.
Figures 12, 14, and 16 show that the total error of the
Nice and Nicer methods increases very rapidly at large
column densities, where the contamination by foreground
stars is very likely; the maximum-likelihood estimator, in-
stead, has an almost flat error curve. Hence, our novel
method is able to "recognize" the presence of foreground
stars; moreover, the inclusion of the background density in
the likelihood expression allows this estimator to "switch"
to the number count technique in regions with large ex-
tinction. Figure 12, in particularly, shows that even in ex-
treme cases with a large foreground star contamination
the maximum-likelihood method is still very reliable and
accurate. To better appreciate this point, we note that, in
our simulations, for AV = 25 and f = 0.2 on average only
one tenth of the ∼ 25 stars are background to the cloud.
4.4. Limitations
The maximum-likelihood approach to the extinction mea-
surements presents clear advantages with respect to the
standard techniques in the simplified framework consid-
ered in this section (uniform AV over the cloud patch
analysed, thin-cloud approximation, simple model for the
star intrinsic magnitude distribution). One could thus le-
gitimately ask whether the maximum-likelihood technique
is applicable to more realistic and complex situations.
4.4.1. Small-scale inhomogeneities
Most clouds present clear sign of substructure at dif-
ferent scales and a statistical analysis of the radio and
NIR observations seems to indicate the that these ob-
jects can be well described in terms of turbulent mod-
16
M. Lombardi: Optimal column density measurements from multiband near-infrared observations
els (see, e.g. Miesch & Bally 1994; Padoan et al. 1997). In
presence of significant inhomogeneities, most methods (in-
cluding the maximum-likelihood one) are expected to be
biased toward low values because the background stars
will no longer be randomly distributed in the patch of
the sky used to estimate the local extinction value, but
will be preferentially detected in low-extinction regions
[cf. Eq. (43)]. Although a detailed discussion of this effect
is behind the scope of this paper, it is worth considering
the following points:
-- For the HK Nice, for the JHK Nicer, and for the
JHK maximum-likelihood methods, small scale inho-
mogeneities become important when the local varia-
tions of AV on the patch of the sky considered are
of order ∆ = 1/(αk2 ln 10), where k2 = AH /AV
[cf. Eq. (8)]. For typical 2MASS observations we find
∆ ≃ 7.3 mag, and hence all methods should still be ap-
plicable to the analysis of regions with relatively low
column densities (approximately AV < 10 mag).
-- The effect of substructures is studied in detail in a
forthcoming paper (Lombardi 2005), where it is also
presented a method to avoid the bias introduced by
small-scale inhomogeneities. The application this novel
technique to a Nicer map of the Pipe nebula con-
firms the expectations summarized in the last item.
In particular,
for the Pipe nebula the "standard"
Nicer has a negligible small bias (below 0.2 mag) for
AV < 10 mag, while the bias increases dramatically
for larger extinction values (e.g., it reaches 1 mag at
AV = 15 mag). A similar bias behavior is expected
for the maximum-likelihood technique described in this
section.
-- Since small-scale inhomogeneities are believed to be
due to turbulent motions (Larson 1981; Padoan et al.
1997; Heyer & Brunt 2004), the probability distribu-
tion pAV (AV ) is expected to be a log-normal:
(ln AV − M )2
pAV (AV ) =
1
√2πSAV
exp(cid:20)−
2S2
(cid:21) . (65)
Independently from the exact form of pAV (AV ), the
maximum-likelihood approach can be used in this more
general framework to obtain the relevant parameters
of pAV (e.g., M and S in the case of the log-normal
of Eq. (65)); moreover, the parameters estimated are
asymptotically unbiased (see Eadie et al. 1971). We
will consider the use of a non-trivial extinction proba-
bility distributions similar to the one of Eq. (65) in a
follow-up paper.
4.4.2. Star magnitude distribution
In the implementation of the maximum-likelihood tech-
nique discussed in this section, we used the simple model
for the magnitude probability distribution pM (M ) [cf.
Eq. (34)]. However, the functional form of pM (M ) used
can be inaccurate in describing real data for several rea-
sons:
-- Different stellar populations can produce multiple
peaks in the color distribution of stars. For example,
giant and dwarf stars produce two distinct peaks in
the J − H histogram (the two peaks are also clearly
visible in a JHK color-color diagram).
-- Different stellar populations can also have different
slopes of the number counts, which could different sig-
nificantly from the "nominal" value α ≃ 0.34 (see, e.g.
Cambr´esy et al. 2002). If the various stellar popula-
tions also have different intrinsic mean colors, then the
simple model (34) could lead to inaccurate extinction
measurements.
-- The average NIR colors of stars in regions free of ex-
tinction are not completely independent of the star
luminosity. For example, a color-magnitude plot of
2MASS stars shows that the average J − K color
slightly increases as the K magnitude increases. Since
this effect is rather small, the associated bias is prob-
ably negligible in most cases; moreover, this effect can
be included in the expression (35) with a suitable
choice of the coefficients P , Q, and R.
All issues described above are strictly related to the sim-
plified description for the probability pM (M ) used here,
and not to the maximum-likelihood method itself. In other
words, it is possible (and relatively easy) to implement
a maximum-likelihood estimator based on a more realis-
tic probability distribution for the star magnitudes (e.g.,
synthetic stellar population models; see Robin et al. 2004;
Jarrett et al. 1994). For this purpose, we note that the
most computationally effective way to generalize pM (M )
is by writing it as a linear combination of functions of
the form (34) (with each function representing, de facto,
a different star population).
4.4.3. General remarks
On of most significant drawback of the maximum-
likelihood approach is its speed. The implementation used
in this paper is approximately one order of magnitude
slower than the Nicer method (at least on a typical work-
station), and this might prevent large applications of the
method proposed here. On the other hand, the fast tech-
nological progress in the computer speed justifies the work
presented in this paper, in the sense that soon it will be
possible to use the maximum-likelihood method on large
datasets composed of millions of stars.
Another possible issue related to the technique pre-
sented here is the need for a more detailed knowledge of
the properties of the data used. As a comparison, we note
that the original Nice technique makes use only of the
H and K magnitudes of stars and of the average color
hH − Ki of stars in the control field. In addition to these
data, Nicer also requires the estimated errors of star mag-
nitudes in the NIR bands used. Finally, the maximum-
likelihood method requires a detailed knowledge of the
probability distribution pM (M ), of the measurement er-
rors pǫλ( mλmλ) of each star, and of the completeness
M. Lombardi: Optimal column density measurements from multiband near-infrared observations
17
functions cλ( mλ). In reality, in the Bayesian approach im-
plicitly adopted in this paper, the maximum-likelihood
method can also be used if an approximate knowledge of
these parameters is available. Suppose, for example, that
the parameters P , Q, and R that characterize pM (M ) [see
Eq. (35)] are only approximately known from the mea-
surements in the control field. In this case we can take
these parameters as unknowns in the expression for the
likelihood, and we can thus minimize L with respect to
them as well as with respect to the parameters that char-
acterize pAV (AV ) (AV and f in the case considered here).
As customary in standard maximum-likelihood problems,
we include the knowledge on P , Q, and R as a prior in
the function L. Note that this way the dimension of the
space over which we need to minimize L greatly increases
(and this can pose severe computational problems), but in
principle the schema proposed is applicable to real cases.
Interestingly, the Nice and Nicer technique can be
seen as special cases of the maximum-likelihood method
when no prior knowledge on R is available: the complete
lack information on the normalization of pM (M ) forces
our method to use the only color information of the stars.
Similarly, the number counts method can be recovered as
special case when the knowledge on the average colors χa
is absent [or, equivalently, when the scatter in the colors
Cab is very large; cf. Eq. (34)]. This suggests that the
Nice(r)) techniques are not more robust of the maximum-
likelihood one, but just simpler.
Finally, we mention that the presence of young stellar
objects (that could be for example embedded in the dense
cores) can in principle affect the extinction measurements.
Hence, these objects should be "manually" removed before
using any extinction measurement method, including the
maximum-likelihood described in this paper.
5. Discussion and conclusions
In this paper we have considered the problem of an accu-
rate determination of the extinction toward a dark molec-
ular cloud. The results obtained here can be summarized
in the following items:
-- The extinction and the reddening of background stars
have been considered from a general statistical point
of view.
-- The bias and uncertainties of the two main NIR tech-
niques used to map the extinction, the star count and
the color excess methods, have been discussed in de-
tail. We have shown that, although the color excess
method has generally a smaller error, it is affected by
a large bias in presence of contamination by foreground
stars. We have also shown that both Nice and Nicer
are affected by a relatively small bias (approximately
0.2mag in AV) as a result of selection effects.
-- A new optimal maximum-likelihood method has been
presented and tested with extensive simulations.
The simulations described in Sect. 4.3 have shown that
in the simple case of a thin cloud the maximum-likelihood
estimator performs significantly better than the Nice and
Nicer estimators (since the number count method is
known to have a larger noise, we did not report a de-
tailed comparison with this method here). However, the
maximum-likelihood approach also allows us to consider
more general cloud configurations, which cannot be easily
dealt with using standard techniques. For example, the
maximum-likelihood techniques could be used to measure
directly on the same patch of the sky both the column
density AV and the fraction of foreground stars f ; alter-
natively, it would be also possible to determine f globally
on the cloud and take it as a constant on the whole field
(a good approximation for small clouds).
In Sect. 4.4.1 we discussed one of the most serious lim-
itations of NIR color excess studies in molecular clouds,
namely the bias introduced by substructures. In our orig-
inal formulation, the maximum-likelihood method can
be used not only in the thin cloud approximation, but
in general for any functional form of pAV (AV ). Hence,
as mentioned above, we could implement the maximum-
likelihood method using a more realistic probability dis-
tribution for AV , such as the one of Eq. (65).
Another possibility offered by the maximum-likelihood
approach is the generalization of the thin-cloud approxi-
mation to a multi-layer case, where two or more (thin)
clouds located at different distances are observed on over-
lapping areas of the sky. For example, in case of a double
cloud we could write [cf. Eq. (17)]
pAV (AV ) = f (1)δ(AV ) + f (2)δ(AV − A(1)
V )
+(cid:0)1 − f (1) − f (2)(cid:1)δ(cid:0)AV − A(1)
V − A(2)
V (cid:1) .
(66)
This configuration, for example, might be appropriate to
study clouds close to the galactic center, where the su-
perposition of different complexes is very likely. Such a
method could effectively disentangle the effects of the two
clouds provided the values of f (1) and f (2) are sufficiently
different. Alternatively, one could use a double cloud as a
null test, i.e. to check that indeed the thin-cloud approx-
imation is appropriate (in this case one expects to find
A(1)
V ≃ 0, A(2)
V ≃ 0, or f (1) ≃ 0). Some of these possibili-
ties will be investigated in detail in a follow-up paper by
using 2MASS data.
Acknowledgements. We are very grateful to G. Bertin and
J. Alves for useful and stimulating discussions, and to the ref-
eree, L. Cambresy for providing stimulating comments and for
helping us improve the paper significantly.
Appendix A: The median and related estimators
The goal of this appendix is to derive the probability distribu-
tion of the median of n independent identical random variables.
This analysis is useful to address some of the issues discussed
in Sect. 3.2.
18
M. Lombardi: Optimal column density measurements from multiband near-infrared observations
A.1. Notation
In the following we will often deal with ordered and unordered
n-tuples. The latter can be taken to be a set of n elements,
where typically each element is a random variable or an element
of another tuple; the former can be taken to be a list of n
elements, where thus each element is associated to a position
in the list. Hence, given a positive integer k ≤ n, it makes sense
to talk about the k-th element of an ordered n-tuple, while this
is meaningless for an unordered tuple.
We will denote an ordered n-tuple using brackets, as in
[x1, . . . , xn]; instead, we will reserve braces for the unordered
tuples, as in {x1, . . . , xn}. Note that, for consistency with the
definition, we identify unordered tuples if these differ just by a
permutation of the elements: thus {x1, x2, x3} is, for instance,
identical to {x2, x1, x3}.
A.2. P n
k and the median estimator
Let us call p(x) the probability distribution for the real ran-
dom variable x, and let us consider the generation of unordered
n-tuples {x1, x2, . . . , xn} of such variables. Suppose that, af-
ter the random generation of a tuple, we order the tuple such
that x1 ≤ x2 ≤ · · · ≤ xn. We consider now the probability
distribution pn
k (xk) for the k-th element of the ordered tuple
[x1, x2, . . . , xn], which can be written as
pn
k (xk) = n n − 1
k − 1!p(xk)(cid:2)P (xk)(cid:3)k−1(cid:2)1 − P (xk)(cid:3)n−k , (A.1)
where P (x) is the cumulative probability distribution of x:
P (x) =Z x
−∞
p(x′) dx′ .
(A.2)
The expression appearing in Eq. (A.1) can be explained as
follows. Let us consider an element (for example the first) of
the unordered n-tuple. The probability that this element is the
k-th in the ordered tuple and that it has a value in the range
[x, x + dx] is the product of three terms:
-- p(x) dx to takes into account the intrinsic probability dis-
tribution of x;
-- (cid:0)n−1
k−1(cid:1)(cid:2)P (xk)(cid:3)k−1(cid:2)1−P (xk)(cid:3)n−k, which is a simple binomial
distribution giving the probability that the value chosen
has k − 1 elements of the order n-tuple at its left, and n − k
elements at its right.
-- Finally, we have to multiply the whole result by n in order
to take into account the arbitrary choice of the k-th element
in the n-tuple.
The cumulative distribution P n
k (xk) is given by
1
0.8
0.6
0.4
0.2
0
)
k
x
(
nk
P
1
2
3
4
5
6
7
0
0.2
0.4
0.6
0.8
1
P (xk)
Fig. A.1. The cumulative probability distribution P n
k (xk)
as a function of P (x) for n = 7 and various values of k
(marked close to the relative curves).
1
0.8
0.6
0.4
0.2
0
)
k
x
(
nk
P
0
0.2
0.4
0.6
0.8
1
P (xk)
Fig. A.2. The cumulative probability distribution P n
k (xk)
as a function of P (x) for k ∈ {1, 2, 3, 4, 5, 6, 7} and
n = 2k − 1. Note that all curves pass through the point
(0.5, 0.5).
In the last step we changed the index variable into m = k + ℓ.
The final result obtained in Eq. (A.3) has the advantage to be
a simple (polynomial) expression in P (x).
Figures A.1 and A.2 show the polynomials P n
k (xk) as a
function of P (xk) for various values of n and k. These figures
suggest a number of properties for P n
k (xk) that can be verified
analytically with the help of the equations written above:
-- P n
k (x) = 0 if P (x) = 0, and P n
k (x) = 1 if P (x) = 1;
k is a monotonic increasing function of P . This
moreover, P n
implies that pn
k (x) vanishes where p(x) vanishes.
P n
k (xk) = Z xk
−∞
pn
k (x′
k) dx′
k
-- Using Eq. (A.1), it is possible to show that [see below
Eq. (A.10)]
=
=
n!
(k − 1)!(n − k)!
ℓ !
(−1)ℓ n − k
n−k
Xℓ=0
k) dx′
k
×Z xk
Xm=k
n
−∞(cid:2)P (x′
(−1)m+k(cid:18) n
k)(cid:3)k−1+ℓp(x′
m(cid:19)(cid:18)m − 1
k − 1(cid:19)(cid:2)P (xk)(cid:3)m .
1
n
n
Xk=1
P n
k (x) = P (x) ,
1
n
n
Xk=1
pn
k (x) = p(x) .
(A.4)
-- As suggested by Fig. A.1 and by Eq. (A.1), we have
P n
k (x) = 1 − P n
n+1−k(x′) ,
(A.5)
(A.3)
provided that P (x) = 1−P (x′). This, in particular, implies
that P 2k−1
(x) = 1/2 if P (x) = 1/2 (see Fig. A.2).
k
M. Lombardi: Optimal column density measurements from multiband near-infrared observations
19
-- Equation (A.3) specialized to the cases k = 1 and k = n is
distributions. Let us then evaluate the sum
P n
1 (x1) =(cid:2)P (x1)(cid:3)n , P n
n (xn) = 1 −(cid:2)1 − P (xn)(cid:3)n .
(A.6)
The two last properties have a close relationship with the me-
dian estimator. We note, in fact, that for odd n the median
estimator is in our notation x(n+1)/2; as a result, the probabil-
ity distribution p2k−1
(xk) is just the probability distribution of
the median for n = 2k − 1. Hence, the property (A.5) written
above can be rephrased as
k
P n
≥k(x) =
=
1
1
n
n
P n
k (x)
Xk′=k
m(cid:19)(cid:2)P (x)(cid:3)m
Xm=k(cid:18) n
k′ − 1(cid:19) .
(−1)k+m(cid:18)m − 1
n − k + 1
n − k + 1
×
m
Xk′=k
(A.9)
The median of the probability distribution of the me-
dian estimator of the random variable x is the median
of the probability distribution of x.
We now consider two cases. If k = 1, then the last sum can
be taken to be the binomial expansion of (1 − 1)m−1, which
vanishes for all integers m > 1. Hence we are left just with the
case m = 1, for which we obtain
Note that here we use the term "median" to denote both the
usual median of an n-tuple {x1, . . . , xn}, and the median of a
distribution, defined as the value x such that the value cumu-
lative distribution is 1/2.
A.3. P n
≤k and P n
≥k
For our purposes, it is also useful to define and study two prob-
ability distributions closely related to P n
k (xk). Let us consider
again the ordered n-tuple [x1, . . . , xn], where each element is
drawn from a probability distribution p(x). Now let us retain
only the first k elements of the ordered n-tuple, and let us call
pn
≤k(x≤k) the probability distribution for each element of the
unordered k-tuple {x1, . . . , xk}; similarly, we call pn
≥k(x≥k) the
probability distribution for the tuple {xk, . . . , xn} where only
the last (n − k + 1) elements of the original ordered n-tuple are
retained.
The evaluation of the probability distribution pn
≥k(x) can
be broken into two parts. Consider the (k − 1)-th element xk−1
of the tuple [x1, . . . , xn] (i.e., the last element discarded); then,
clearly, each element of the (n − k + 1)-tuple {xk, . . . , xn} can-
not be smaller than xk−1. Moreover, each element of this un-
ordered tuple is distributed between xk−1 and ∞ according to
the (truncated) original probability distribution p(x). Finally,
by repeating this argument for each possible value of xk−1
(weighted by pn
k−1), we obtain the expression
pn
≥k(x) =Z x
−∞
pn
k−1(xk−1)
p(x)
1 − P (xk−1)
dxk−1 .
(A.7)
Note that the term 1 − P (xk−1) is introduced here to correctly
normalize the truncated probability distribution p(x).
As usual in this appendix, it is more convenient to consider
the cumulative probability distributions. We can thus integrate
pn
≥k(x) over x and obtain, after a few manipulations, a closed
expression for P n
≥k(x). Here, however, we prefer to follow a
different and simpler path.
Let us consider again the tuple [xk, . . . , xn]. Each element
of this ordered tuple follows the probability distribution pn
k (xk)
discussed in the previous section. Hence, if we consider the un-
ordered tuple, the elements {xk, . . . , xn} will follow the average
distribution
pn
≥k(x) =
1
n − k + 1
pn
k (x) .
n
Xk′=k
(A.8)
By integrating both sides of this equation on dx, we can verify
that the same relation holds for the cumulative probability
P n
≥1 =
1
n
n
Xk′=k
P n
k (x) = P (x) .
(A.10)
This shows that each element of the unordered n-tuple
{x1, . . . , xn} follows the original distribution p(x), a very nat-
ural result indeed.
If, instead, k > 1, then using the properties of the binomial
coefficient we find
P n
≥k(x) =
1
n − k + 1
n
Xm=k
(−1)m+k(cid:18) n
m(cid:19)(cid:18)m − 2
k − 2(cid:19)(cid:2)P (x)(cid:3)m .
(A.11)
The similarity of this result with the last line of Eq. (A.3) is
rather surprising.
Similarly, we wish to investigate the cumulative probability
≤k(x) for the k-tuple {x1, . . . , xk}. This quantity
distribution P n
is better evaluated using P n
≥k(x):
P n
≤k(x) =
=
=
=
k
1
k
n
1
P n
P n
k′ (x)
k′ (x) −
Xk′=1
k(cid:20) n
Xk′=k+1
Xk′=1
k(cid:20)nP (x) − (n − k)P n
Xm=k+1(cid:18) n
P (x) +
n
k
1
k
1
n
P n
k′ (x)(cid:21)
≥k+1(x)(cid:21)
m(cid:19)(cid:18)m − 2
×(cid:2)P (x)(cid:3)m .
k − 1(cid:19) (−1)m+k
(A.12)
As an application of the results obtained in this section,
we evaluate the bias introduced in the estimate of the column
density when using the technique described in Sect. 3.2 to re-
move foreground stars [see Eq. (54)]. Suppose that we observe
stars in a region with no significant extinction, so that both
foreground and background stars have the same distribution
in colors. For simplicity, we assume that the reddening esti-
mates AV for each individual star is a Gaussian distribution
with vanishing average and variance σ AV
. In this situation, if
we exclude the k bluer stars (because they are taken to be
foreground), we will bias the estimate of AV toward large col-
umn densities. In particular, the column densities for the stars
left will be distributed according to pn
≥k+1. The median of this
20
M. Lombardi: Optimal column density measurements from multiband near-infrared observations
distribution can be evaluated as follows. At AV = 0 we have
P (AV ) = 1/2, while
Kleinmann, S. G., Lysaght, M. G., Pughe, W. L., et al. 1994,
Experimental Astronomy, 3, 65
P n
≥k+1(0) =
1
n − k(cid:20) n
2
−
k
Xk′=1
P n
k′ (1/2)(cid:21) .
(A.13)
Assuming n ≫ k, then each term in the sum above is approx-
imately unity (cf. above Fig. A.1). Hence we obtain
P n
≥k+1(0) ≃
n − 2k
2n − 2k
.
(A.14)
The fact that this quantity is not exactly 1/2 is telling us that
there is a bias. The exact amount of this bias can be evalu-
ated by solving the equation P n
≥k+1(x) = 1/2. Using Newton's
method we obtain then the approximate solution
Lada, C. J., Alves, J., & Lada, E. A. 1999, ApJ, 512, 250
Lada, C. J., Lada, E. A., Clemens, D. P., & Bally, J. 1994,
ApJ, 429, 694
Larson, R. B. 1981, MNRAS, 194, 809
Li, A. & Greenberg, J. M. 2003, in Solid State Astrochemistry,
37, astro-ph/0204392
Lilley, A. E. 1955, ApJ, 121, 559
Lombardi, M. 2005, submitted to A&A
Lombardi, M. & Alves, J. 2001, A&A, 377, 1023
Mathis, J. S. 1990, ARA&A, 28, 37
Mattila, K. 1986, A&A, 160, 157
Miesch, M. S. & Bally, J. 1994, ApJ, 429, 645
Padoan, P., Jones, B. J. T., & Nordlund, A. P. 1997, ApJ, 474,
730
416, 157
Schultheis, M., Ganesh, S., Simon, G., et al. 1999, A&A, 349,
L69
Trumpler, R. J. 1930, Lick Observatory Bulletin, 14, 154
van de Hulst, H. C. 1946, Recherches Astronomiques de
l'Observatoire d'Utrecht, 11, 2
Wolf, M. 1923, Astronomische Nachrichten, 219, 109
x ≃(cid:20) dP n
≥k+1
dx
(cid:21)−1
x=0(cid:20) 1
2
− P n
≥k+1(0)(cid:21) ≃
k
2np(0)
.
(A.15)
Rieke, G. H. & Lebofsky, M. J. 1985, ApJ, 288, 618
Robin, A. C., Reyl´e, C., Derri`ere, S., & Picaud, S. 2004, A&A,
Note that in this equation we used the relationship between
cumulative and differential probability distributions; moreover,
we simplified pn
k (0) ≃ 0 at low k, and retained only terms linear
in k/n. The final result obtained is thus given by Eq. (54).
References
Alves, J., Lada, C. J., & Lada, E. A. 1999, ApJ, 515, 265
Alves, J., Lada, C. J., & Lada, E. A. 2001, Nature, 409, 159
Andreazza, C. M. & Vilas-Boas, J. W. S. 1996, A&AS, 116, 21
Barnard, E. E. 1919, ApJ, 49, 1
Bohlin, R. C., Savage, B. D., & Drake, J. F. 1978, ApJ, 224,
132
Bok, B. J. 1937, The distribution of the stars in space (Chicago:
University of Chicago Press, 1937)
Bok, B. J. 1956, AJ, 61, 309
Cambr´esy, L., Beichman, C. A., Jarrett, T. H., & Cutri, R. M.
2002, AJ, 123, 2559
Cambr´esy, L. 1999, A&A, 345, 965
Cambresy, L., Epchtein, N., Copet, E., et al. 1997, A&A, 324,
L5
Casali, M. M. 1986, MNRAS, 223, 341
Cressie, N. A. C. 1993, Statistics for Spatial Data (Wiley &
sons, New York)
Dickman, R. L. 1978, AJ, 83, 363
Eadie, W., Drijard, D., James, F., Roos, M., & Sadoulet,
B. 1971, Statistical Methods in Experimental Physics
(Amsterdam New-York Oxford: North-Holland Publishing
Company)
Epchtein, N., de Batz, B., Capoani, L., et al. 1997, The
Messenger, 87, 27
Frerking, M. A., Langer, W. D., & Wilson, R. W. 1982, ApJ,
262, 590
Harjunpaa, P., Lehtinen, K., & Haikala, L. K. 2004, A&A, 421,
1087
Heyer, M. H. & Brunt, C. M. 2004, ApJ, 615, L45
Horner, D. J., Lada, E. A., & Lada, C. J. 1997, AJ, 113, 1788
Indebetouw, R., Mathis, J. S., Babler, B. L., et al. 2004, ArXiv
Astrophysics e-prints
Jarrett, T. H., Dickman, R. L., & Herbst, W. 1994, ApJ, 424,
852
Jones, T. J., Hyland, A. R., & Bailey, J. 1984, ApJ, 282, 675
Jones, T. J., Hyland, A. R., Robinson, G., Smith, R., &
Thomas, J. 1980, ApJ, 242, 132
|
0810.4793 | 1 | 0810 | 2008-10-27T11:48:54 | On the Low Surface Magnetic Field Structure of Quark Stars | [
"astro-ph",
"cond-mat.supr-con"
] | Following some of the recent articles on hole super-conductivity and related phenomena by Hirsch \cite{H1,H2,H3}, a simple model is proposed to explain the observed low surface magnetic field of the expected quark stars. It is argued that the diamagnetic moments of the electrons circulating in the electro-sphere induce a magnetic field, which forces the existing quark star magnetic flux density to become dilute. We have also analysed the instability of normal-superconducting interface due to excess accumulation of magnetic flux lines, assuming an extremely slow growth of superconducting phase through a first order bubble nucleation type transition. | astro-ph | astro-ph | On the Low Surface Magnetic Field Structure of Quark Stars
Nandini Nag, Sutapa Ghosh, Roni Saha and Somenath Chakrabarty†
Department of Physics, Visva-Bharati, Santiniketan 731 235
West Bengal, India, ‡E-mail:[email protected]
Abstract
Following some of the recent articles on hole super-conductivity and related phenomena by
Hirsch [1, 2, 3], a simple model is proposed to explain the observed low surface magnetic field of
the expected quark stars. It is argued that the diamagnetic moments of the electrons circulating in
the electro-sphere induce a magnetic field, which forces the existing quark star magnetic flux density
to become dilute. We have also analysed the instability of normal-superconducting interface due to
excess accumulation of magnetic flux lines, assuming an extremely slow growth of superconducting
phase through a first order bubble nucleation type transition.
PACS numbers: 97.60.Jd, 97.60.-s, 75.25.+z
8
0
0
2
t
c
O
7
2
]
h
p
-
o
r
t
s
a
[
1
v
3
9
7
4
.
0
1
8
0
:
v
i
X
r
a
1
1.
INTRODUCTION
More than two decades ago, Witten had conjectured in an outstanding article [4] (see also
[5]) that a flavor symmetric quark matter at zero temperature and at zero pressure would be
energetically the most stable configuration. This exotic phase is known as the strange quark
matter (SQM). Since then a large number of articles have been published on the theoretical
studies of the properties of bulk SQM as well as droplets (known as strange-lets) of SQM
[6, 7, 8].
In those investigations various kinds of confinement models were used.
It has
further been shown that SQM phase can exist in compact stellar objects, known as strange
quark stars or strange stars [9, 10]. The finite size strange-lets are expected to be formed
both in ultra-relativistic heavy ion collisions and also during QCD phase transition in the
early universe [11, 12]. The strange-lets expected to have formed in the early universe, are
also called the strange nuggets, believed to be the relics from primordial quark-hadron phase
transition. These relics are also one of the viable candidates for baryonic dark matter. Inside
the strange stars, three light flavors: up (u), down (d), strange (s), along with small quantity
of electrons are in β-equilibrium. The presence of electrons make the system electrically
charge neutral. So far the gross properties of such compact objects are concerned, it is
almost impossible to distinguish strange stars from the conventional neutron stars. Except,
it is believed that these compact objects, which are expected to be strange stars, are fast
rotating (milli second or sub-milli second pulsars) and the surface magnetic field is quite
low ( ≤ 108G) [13]. It is further assumed that strange stars are very old and also extremely
cold objects, formed by the accretion of matter from the companion binary counter part.
However, for a strange star, the compositional structure in the microscopic scale is extremely
complex in nature. Since the density at the core region is very high (3 − 4 times normal
nuclear density), the corresponding quark matter is almost flavour symmetric. As one goes
radially outward, the density of matter decreases and since s-quarks are much more heavier
than u and d-quarks (current masses for u and d-quarks are 5−10MeV, whereas, for s-quark,
it is ≈ 150MeV), in this region, it is therefore energetically not favorable to produce enough
s-quarks through weak processes. The abundance of s-quark is therefore steadily decreases
as one moves towards the surface region. The charge neutrality of quark matter near the
outer core or the crustal region is therefore mainly maintained by electrons instead of strange
quarks. The density of electron is negligibly small at the core and is maximum at the crustal
2
region. Of course it is not very high (≤ 10−4 times normal nuclear density). These electrons
are bound to the positively charged quark matter by electromagnetic force (of which the
Coulomb force is the non-relativistic scenario) and are allowed to move freely across the
strong Coulomb field. Because of this strong electromagnetic force, they can not move far
away from the quark matter surface. The electron gas at the surface of strange stars, form a
very thin layer of width a few thousand Fermi, known as electro-sphere. Therefore, the gross
compositional structure of a strange star is very simple: a positively charged quark matter
region covered by a thin electro-sphere. Since the strange quark matter is energetically
more favorable, in the case of strange stars, it is unlikely to have a crust of dense iron like
matter, which is expected to be present in neutron stars. Further, depending on the internal
temperature and density of strange star, the quark matter may be in normal phase or in
super-conducting state or in color flavor locked phase (CFL) [14]. The CFL phase at the
core region is expected if the density is extremely high. Since the kinetic energies of the
electrons are much larger than the corresponding super-conducting gap energy, they never
show super-conductivity, either inside the star or at the electro-sphere. Which actually
means that the critical temperature for superconducting transition of electron is much less
than the corresponding critical temperature for quark matter.
In this article we shall propose a mechanism by which very low surface magnetic field of
strange stars can be obtained. We have assumed a type-I superconducting transition and
consider two possible flux expulsion processes. We shall show that extremely slow expulsion
mechanism leads to an instability at the normal-superconductor interface of quark matter.
In this article we have also analysed the interface instability. It is further noticed that by a
slow expulsion process, at several points on the interface, since the strength of magnetic field
becomes > Bc, the critical strength to destroy type-I superconductivity of quark matter,
the growth of superconducting quark matter phase will stop abruptly. On the other hand,
if it is an extremely fast process, we ultimately get a stable quark star of very low surface
magnetic field.
The paper is organized in the following manner: In the next section we shall compare a
strange star with a giant atom. In this section we shall discuss qualitatively the physical
processes which may take place at the surface of strange star and compare them with a giant
ultra-heavy atom. In section 3. we shall make a comparative study of superconducting quark
matter in strange star and the theory of hole superconductivity. The comparison is again
3
purely qualitative. The mathematical formalism will be given in section 6. We shall discuss
the mechanism of slow flux expulsion and give a mathematical formalism in sections 5 and
6 respectively. In the last section we have given the conclusion of the present work.
2. A GIANT ATOM MODEL
In this section we shall try to explain the possible low magnetic surface structure of quark
stars. In this model calculation, we assume a prompt phase transition to superconducting
phase and the strange star is treated as an equivalent giant atom with positively charged
quark matter (along with some admixture of electrons) as the nucleus and the electrons
in the electro-sphere are treated as orbital electrons. Unlike normal atoms, here size of
the nucleus is ≫ than the volume occupied by the electrons in the electro-sphere. Now
it is very easy to show from the β-equilibrium and charge neutrality conditions that the
net positive charge content of the quark matter nucleus is ≫ 172, the critical value of Z,
the atomic number of a typical super-heavy Dirac-atom. At the quark matter surface, the
vacuum will therefore become unstable, and spontaneous e+e−-pairs will be created if there
are unoccupied energy shell [15] (see also [16]).
Now for a many body fermion system, the microscopic theory of superconductivity sug-
gests that if the interaction favors the formation of pairs at low temperature, the system
may undergo a phase transition to a super-fluid state. In astrophysics, this is expected to
occur in the dense neutron matter of cold neutron stars. Whereas the small percentage of
protons (∼ 4%) inside neutron stars undergo a transition to type-II superconducting phase
[17, 18]. In the case of a many body system of fermions the well known BCS theory is gen-
erally used to study the super-conducting properties due to fermion pairing. One fermion
of momentum ~p and spin ~s combines with another one of momentum −~p and spin −~s and
form a Cooper pair. In the case of type-I electronic super-conductors the coupling is me-
diated by the electron-phonon interaction. In the case of quark matter, however, the basic
quark-quark interaction at large distance favors the formation of Cooper pairs. In the case
of quark matter, since the force is mediated via gluons, it gives rise to what is known as color
superconductivity. For a highly degenerate fermion system, which is true for strange star
matter, the pairing takes place near the Fermi surface. The other important condition that
must be satisfied for the formation of Cooper pairs is that the temperature (T ) of the system
4
should be much less than the super-conducting energy gap (∆). In the case of strange stars,
since the Fermi levels are not identical for u, d and s-quarks, only same type of quarks can
form Cooper pairs at the Fermi surface and give rise to color super-conductivity [19]. It is
interesting to note that the electrons, whose density is extremely low compared to quark
matter, may also be treated as highly degenerate relativistic plasma, but are unlikely to
form Cooper pairs.
In the next section we have assumed a type-I super-conducting phase of quark matter
within the strange stars and develop a mechanism by which the expelled magnetic field
is reduced at the electro-sphere. Since we are not investigating any of the properties of
super-conducting quark matter, rather, we are interested to study some of the important
features of normal electron gas layer at the surface region / inside the electro-sphere, which
are essential to have low surface magnetic field, then instead of standard relativistic version
of BCS theory [19], here, we have followed the interesting idea of "theory of hole super-
conductivity", proposed by Hirsch in a series of articles [1, 2, 3, 20, 21]. In the case of strange
stars the charge separation or the charge asymmetry takes place by the combined effect of
charge neutrality and β-equilibrium conditions; completely different physical mechanism.
Further, the properties of both positively charged quark matter inside the star and the
electrons in the electro-sphere are just opposite to the predicted nature of hole states and
the electrons for the hole super-conductor. In fact, we have noticed that the theory of hole
super-conductivity is not applicable for super-conducting quark matter of strange stars. The
reason is the strong interaction which bind the quarks within the strange stars.
3. THE HOLE SUPERCONDUCTIVITY AND COMPACT STRANGE STARS
To develop a mechanism of strange star magnetic field suppression at the electro-sphere,
we consider the charge asymmetric strange star as an equivalent hole super-conductor. We
have noticed, that there are a lot of similarities and also dissimilarities between the labo-
ratory samples and the strange stars within the theory of hole super-conductivity. In the
case of laboratory superconducting samples, according to the theory proposed by Hirsch, the
inner part is positively charged holes in normal phase. The electrons at the surface region
are super-conducting. The surface layer is bound to the inner region by strong electromag-
netic force. In the case of strange stars, the inner region is positively charged quark matter.
5
But unlike the laboratory superconductors, it is in super-conducting phase. Whereas, the
outer layer, the electro-sphere, which is a degenerate electron gas, is in normal phase. The
force between these two regions is again electromagnetic in nature. Going a step further,
we assume following the model by Hirsch, that the electrons in the electro-sphere gyrate
about the magnetic lines of force, expelled from the super-conducting interior of the strange
star. The origin of such orbital motion of these normal electrons in electro-sphere about the
magnetic flux lines is the well known Lorentz force. This orbital motion of electrons will
generate a magnetic field at the surface of the strange star. According to Lenz's law, the
induced field must be in the direction opposite to that of existing one. At the surface or
in the electro-sphere, therefore, the magnetic field from the tiny magnets produced by the
gyration of electrons about the existing lines of forces reduce the strange star magnetic field
by diamagnetic effect. Therefore, unlike the hole super-conductor, in the case of strange
stars, the magnetic field expelled from the super-conducting quark matter are suppressed
by the electron gas, which is in normal phase. By the repulsive diamagnetic action, there
will be an effective dilution of magnetic flux lines inside the electro-sphere. The dilution of
magnetic flux will naturally increase the size (width) of the electro-sphere. Analogous to
the phenomenon of frozen-in magnetic field, here, the "frozen-in electron gas" will be pulled
away by the flux lines. The increase in size of the electro-sphere will depend both on the
repulsive magnetic force and the attractive coulomb force and in equilibrium configuration,
these two will balance each other. The width of the electro-sphere of a typical quark star of
surface magnetic field ≤ 108G can be a few tens of km, which is at least an order of mag-
nitude less than the radius of the light cylinder, ∼ 100km, for a milli second pulsar. Now
this increase in size of the electro-sphere will reduce the electron density within the system.
The relation between the induced field and the existing field is given by eqn.(39). Further,
because of the motion of electrons around the quark matter nucleus, the actual trajectory of
an electron in the electro-sphere is more or less like a closed helical spring produced by the
motion of circular orbit along the lines of force. Since the magnetic field strengths at the
poles are very high, the electrons in the helical trajectory will be reflected back from these
region. The two ends will therefore behave like magnetic mirrors. It is also obvious from
eqn.(39) that the radius of gyration is larger at the equatorial region compared to the polar
values. Since the diamagnetism of the electrons inside the electro-sphere is caused by Lenz's
law, the sign of ~L.~p/~L.~p, which may be called as the effective "orbital-helicity" will change
6
sign after each reflection by the magnetic mirrors at the poles. Here ~p and ~L are respectively
the linear momentum and the angular momentum of the electrons. Further, when a balance
between the existing field and the induced field will be established, the electrons within each
half of the electro-sphere will under go steady helical motion.
4. SLOW EXPULSION OF MAGNETIC FLUX LINES
In this section we shall give a qualitative picture of very slow magnetic flux expulsion from
the super conducting region. It has been shown in ref.[19] in a relativistic version of BCS
theory, that if a normal quark matter system undergoes a superconducting phase transition,
the newly produced quark matter phase will be a type-I superconductor. They have also
shown in that the critical magnetic field to destroy such pairing is ∼ 1016G for n ∼ 3 − 4n0,
with n0 = 0.17fm−3, the normal nuclear density. This magnetic field strength is indeed much
larger than the typical pulsar magnetic field. The corresponding critical temperature is ∼
109−1010K, which is again high enough for strange stars, which are expected to be extremely
cold objects.
In this section, instead of investigating the superconducting properties of
quark matter inside strange stars, we shall give a possible mechanism of flux expulsion
by Meissner effect, assuming that the growth of superconducting phase is extremely slow.
Further, we assume that the magnetic field strength at the core region of a strange star
are much less than the corresponding critical value for the destruction of superconducting
property and the temperature is also low enough. Then during such a type-I superconducting
phase transition, the magnetic flux lines from the superconducting quark matter of the
strange star will be pushed out towards the normal crustal region. Now for a small type-
I superconducting laboratory sample placed in an external magnetic field less than the
corresponding critical value, the expulsion of magnetic field takes place instantaneously.
Whereas in the bulk strange star scenario, the picture may be completely different.
It
may take several thousands of years for the magnetic flux lines to get expelled from the
superconducting core. Which further means, that the growth of superconducting phase in
strange stars may not be instantaneous. A simple estimate shows that the expulsion time due
to ohmic diffusion is ∼ 104yrs [22]. It was shown by Chau using Ginzberg-Landau formalism
that the time for expulsion of magnetic lines of force accompanied by the enhancement of
magnetic field non-uniformities at the crustal region gets prolonged to 107yrs [23]. Alford
7
et. al. investigated the expulsion of magnetic lines of force from the colour superconducting
region by considering the pairing of like and unlike quarks and obtained the expulsion time
much larger than the age of the Universe [24].
In our investigation of magnetic flux expulsion from growing superconducting core of a
strange star, the idea of impurity diffusion in molten alloys or the transport of baryon num-
bers from hot quark matter soup to hadronic matter during quark-hadron phase transition
in the early universe, expected to occur micro-second after big bang (the first mechanism is
used by the material scientists and metallurgists [25], whereas the later one is used by cosmol-
ogists working in the field of big bang nucleo-synthesis [26, 27]) are assumed. In the present
section, we shall further show the possibility of Mullins-Sekerka normal-superconducting
interface instability [28, 29] in quark matter. This is generally observed in the case of solidi-
fication of pure molten metals at the solid-liquid interface, if there is a temperature gradient.
The interface will always be stable if the temperature gradient is positive, otherwise it will
be unstable. In alloys, the criteria for stable / unstable behaviour is more complicated. It is
seen that, during solidification of an alloy, there is a substantial change in the concentration
ahead of the interface. Here solute diffusion as well as the heat flow effects must be consid-
ered simultaneously. The particular problem we are going to investigate here is analogous
to solute diffusion during solidification of an alloy.
5. A FORMALISM FOR SLOW NUCLEATION
It has been assumed that the growth of superconducting quark bubble started from the
centre of the star and the nomenclature controlled growth for such phenomenon has been
used.
If the magnetic field strength and the temperature of the star are a few orders of
magnitude less than their critical values, the normal quark matter phase is thermodynami-
cally unstable relative to the corresponding superconducting one. Then due to fluctuation,
a droplet of superconducting quark matter bubble may be nucleated in metastable nor-
mal quark matter medium. If the size of this superconducting bubble is greater than the
corresponding critical value, it will act as the nucleating centre for the growth of supercon-
ducting quark core. The critical radius can be obtained by minimising the free energy. Then
8
following the work of Mullins and Sekerka, we have [28, 29]
,
(1)
rc =
16πα
B(c)2(cid:20)1 −(cid:16) B
B(c)(cid:17)2(cid:21)
where α is the surface tension or the surface energy per unit area of the critical supercon-
ducting bubbles, (from this expression it is possible to obtain the critical size of the quark
matter bubble by considering 10−3 ≤ α ≤ 1 (in MeV/fm−3) as the range of surface tension)
which is greater than zero for a type-I superconductor-normal interface, B(c) is the critical
magnetic field. In presence of a magnetic field B < B(c), the normal to superconducting
transition is first order in nature. As the superconducting phase grows continuously, the
magnetic field lines will be pushed out into the normal quark matter crust. This is the usual
Meissner effect observed in type-I superconductor. We compare this phenomenon of mag-
netic flux expulsion from a growing superconducting quark matter core with the diffusion of
impurities from the frozen phase of molten metal or the transport of baryon numbers from
hot quark matter soup during quark-hadron phase transition in the early universe. The
formation of superconducting zone is compared with the solidification of molten metal or
with the transition to hadronic phase with almost zero baryon number. It is known from
the simple thermodynamic calculations that if the free energy of molten phase decreases in
presence of impurity atoms, then during solidification they prefer to reside in the molten
phase otherwise they go to the solid phase. In this particular case the magnetic field lines
play the role of impurity atoms and because of less free energy, they prefer to remain in
normal quark matter phase. The normal quark matter phase plays the role of molten metal
or the hot quark soup. Whereas the superconducting phase can be compared with the frozen
solid phase or the hadronic phase. This idea was applied to baryon number transport dur-
ing first order quark-hadron phase transition in the early Universe, where baryon number
replaces impurity, quark phase replaces molten metal and hadronic matter replaces that of
solid metal [26, 27]. Of course the baryon number prefers to stay in the quark phase because
of Boltzmann suppression factor in the hadronic phase. Since the magnetic flux lines prefer
to reside in the normal phase, the well known Meissner effect can therefore be restated as the
solubility of magnetic flux lines in the superconducting phase is zero with a finite penetration
depth.
The dynamical equation for the flux expulsion can be obtained from the simplified model
of sharp normal-superconducting interface. The expulsion equation is given by the well
9
known diffusion equation [30]
∂B
∂t
= D∇2B
where B is the magnetic field intensity and D is the diffusion coefficient, given by
D =
c2
4πσn
,
(2)
(3)
σn is the electrical conductivity of the normal quark matter and c is the velocity of light.
The electrical conductivity of quark matter for B = 0 is given by [31, 32]
σn ∼ α−3/2
s
T −2
10
or [33]
σn ∼ (αsT10)−5/3
(4)
expressed in sec−1, where αs is the strong coupling constant and T10 = T /1010K, the nu-
merical value for this electrical conductivity in the case of quark matter relevant for quark
star density is ∼ 1026 sec−1. In the order of magnitude estimate, we have used this numer-
ical value for electrical conductivity. In this context we must mention that in presence of
strong quantizing magnetic field, σn may not be a scalar quantity, magnetic field destroys
the isotropy of the electromagnetic properties of the medium. In particular, for extremely
large B, the components of electric current vector orthogonal to B become extremely small.
Which indicates that the quarks can move only along the direction of magnetic field or in
other words, across the field the resistivity becomes extremely high. When the conditions
for charge neutrality and β-equilibrium are considered together, since the mass of s-quark
is assumed to be 150Mev, the electron density can not become exactly zero, but it is a few
orders of magnitude less than the s-quark density. Therefore, one can neglect the electron
contribution to electrical conductivity.
A solution of eqn.(2) with spherical symmetry can be obtained by Greens' function tech-
nique, and is given by (for a general topological structure, no analytical solution is possible)
B(r, t) =
B(0)(r′)
1
2r(πDt)1/2 Z ∞
0
[ exp(−u2
−) − exp(−u2
+)]r′dr′
(5)
where u± = (r ± r′)/2(Dt)1/2 and B(0)(r) is the magnetic field distribution within the star at
t = 0, which is of course an entirely unknown function of radial coordinate r. To obtain an
10
estimate of magnetic field diffusion time scale (τD), we assume B(0)(r) = B(0) = constant.
Then we have from eqn.(5)
B(r, t) = B(0)(cid:20)1 −
2
r
(πDt)1/2(cid:21)
(6)
Hence, if we put B(r, t) = 0 (field free condition), the estimated time scale for the expulsion
of magnetic flux lines is ∼ 105 − 106yrs. Which is of the same order of magnitude as
the Ohmic decay scale. From this simple estimate, this is the approximate time scale for
complete expulsion of magnetic field lines. Which is of course quite large. Unfortunately,
nothing else can be inferred about the growth of superconducting zone and the associated
expulsion of magnetic flux lines from this region. The reason behind such uncertainty is
our lack of knowledge or definite ideas on the numerical values of the parameters present in
eqn.(5).
To get some idea of the effect of magnetic field on the structure of growing superconduct-
ing zone, we shall now investigate the morphological instability of normal-superconducting
interface of quark matter. The motion of normal-superconducting interface is extremely im-
portant in this case and has to be taken into consideration. Then instead of eqn.(2) which is
valid in the rest frame, an equation expressed in a coordinate system which is moving with
an element of the boundary layer is the correct description of such superconducting growth,
known as Directional Growth. The equation is called Directional Growth Equation, and is
given by
∂B
∂t
− v
∂B
∂z
= D∇2B
(7)
where the motion of the plane interface is assumed to be along z-axis and v is the velocity of
the front. This diffusion equation must be supplemented by the boundary conditions at the
interface. The first boundary condition is obtained by combining Ampere's and Faraday's
laws at the interface, and is given by
Bv s= −D(∇B).n s
(8)
where n is the unit vector normal to the interface directed from the normal phase to the
superconducting phase. This is nothing but the continuity equation for magnetic flux diffu-
sion. For the normal growth of superconducting zone, the rate at which excess magnetic field
lines are rejected from the interior of the phase is balanced by the rate at which magnetic
flux lines diffuses ahead of the two-phase interface. Therefore the boundary layer between
11
superconducting-normal quark matter phases will become unstable if excess magnetic field
lines are present on the surface of the growing superconducting bubble, i.e., if the rate of
diffusion of magnetic flux lines is slow enough compared to the rate at which they are ex-
pelled from the superconducting zone. Local thermodynamic equilibrium at the interface
gives (Gibbs-Thompson criterion)
B s≈ B(c)(cid:18)1 −
4πα
RB(c)2(cid:19) = B(c) (1 − δC)
(9)
where δ is called capillary length with α the surface tension, C is the curvature = 1/R (for
a spherical surface), and B(c) is the thermodynamic critical field.
To investigate the stability of superconducting-normal interface, we shall follow the origi-
nal work by Mullins and Sekerka [28, 29], and consider a steady state growth of superconduct-
ing core. Then the time derivative in eqn.(7) will not appear. Introducing r⊥ = (x2 + y2)1/2
as the transverse coordinate at the interface, we have after rearranging eqn.(7)
" ∂2
∂r2
⊥
+
1
r⊥
∂
∂r⊥
+
∂2
∂z2 +
v
D
∂
∂z# B = 0
(10)
The approximation that the solidification is occurring under steady state condition used
in the freezing of molten material will be followed in the present case of normal to super-
conducting phase transition. Now if it is assumed that these two phenomena taking place
in two completely separate physical world are almost identical natural processes, then the
concentration of magnetic flux lines and normal-superconducting interface morphology will
be independent of time. The main disadvantage of this assumption is that there will be no
topological evolution of the interface shape. As a consequence of this constraint the solution
to the basic diffusion problem is indeterminate and a whole range of morphologies is per-
missible from the mathematical point of view. In order to distinguish the solution which is
the most likely to correspond to reality, it is necessary to find some additional criteria. The
examination of the stability of a slightly perturbed growth form is probably the most rea-
sonable manner in which this situation may be treated. In the following we shall investigate
the morphological instability of normal-superconducting interface from eqn.(10). Assuming
a solution of this equation expressed as the product of separate functions of r⊥ and z and
setting the separation constant equal to zero and using the boundary condition given by
eqn.(9), we have for an unperturbed boundary layer moving along z-axis
B = B(s) exp(−zv/D) = B(s) exp(−2z/l)
(11)
12
where l = 2D/v is the layer thickness, which is very small for small D. Mathematically, the
thickness of this layer is infinity. For practical purpose an effective value l can be taken.
The order of magnitude estimates or limiting values for the three quantities D, v and l can
be obtained from the stability condition of planer interface.
Due to excess magnetic flux lines at the interface, the form of the planer normal-
superconducting interface described by the equation z = 0 is assumed to be changed by
a small perturbation represented by the simple sine function
z = ǫ sin(~k.~r⊥)
(12)
where ǫ is very small amplitude and ~k is the wave vector of the perturbation. Then the
perturbed solution of the magnetic field distribution near the interface can be written as
B = B(s) exp(−vz/D) + Aǫ sin(~k.~r⊥) exp(−bz)
(13)
where A and b are two unknown constants. Since the solution should satisfy the diffusion
equation (eqn.(10)), we have
b =
v
2D
+"(cid:18) v
2D(cid:19)2
+ k2#1/2
(14)
To evaluate A, we utilise the assumption that both ǫ and ǫ sin(~k.~r⊥) are small enough so
that we can keep only the linear terms in the expansion of exponentials present in eqn.(13).
Then at the interface, we have after some straight forward algebraic manipulation
A =
v
D
B(s)
(15)
The expression describing the magnetic field distribution ahead of the slightly perturbed
interface then reduces to
B = B(s)(cid:20)exp(−vz/D) +
v
D
ǫ sin(~k.~r⊥) exp(−bz)(cid:21)
Now from the other boundary condition (eqn.(9)) we have
B(s) = B(c) −
4παB(c)
B(s)2 C
(16)
(17)
where C = z ′′/(1 + z ′2)3/2 is the curvature at z = ǫ sin(~k.~r⊥) and prime indicates derivative
with respect to r⊥.
13
Neglecting z ′2, which is small for small perturbation, we have
B(s) = B(c) + Γk2S
(18)
where Γ = 4παB(c)/B(s)2 and we have replaced ǫ sin(~k.~r⊥) by S. Since the amplitude of
perturbation ǫ is extremely small, the quantity S is also negligibly small.
Now eqn.(18) can also be written as
B(s) = B(c) + GS
where
G =
dB
dz
z=S= −
v
D (cid:18)1 −
vS
D (cid:19) B(s) − bAS(1 − bS)
Combining these two equations, we have
k2Γ +
v
D (cid:18)1 −
vS
D (cid:19) B(s) −
bv
D
B(s)S(1 − bS) = 0
(19)
(20)
(21)
This expression determines the form (values of k) which the perturbed interface must assume
in order to satisfy all of the conditions of the problem. To analyse the behaviour of the roots,
we replace right hand side of eqn.(21) by some parameter −P . (We have taken −P in order
to draw a close analogy with the method given in refs.[26, 27]). Then rearranging eqn.(21),
we have
− k2Γ −
v
D (cid:18)1 −
vS
D (cid:19) B(s) +
bvB(s)S
D
(1 − bS) = P
(22)
(in refs.[26, 27] the parameter P is related to the time derivative of ǫ, the amplitude of
small perturbation). If the parameter P is positive for any value of k, the distortion of the
interface will increase, whereas, if it is negative for all values of k, the perturbation will
disappear and the interface will be stable. In order to derive a stability criterion, it only
needs to know whether eqn.(20) has roots for positive values of k. If it has no roots, then
the interface is stable because the P − k curve never rises above the positive k-axis and P is
therefore negative for all wavelengths. We have used Decarte's theorem to check how many
positive roots are there. It is more convenient to express k in terms of b and then replacing
b by ω + v/D. Then we have from eqn.(22)
− ω2 Γ +
vB(s)S2
D ! − ω Γ +
2vB(s)S2
D
v
D
B(s)(cid:18)1 −
−
v
D
14
− B(s)S! v
S(cid:19)2
= P
D
(23)
This is a quadratic equation for ω. The first and the third terms are always negative. The
second term will also be negative if
Γ +
2vB(s)S2
D
− B(s)S > 0
(24)
Then it follows from Decart's rule that if the condition (24) is satisfied, there can not be
any positive root. Which implies that the small perturbation of the interface will disappear.
Since the amplitude of perturbation is assumed to be extremely small, the quantity S =
ǫ sin(~k.~r⊥) is also negligibly small. Under such circumstances the middle term of eqn.(23) is
much smaller than rest of the terms. The Decart's rule given by the condition (24) can be
re-written as
Γ > B(s)S
(25)
Which after some simplification gives the stability criterion for the plane unperturbed in-
terface, given by
α >
B(s)3S
4πB(c)
(26)
From the stability criterion, it follows that the normal-superconducting interface en-
ergy/area of quark matter has a lower bound, which depends on the interface magnetic field
strength, critical field strength and also on the perturbation term S. An order of magnitude
of this lower limit can be obtained by assuming B(s) = 10−3B(c). (Since the critical field
B(c) ∼ 1016G, and the neutron star magnetic field strength B ∼ 1013G, we may use this
equality). Then the lower limit is given by
αL ≈ 10−9 MeV/fm2 (cid:18) S
fm(cid:19)
On the other hand for B(s) = 0.1B(c), we have
αL ≈ 10−3 MeV/fm2 (cid:18) S
fm(cid:19)
The approximate general expression for the lower limit is given by
αL ≈ h3 MeV/fm2 (cid:18) S
fm(cid:19)
where h = B(s)/B(c). Therefore the maximum value of this lower limit is
αmax.
L ≈ 1 MeV/fm2 (cid:18) S
fm(cid:19)
15
(27)
(28)
(29)
(30)
when the two phase are in thermodynamic equilibrium. Of course for such a strong mag-
netic field, as we have seen [34, 35] that there can not be a first order quark-hadron phase
transition.
On the other hand if we do not have control on the interface energy, which can in principle
be obtained from Landau-Ginzberg model, we can re-write the stability criteria in terms of
interface concentration of magnetic field strength B(s), and is given by
1/3
B(s) <
4παB(c)
S(cid:16)1 − 2v
D S(cid:17)
(31)
This is more realistic than the condition imposed on the surface tension α. Now for a type-I
superconductor, the surface tension α > 0, which implies 1 − 2vS/D > 0. Therefore, we
have 2vS/D < 1. For the typical value of σn ∼ 1026 sec−1 for the electrical conductivity
of normal quark matter, the profile velocity v < D/2S ∼ 10−6/Scm/sec ∼ 1cm/sec for
S ∼ 10−6cm. Therefore the interface velocity < 1cm/sec for such typical values of σn and
S to make the planer interface stable under small perturbation. Now the thickness of the
layer at the interface is l = 2D/v > 10−6cm for such values of D (or σn) and v. Here S is
always greater than 0, otherwise, the magnetic field strength at the normal-superconductor
interface becomes unphysical. As before, if the second term of eqn.(23) is negligibly small
compared to other two terms, we have
B(s) < " 4παB(c)
S
#1/3
(32)
For the sake of illustration, we have shown in fig.(1), the distribution of magnetic field
in a very small portion of horizontal plane at the perturbed interface. We have solved
eqn.(17) numerically and use the typical parameter set as given above. Along z-axis, we
have plotted the ratio of the strength of surface magnetic field and the critical strength
for the destruction of type I quark matter superconductivity and x and y axes represent
the x-y coordinates of this tiny horizontal plane. The distribution is extremely chaotic in
nature and some of the magnetic field peaks are stronger that the critical strength to destroy
type-I superconductivity. Then analogous to the case of crystal growth in an impure molten
alloy, where the process is stopped because of high density of accumulated impurities at the
interface, here also, further growth of superconducting phase will be stopped. Therefore, it
is a kind of instability at the interface, which forces the growth process to stop abruptly.
16
Since the slow growth of superconducting phase does not give a stable quark star with very
low surface magnetic field, is therefore not a physically acceptable phenomenon.
6. THE PROMPT PROCESS OF FLUX EXPULSION
We have noticed that the slow expulsion process, in which diffusion mechanism of mag-
netic lines of force may be applicable, leads to a kind of instability at the interface.
It
ultimately stops abruptly the growth process of superconducting zone. As a consequence
we will not get a stable quark star with very low surface magnetic field.
In this section, we assume that the superconducting phase of quark matter grows by a
faster process; so that the diffusion model for the magnetic lines of force is not applicable.
Which actually means, that the expulsion also occurs with a faster rate. Now during this
process, since magnetic flux changes rapidly with time, an emf will be induced in the system.
Although, the superconducting quark Cooper pairs do not feel electromagnetic force, a
Lorentz force will be exerted on the electrons near the surface region, which are in normal
phase. These electrons will try to nullify the force following Lenz's law. As a consequence
they will start gyrating about the magnetic lines of force at the surface / in the electro-
sphere. We have already argued that these circulation of the electrons about the magnetic
field lines will produce a diamagnetic effect to oppose the Lorentz force by reducing the
strength of magnetic field at the surface.
To get an estimate for induced magnetic field strength at the strange star surface /
electro-sphere and the corresponding reduction in magnetic field strength, we follow the
recent article by Hirsch [20]. The magnetic dipole moment per electron in the electro-sphere,
due to the tiny orbital motion is given by
µ =
eu
2
a
(we have assumed that c = 1)
(33)
where u is the orbital velocity and a is the radius of the orbit. The induced emf in the
electronic orbit is then given by
E =
a
2
∂B
∂t
Then the corresponding change in orbital speed is given by
∆u =
ea
2me
B
17
(34)
(35)
Hence the change in magnetic moment per electron is
∆µ =
e2a2
4me
B,
(36)
and the associated average value of the induced magnetization per unit volume in the electro-
sphere is given by
M =
nee2a2
4me
B
(37)
where ne is the average electron density in the electro-sphere (in reality, ne must be maximum
near the quark matter surface and is minimum at the outer surface of the electro-sphere).
Hence one can obtain the strength of induced magnetic field at the surface region, which is
given by
Bind = 4πM
(38)
Therefore the ratio of induced magnetic field by the electrons in the electro-sphere to the
existing strange star magnetic field is given by
Bind
B
=
nee2
me
πa2
(39)
The strength of induced magnetic field is therefore a monotonically increasing function of
electron density. It is very easy to verify that for complete suppression of magnetic field of
a strange star, with a typical average electron density ne = 10−5nB, where nB = 0.17fm−3,
the normal nuclear density, the orbital radius a ∼ 7fm. Which is about three orders of
magnitude less than the width of the electro-sphere. Therefore it is quite obvious that the
electro-sphere of width ∼ 1000fm can accommodate a large number of tiny magnets.
It
may be argued that the actual form of trajectories for such tiny magnets in each half of the
electro-sphere are closed helical spring.
Now, if there is no reduction, the magnetic field strength at the surface could be as high
as 1030 times the magnetic field of a typical neutron star. One can obtain this number by
considering the ratio R2/d2, where R = 10Km, the radius of the star and d = 1000fm, the
width of the electro-sphere and assuming that the magnetic flux remains conserved during
Meissner expulsion from super-conducting quark matter. This value is far above the upper
limit predicted by Shabad and Usov [36].
Now it has been discussed in the literature that if the density of quark matter, particularly,
at the core region is sufficiently high, the color super-conducting quark matter undergoes a
18
phase transition to what we call the CFL phase [37, 38, 39]. It is therefore expected that
at the ultra-dense interior, there will be a color as well as charge neutral CFL phase. The
number of quarks in the CFL cluster is even and divisible by three. It is found that in this
new phase all the three flavors u, d and s can form pairs with the same flavor as well as
with other components, i.e., their Fermi levels are identical. In this case the excess u-quarks
will therefore go to the outer region. The gross structure as mentioned at the beginning of
this article may not therefore be correct for a strange star with extremely high core density;
at the interior, it could be the CFL phase, at the outer core or inner crust region, it is the
positively charged color neutral quark matter in non-CFL color super-conducting phase and
finally the thin outer surface, the electro-sphere, is the normal electron gas. Whether the
formation of CFL phase at the inner part of a super-conducting strange star is an equivalent
QCD picture of the so called Tau-effect [40, 41] needs further investigation.
Therefore we expect that depending on the density of quark matter, the inner part of a
strange star is either a type-I super-conductor or in CFL phase, the outer region is color
neutral type-I super-conducting non-CFL phase, mainly dominated by u-quarks and some
electrons, and finally the outer surface is a thin layer of electron gas in normal state. The
magnetic field lines are expelled from quark matter phase to the electron gas layer with a
small penetration at the outer surface of the quark matter. We have conjectured that the
classical diamagnetism of the normal electron gas suppresses the magnetic field of a strange
star by diluting the existing magnetic flux lines in the electro-sphere by the repulsive action
of the induced magnetic field.
If B ≈ Bind, we have almost total suppression.
In this
ideal situation, so far the emission due to magnetic activities are concerned, the object
becomes dark to the observers. However, because of vacuum instability, non-thermal surface
γ emission is possible. Since Z is very high, the Coulomb field at the quark matter surface
will be extremely strong, then e+e− pairs may be produced from vacuum by Schwinger
mechanism [42]. The created e+ gets annihilated with one of the electrons in the electro-
sphere and e− will occupy one of the empty energy levels. Therefore, γ emission will also
be forbidden if the star is extremely cold, when there will be no more vacant states for the
produced electrons [43].
An alternative explanation for low magnetic field of strange stars can also be obtained
from a very simple model of magnetic circuits. The combination of electro-sphere and the
magneto-sphere may be treated as an equivalent magnetic circuit with varying reluctance.
19
Of course, there are some open flux lines goes out after cutting the light cylinder. In that
case, we may assume that the whole universe is a complicated network of magnetic circuits
with varying reluctance and the strength of sources of magneto motive force. From the
definition, we have total flux
N =
MMF
R
(40)
where MMF is the effective magneto motive force and R is the reluctance, which is given by
R = I dl
αµ
(41)
where dl is the length of some flux carrying element of cross section α and magnetic perme-
ability µ. Let us consider the passage of expelled magnetic flux lines through surface region,
when there is no gyrating motion of electrons. Here we may take µ ≈ 1, the reluctance is
then given by.
+ Rms
(42)
where es indicates the electro-sphere and ms is the magneto-sphere. On the other hand,
with the gyrating electrons, since there is a diamagnetic effect at the surface / electro-sphere,
we must have the permeability µ < 1 in this region. In Gaussian unit, we have µ = 1 − 4πχ.
Where from eqn.(37) the susceptibility is given by
R1 = Z dl
α(cid:12)(cid:12)(cid:12)es
(43)
(44)
χ =
∂M
∂B
=
nee2a2
4me
The reluctance is then given by
R2 = Z dl
µα′(cid:12)(cid:12)(cid:12)es
+ Rms
where α′ is now the new cross sectional area of the flux carrying element. Since the MMF
source and the total flux does not change by the gyration of electrons at the surface, we
have effectively
which gives
R1 = R2
α′ = α
1
µ
(45)
(46)
Since µ ≪ 1, it is therefore quite obvious that the area of cross section of the electro-sphere
will increase by several orders of magnitude. As a result the magnetic flux density will
also be reduced by the same factor. Now it is well known that Larmor radius a ∼ B−1,
20
here B is the surface magnetic field. Therefore it is quite possible in the parameter space
(electron density and the corresponding Larmor radius) for the physically acceptable values
for electron density and Larmor radius, the surface magnetic field can be as low as 108G.
For the typical values of electron density ne ∼ 10−12n0 and Larmor radius a ∼ 10 A, the
magnetic permeability becomes ∼ 0.012. Which gives surface magnetic field a few times
107G. Since the Larmor radius increases with the decrease in magnetic field strength, it will
further increase the induced magnetic field (see eqn.(39)), if it dominates over the electron
density. The induced magnetic field, which is diamagnetic in nature will further reduce the
existing field. So the solution must be self-consistent in nature. We believe that ultimately
a steady state will be established in the electro-sphere. This is true for both the models.
7. CONCLUSION
From this investigation, we may conclude in a straight forward way that if strange stars
really exists with very low surface magnetic field, the type-I superconducting transition in
the quark matter phase must have occurred within the star. Further, the transition process
must be very prompt, so that diffusion model for the expulsion of magnetic flux lines is not
applicable, where the later gives rise to some kind of instability at the interface and finally
stops the growth of superconducting phase abruptly. Further, the slow transition model
does have any mechanism to reduce the surface field strength, which is ≤ 108G.
Our next conclusion is that the transition is quite fast. The classical diamagnetism of
gyrating electrons at the surface / electro sphere may be the possible source of opposing
field to reduce the expelled surface magnetic field by several orders of magnitude. It is quite
obvious that with this prompt transition model, using the idea of Hirsch, one can obtain a
surface magnetic field for the expected quark stars as low as 108G.
[1] J.E. Hirsch, Phys. Lett. A134, 451 (1989).
[2] J.E. Hirsch, Phys. Rev. B68, 184502 (2003).
[3] J.E. Hirsch, Phys. Rev. B71, 184521 (2005); Phys. Lett. A345, 453 (2005).
[4] E. Witten, Phys. Rev. D30, 272 (1984).
[5] A.R. Bodmer, Phys. Rev. D4, 1601 (1971).
21
[6] E. Farhi and R.L Jaffe, Phys. Rev. D30, 2379 (1984).
[7] S. Chakrabarty, S. Raha and B. Sinha, Phys. Lett. B229,112 (1989); S. Chakrabarty, Phys.
Rev. D43, 627 (1991).
[8] G. Baym, R.L. Jaffe, E.W. Kolb, L. Mclerran and T.P. Walker, Phys. Lett. B160, 181 (1985);
C. Alcock and E. Farhi, Phys. Rev. D32, 1273 (1985).
[9] P. Haensel, J.L. Zdunik and R. Schaeffer, Astr. and Astrophys. 160, 121 (1986).
[10] see for example the Proceeding of the Workshop on Strange Quark Matter in Physics and
Astrophysics, May 20-24, 1991, University of Aarhus, Denmark, Nucl. Phys. B24 (1991).
[11] J. Schaffner-Bielich, C. Greiner, A. Diener and H. Stoecker, Phys. Rev. C55, 3038 (1997).
[12] J. Madsen, Phys. Rev. Lett. 85, 4687 (2000).
[13] Observations released by the Chandra X-ray observatory predicted that compact objects
RXj1856.5-3754 and 3C58, where the later one is associated with the supernova SN1181 could
be the strange quark stars. Another such object XTEj1739-285, observed by a team led by
Philip Karret. It is a fast rotating compact object (1122 rotation per second). stars.
[14] M. Alford, K. Rajagopal, T. Schaefer and A. Schmitt, arXiv:0709.4632 (hep-ph) (to appear
in Rev. Mod. Phys.).
[15] B. Muller and J. Rafelski, Phys. Rev. Lett. 34, 349 (1975); B. Muller, Ann. Rev. Nucl. Sci.
26, 351 (1976).
[16] J. Madsen, Phys. Rev. Lett. 100, 151102 (2008).
[17] G. Baym, C. Pethick and D. Pines, Nature 224, 673 (1969); S.L. Shapiro and S.A. Teukolsky,
Black Holes, White Dwarfs and Neutron Stars, Wiley, New York (1983).
[18] M. Baldo, et. al., Nucl. Phys. A536, (1992) 349; S. Chakrabarty, Astrophysics & Space
Science, 314, (2008) 105.
[19] D. Bailin and A. Love, Phys. Rep. 107, 325 (1984).
[20] J. Hirsch, arXiv:cond-mat/0301611.
[21] J. Hirsch, arXiv:cond-mat/0508529; arXiv:0803.2054.
[22] Bailin B. and Love A., 1982, Nucl. Phys. B205, 119.
[23] Chau A.F., 1997, Ap. J. 479, 886.
[24] Alford M.G., Berges J. and Rajagopal K., 2000, Nucl. Phys. B571, 269.
[25] Losert W., Shi B.Q. and Cummins H.Z., 1998, Proc. Natl. Acad. Sci. USA (Applied Physical
Science) 95, 431.
22
[26] Adams F.C, Freese K. and Langer J.S., 1993, Phys. Rev. D47, 4303.
[27] Kamionkowski Marc and Freese K., 1992, Phys. Rev. Lett. 69, 2743.
[28] Mullins W.W. and Sekerka R.F., 1963, Jour. Appl. Phys. 34, 323.
[29] Mullins W.W. and Sekerka R.F., 1964, Jour. Appl. Phys. 1964; 35, 444.
[30] Langer J.S., 1980, Rev. Mod. Phys. 52, 1.
[31] Haensel P. and Jerzak A.J., 1989, Acta Phys. Pol. B20, 141.
[32] Haensel P., 1991, Nucl. Phys. B24 (proc. of the Int. Workshop on Strange quark Matter in
Physics and Astrophysics, Univ. of Aarhus, Denmark, May 20-24, 1991), 23.
[33] Heiselberg H and Pethick C.J, 1993, Phys. Rev. D48, 2916.
[34] Chakrabarty S., 1995, Phys. Rev. D51, 4591.
[35] Ghosh T. and Chakrabarty S., 2001, Phys. Rev. D63 043006.
[36] A.E. Shabad and V.V. Usov, Phys. Rev. Lett. 96, 180401-1 (2006); astro-ph/0601542.
[37] M. Alford, K. Rajagopal and F. Wilczek, Phys. Lett. B422, 247 (1998).
[38] M. Alford, K. Rajagopal, T. Schafer and A. Schmitt, arXiv:0709.4635.
[39] R. Rapp, T. Schafer, E. Shuryak and M. Velkovsky, Phys. Rev. Lett. 81, 53 (1998).
[40] J.E. Hirsch, Phys. Rev. Lett. 94, 187001 (2005).
[41] R. Tau, X. Zhang, X. Tang and P.W. Anderson, Phys. Rev. Lett. 83, 5575 (1999).
[42] J. Schwinger, Phys. Rev. 82, 664 (1951).
[43] V.V. Usov, Astrophys. Jour 550, L179 (2001).
23
4
3.5
3
2.5
2
1.5
1
0.5
0
0
10
20
30
40
50 0
20
10
50
40
30
FIG. 1: The variation of magnetic field along z-axis (plotted as a ratio of B/Bcr, where
Bcr ∼ 1016G, the critical strength to destroy the type-I quark matter superconductivity) with
the orthogonal coordinates x and y.
24
|
astro-ph/9710129 | 1 | 9710 | 1997-10-13T18:01:39 | Breaking the degeneracy between anisotropy and mass: The dark halo of the E0 galaxy NGC 6703 | [
"astro-ph"
] | (abridged) We have measured line-of-sight velocity profiles (VPs) in the E0 galaxy NGC 6703 out to 2.6 R_e. From these data we constrain the mass distribution and the anisotropy of the stellar orbits in this galaxy.
We have developed a non-parametric technique to determine the DF f(E,L^2) directly from the kinematic data. From Monte Carlo tests using the spatial extent, sampling, and error bars of the NGC 6703 data we find that smooth underlying DFs can be recovered to an rms accuracy of 12%, and the anisotropy parameter beta(r) to an accuracy of 0.1, in a given potential. An asymptotically constant halo circular velocity v_0 can be determined with an accuracy of +- \lta 50km/s.
For NGC 6703 we determine the true circular velocity at 2.6 R_e to be 250 +- 40km/s at 95% c.l., corresponding to a total mass in NGC 6703 inside 78'' (13.5 h_50^-1 kpc), of 1.6-2.6 x 10^11 h_50^-1 Msun. No model without dark matter will fit the data; however, a maximum stellar mass model in which the luminous component provides nearly all the mass in the centre does. In such a model, the total luminous mass inside 78'' is 9 x 10^10 Msun and the integrated M/L_B=5.3-10, corresponding to a rise from the center by at least a factor of 1.6.
The anisotropy of the stellar distribution function in NGC 6703 changes from near-isotropic at the centre to slightly radially anisotropic (beta=0.3-0.4 at 30'', beta=0.2-0.4 at 60'') and is not well-constrained at the outer edge of the data.
Our results suggest that also elliptical galaxies begin to be dominated by dark matter at radii of \sim 10kpc. | astro-ph | astro-ph | Mon. Not. R. Astron. Soc. 000, 000 -- 000 (0000)
Printed July 5, 2021
(MN LATEX style file v1.4)
Breaking the degeneracy between anisotropy and mass:
The dark halo of the E0 galaxy NGC 6703
Ortwin Gerhard1,2, Gunther Jeske2, R. P. Saglia3,⋆, Ralf Bender3,⋆
1 Astronomisches Institut, Universitat Basel, Venusstrasse 7, CH-4102 Binningen, Switzerland
2 Landessternwarte, Konigstuhl, D-69117 Heidelberg, Germany
3 Institut fur Astronomie und Astrophysik, Scheinerstr. 1, D-81679 Munchen, Germany
7
9
9
1
Accepted October 1997. Received May 1997, in original form June 1995.
t
c
O
3
1
1
v
9
2
1
0
1
7
9
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
ABSTRACT
We have measured line-of-sight velocity profiles (vps) in the E0 galaxy NGC 6703
out to 2.6Re. Comparing with the vps predicted from spherical distribution functions
(dfs), we constrain the mass distribution and the anisotropy of the stellar orbits in
this galaxy.
We have developed a non-parametric technique to determine the df f (E, L2)
directly from the kinematic data. We test this technique on Monte Carlo simulated
data with the spatial extent, sampling, and error bars of the NGC 6703 data. We find
that smooth underlying dfs can be recovered to an rms accuracy of 12% inside three
times the radius of the last kinematic data point, and the anisotropy parameter β(r)
to an accuracy of 0.1, in a known potential. These uncertainties can be reduced with
improved data.
By comparing such best-estimate, regularized models in different potentials,
we can derive constraints on the mass distribution and anisotropy. Tests show that
with presently available data, an asymptotically constant halo circular velocity v0 can
be determined with an accuracy of ± ∼< 50 km s−1. This formal range often includes
high -- v0 models with implausibly large gradients across the data boundary. However,
even with extremely high quality data some uncertainty on the detailed shape of the
underlying circular velocity curve remains.
50 kpc, where h50 ≡ H0/50km/s/Mpc) of 1.6 − 2.6 × 1011h−1
In the case of NGC 6703 we thus determine the true circular velocity at 2.6Re
to be 250 ± 40 km s−1 at 95% confidence, corresponding to a total mass in NGC 6703
inside 78′′ (13.5 h−1
50 M⊙. No
model without dark matter will fit the data; however, a maximum stellar mass model
in which the luminous component provides nearly all the mass in the centre does. In
such a model, the total luminous mass inside 78′′ is 9 × 1010 M⊙ and the integrated
B-band mass -- to -- light ratio out to this radius is ΥB = 5.3 − 10, corresponding to a
rise from the center by at least a factor of 1.6.
The anisotropy of the stellar distribution function in NGC 6703 changes from
near-isotropic at the centre to slightly radially anisotropic (β = 0.3 − 0.4 at 30",
β = 0.2 − 0.4 at 60") and is not well-constrained at the outer edge of the data, where
β = −0.5 − +0.4, depending on variations of the potential in the allowed range.
Our results suggest that also elliptical galaxies begin to be dominated by dark
matter at radii of ∼ 10 kpc.
Key words:
galaxies: kinematics and dynamics -- galaxies: individual -- line profiles
stellar dynamics -- dark matter -- galaxies: elliptical and lenticular --
⋆ Visiting Astronomer of the German-Spanish Astronomical Cen-
ter, Calar Alto, operated by the Max Plank Institut fur As-
tronomie, Heidelberg, jointly with the Spanish National Com-
mission for Astronomy
Current cosmological models predict that, similar to spiral
galaxies, elliptical galaxies should be surrounded by dark
matter haloes. The observational evidence for dark matter
1 INTRODUCTION
c(cid:13) 0000 RAS
2 O.E. Gerhard, G. Jeske, R.P. Saglia, R. Bender
in ellipticals is still weak, however. In a few cases masses
have been determined from X-ray observations (e.g., Awaki
et al. 1994, Kim & Fabbiano 1995) or HI ring velocities
(Franx, van Gorkom & de Zeeuw 1994). In others it has
been possible to rule out constant M/L from extended ve-
locity dispersion data (Saglia et al. 1993), from absorption
line profile measurements (Carollo et al. 1995, Rix et al.
1997), or from globular cluster or planetary nebula velocities
(e.g., Grillmair et al. 1994, Arnaboldi et al. 1994). Gravita-
tional lensing statistics (Maoz & Rix 1993) and individual
image lens separations (Kochanek & Keeton 1997) favour
models with extended dark matter haloes around ellipticals.
Despite of this, the detailed radial mass distribution in ellip-
tical galaxies remains largely unknown. Similarly, although
we know from the tensor virial theorem that giant ellipti-
cals are globally anisotropic (Binney 1978), their detailed
anisotropy structure is only poorly known.
The origin of this uncertainty is a fundamental de-
generacy -- in general, it is impossible to disentangle the
anisotropy in the velocity distribution and the gravitational
potential from velocity dispersion and rotation measure-
ments alone (Binney & Mamon 1982, Dejonghe & Merritt
1992). Tangential anisotropy, for example, can mimic the
presence of dark matter. Recent dynamical studies have in-
dicated, however, that the anisotropy of the stellar distribu-
tion function (df) is reflected in the shapes of the line-of-
sight velocity profiles (vps) in a way that depends on the
gravitational potential (Gerhard 1993, G93; Merritt 1993).
These papers argued that the extra constraints derived from
the vp measurements may be enough to break the degener-
acy and determine the mass distribution.
If this is correct, it provides a new method to investi-
gate the properties of the dark matter haloes around ellipti-
cal galaxies at intermediate radii: vps can now be estimated
from high-quality absorption line measurements out to ∼ 3
effective radii. Dynamical models are then used to disentan-
gle the effects of orbital anisotropy and potential gradient
on the vp shapes. In this paper, we implement these ideas
for the analysis of real data, analyzing the E0 galaxy NGC
6703. This study is part of an observational and theoretical
program aimed at understanding the mass distribution and
orbital structure in elliptical galaxies. Preliminary accounts
of this work have been given in Jeske et al. (1996) and Saglia
et al. (1997a).
We have obtained long -- slit spectroscopy for NGC 6703,
and have measured vps to ∼ 2.6Re with the method of Ben-
der (1990). The results are quantified by a Gauss-Hermite
decomposition (G93, van der Marel & Franx 1993) as de-
scribed by Bender, Saglia & Gerhard (1994; BSG). These
observations are described in Section 2. In Section 3 we use
simple dynamical models to describe the variation of the
vp shapes with anisotropy and potential, generalizing the
results of G93 for scale -- free models. These models, taken
from a systematic study of the relation between df and vps
in spherical potentials (Jeske 1995), are described in Ap-
pendix A. In Section 4 we develop a non-parametric method
for inferring the df and potential from absorption line pro-
file measurements. Tests on Monte Carlo generated data are
used to determine the degree of confidence with which the
df and potential can be inferred from real data. In Section
5 we analyse the kinematic data for NGC 6703. Compari-
son with the dynamical models from Jeske (1995) already
shows that no constant -- M/L model will fit the data. The
non -- parametric method developed in Section 4 is then used
to derive quantitative constraints on the mass distribution
and anisotropy of this galaxy. Finally, Section 6 presents a
discussion of the results and our conclusions.
2 OBSERVATIONS OF NGC 6703
NGC 6703 is an E0 galaxy at a distance D = 36 Mpc
(Faber et al. 1989) for H0 = 50 km/s/Mpc. From a B-
band CCD frame taken at the prime focus of the 3.5m
telescope on Calar Alto and kindly provided by U. Hopp
we have measured the inner surface brightness profile us-
ing the algorithm by Saglia et al. (1997b); this follows an
R1/4 law with Re = 30′′ = 5.2 h−1
50 kpc, or a Jaffe model with
rJ = 46.5′′ = 8.1 h−1
50 kpc, with small residuals (Fig. 1). A
3% increase of the sky value reduces the measured Jaffe ra-
dius to 35.5′′, a 1% decrease of the sky values increases it
to 54′′. Isophote shapes deviate little from circles (ǫ < 0.05,
a4/a < 0.005) and show small twisting (∆P A ≃ 10◦).
From the Jaffe profile fit we derive a fiducial (calibrated and
corrected for galactic absorption following Faber et al. 1989)
MB =−21.07, or luminosity LB = 4.16× 1010h−2
50 L⊙,B. Note
that the values of Re and MB derived here are slightly larger
than those (Re = 24′′, MB = −20.79) given by Faber et al.
(1989).
The spectroscopic observations were carried out in Oc-
tober 1994, May 1995 and August 1995 with the 3.5-m tele-
scope on Calar Alto, Spain. In all of the runs the same setup
was adopted. The Boller & Chivens longslit twin spectro-
graph was used with a 1200 line mm−1 grating giving 36
A/mm dispersion. The detector was a Tektronix CCD with
1024×1024 24µm pixels and the wavelength range 4760-5640
A. The instrumental resolution obtained using a 3.6 arcsec
wide slit was 85 km/s. We collected 1.5 hrs of observations
along the major axis of the galaxy, 4 hrs of observations per-
pendicular to the major axis and shifted to the North-East
of the center by 24 arcsec (0.8Re), and 13 hours of obser-
vations perpendicular to the major axis and shifted to the
North-East of the center by 36 arcsec (1.2Re). Spectra taken
parallel to the minor axis and shifted from the center allow
at the same time a good sky subtraction and the symmetry
check of the data points.
The analysis of the data was carried out following the
steps described by BSG. The logarithmic wavelength cali-
bration was performed at a smaller step (∆v = 30 km/s)
than the actual pixel size (≈ 50 km/s) to exploit the full ca-
pabilities of the Fourier Correlation Quotient method. A sky
subtraction better than 1% was achieved. The heliocentric
velocity difference between the May 1995 and the Septem-
ber 1994 - August 1995 frames was taken into account before
coadding the observed spectra. The spectra were rebinned
along the spatial direction to obtain a nearly constant sig-
nal to noise ratio larger than 50 per resolution element. The
effects of the continuum fitting and instrumental resolution
were extensively tested by Monte Carlo simulations. The
residual systematic effects on the values of the h3 and h4
parameters are expected to be less than 0.01. The resulting
fitted values for the folded velocity v , velocity dispersion
σ, h3 and h4 profiles are shown in Fig. 1 as a function of
the distance from the center, reaching ∼ 2.6Re. The v and
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
The dark halo of the E0 galaxy NGC 6703
3
points) is about 20 % (determined from the scatter of the
estimated signal to noise ratios).
Data points at the largest distances for the different
data sets have lower signal to noise ratio than the mean
and therefore have larger error bars. In addition, these data
points are expected to suffer more from the systematic ef-
fects due to the galaxy light contamination of the sky sub-
traction(see discussion in Saglia et al. 1993). They are in any
case consistent within the error bars with the more accurate
values derived from the other available spectra.
The observed scatter is sometimes slightly larger than
expected from the error estimates. This excess could be real
and due to the faint structures apparent in an unsharped
mask image of the galaxy. In particular this applies to the
asymmetries observed in the h4 profile in the central 5 arc-
sec. The negative h4 values detected for the first data points
of the 24 arcsec shifted spectrum are also real. They are de-
tected in the unbinned major axis spectra at R ≈ 22 arcsec
(dots in Fig. 1).
3 VELOCITY PROFILES IN SPHERICAL
GALAXIES
To better understand the relation between vp -- shape,
anisotropy, and gravitational potential, we have constructed
a large number of anisotropic models for spherical galaxies
in which the stars follow a Jaffe (1983) profile. The gravi-
tational potential was taken to be either that of the stars
(self-consistent case), or one with everywhere constant rota-
tion speed ('halo potential'). The latter case corresponds to
a mass distribution with a dark halo which has equal density
as the stars at r ≃ 0.4rJ , and equal interior mass at r ≃ rJ ,
where rJ is the scale radius of the Jaffe model. Anisotropic
quasi-separable distribution functions (dfs) g(E) h(E, L2)
were calculated by the method of Gerhard (1991; G91), but
contrary to G93 the circularity function (which specifies the
distribution of angular momenta on energy shells) was al-
lowed to vary with energy. These models include dfs in
which the anisotropy changes radially, from tangential to
radial or vice-versa, from isotropic to radial to tangential,
etc. (see Fig. 18). For comparison, we have also constructed
families of Osipkov (1979) - Merritt (1985) models which be-
come strongly radially anisotropic beyond a certain radius.
Details of this model data base and the properties of their
vps are given in Appendix A and in Jeske (1995); here we
only give a brief summary relevant for the comparison with
NGC 6703.
The shapes of the observable vps are most sensitive to
the anisotropy of the df, but depend also on the poten-
tial (G93). For rapidly falling luminosity profiles, the vps
are dominated by the stars at the tangent point. Then ra-
dially (tangentially) anisotropic dfs lead to more peaked
(more flat-topped) vps than in the isotropic case; in terms
of the Gauss-Hermite parameter h4, this corresponds to
h4 > (h4)iso and h4 < (h4)iso, respectively (Figs. 8,9 in G93).
Fig. 2 shows that these trends are also seen in the
present models in which both the luminosity density and the
anisotropy change with radius. An increase in radial (tan-
gential) anisotropy at intrinsic radius r is accompanied by
an increase (decrease) of h4 at projected radius R≃ r. The
correspondence is strongest in the models' outer parts, but
Figure 1. (a) Residuals of a Jaffe model fit to photometry for
NGC 6703. Full line: in surface brightness; dotted line: in the
curve of growth. The vertical dotted line marks the scaling Jaffe
radius rJ . (b) Folded mean velocity, (c) velocity dispersion, (d)
h3, and (e) h4 profiles. Crosses and filled circles refer to the two
sides of the galaxy and the major axis spectrum. The small dots
refer to the unrebinned spectrum (see text). Open and filled tri-
angles refer to the two sides of the galaxy and the spectrum taken
parallel to the minor axis and shifted 24 arcsec from the center.
Open and filled squares refer to the two sides of the galaxy and the
spectrum taken parallel to the minor axis and shifted 36 arcsec
from the center.
σ profiles are folded antisymetrically with respect to the
center for the major axis spectrum, symmetrically with re-
spect to the major axis for the spectra parallel to the minor
axis. Template mismatching was minimized by choosing the
template star which gave the minimal symmetric h3 profile
derived along the major axis of the galaxy. The systematic
effect due to the residual mismatching on the derived h4 val-
ues estimated from the remaining symmetric components is
less than 0.01.
The galaxy shows very little rotation (≈ 0 km/s for
R < Re, ≈ 20 − 30 km/s for R > Re). The (cylindrical) ro-
tation measured parallel to the minor axis is slightly larger
(≈ 35 km/s) along the 24 arcsec shifted spectrum, but con-
sistent with the peak velocity reached along the major axis
at R ≈ 22 arcsec. This is shown by the velocities derived
from the unbinned major axis spectra (dots in Fig. 1). The
velocity dispersion drops from the central ≈ 190 km/s to
≈ 140 km/s at Re/2, slowly declining to about 110 km/s in
the outer parts. The h3 and h4 values are everywhere close
to zero. The error bars are determined from Monte-Carlo
simulations. Noise is added to template stars (rebinned to
the original wavelength pixel size) broadened following the
observed values of σ and h4, matching the power spectrum
noise to peak ratio of the galaxy spectra. The accuracy of
the estimated error bars (the r.m.s. of 30 replica of the data
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
4 O.E. Gerhard, G. Jeske, R.P. Saglia, R. Bender
Jaffe, self
Jaffe, self
Jaffe, self
Jaffe, self
Jaffe, self
Jaffe, self
Jaffe, Halo
Jaffe, Halo
Jaffe, Halo
Jaffe, Halo
Jaffe, Halo
Jaffe, Halo
1
1
1
1
1
1
0.8
0.8
0.8
0.8
0.8
0.8
0.6
0.6
0.6
0.6
0.6
0.6
0.4
0.4
0.4
0.4
0.4
0.4
0.2
0.2
0.2
0.2
0.2
0.2
0
0
0
0
0
0
-0.2
-0.2
-0.2
-0.2
-0.2
-0.2
-0.4
-0.4
-0.4
-0.4
-0.4
-0.4
-0.6
-0.6
-0.6
-0.6
-0.6
-0.6
-0.8
-0.8
-0.8
-0.8
-0.8
-0.8
-1
-1
-1
-1
-1
-1
0.8
0.8
0.8
0.8
0.8
0.8
0.4
0.4
0.4
0.4
0.4
0.4
0
0
0
0
0
0
0.16
0.16
0.16
0.16
0.16
0.16
0.12
0.12
0.12
0.12
0.12
0.12
0.08
0.08
0.08
0.08
0.08
0.08
0.04
0.04
0.04
0.04
0.04
0.04
0
0
0
0
0
0
-0.04
-0.04
-0.04
-0.04
-0.04
-0.04
-0.08
-0.08
-0.08
-0.08
-0.08
-0.08
0.01
0.01
0.01
0.01
0.01
0.01
0.1
0.1
0.1
0.1
0.1
0.1
1
1
1
1
1
1
10
10
10
10
10
10
0.01
0.01
0.01
0.01
0.01
0.01
0.1
0.1
0.1
0.1
0.1
0.1
1
1
1
1
1
1
10
10
10
10
10
10
Figure 2. Fitted projected velocity dispersion σfit, anisotropy parameter β, and vp-parameter h4 for representative Jaffe models in
self-consistent (left) and halo potential (right). The models shown are radially and tangentially anisotropic models constructed with the
method of G91 (dashed lines), the isotropic model (solid), and two Osipkov-Merritt models (dotted lines). Note that while β is a function
of three-dimensional radius r and σfit, h4 are observed quantities depending on projected radius R, there is a close correspondence
between features in these profiles. See text.
is also seen to a lesser extent in the centre of a Jaffe model
where ρ(r) ∝ r−2 -- contrary to a homogeneous core where
radial orbits lead to broadened vps (Dejonghe 1987). Quan-
titatively the correspondence depends also on the anisotropy
gradient. Osipkov-Merritt-models show a reversal of this
trend near their anisotropy radius ra due to the large num-
ber of high-energy radial orbits all turning around near ra;
this leads to flat-topped vps in a small radius range near
ra. However, the properties of these models are extreme and
they are in general not very useful for modelling observed
vps.
Fig. 2 and Figs. 8,9 in G93 also show that as the mass
of the model at large r is increased at constant anisotropy,
both the projected dispersion and h4 increase. Increasing
β at constant potential, on the other hand, lowers σ and
increases h4. This suggests that by modelling σ and h4 both
mass M (r) and anisotropy β(r) can in principle be found.
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
N (R, vk) = Z dzZZ dvxdvyf (E, L2)
= Z dzZZ dvxdvyf [ 1
(1)
2 v2
r + 1
2 v2
t + Φ(r), r2v2
t ].
4 MODELLING ABSORPTION LINE PROFILE
DATA
Having seen the effect of anisotropy and potential variations
on the line profile parameters, we now proceed to construct
an algorithm by which the distribution function and poten-
tial of a spherical galaxy can be constrained from its ob-
served σ and h4 -- profiles. Such absorption line profile data
contain a subset of the information given by the projected
distribution function N (R, vk), which in the spherical case
is related to the full df by
Here the position on the sky is specified by R = (x, y);
R = R. Velocities in the sky plane are denoted by (vx, vy),
z and vk are the line-of-sight position and velocity, and vr
and vt are the intrinsic radial and tangential velocities. In
spherical symmetry, the df f (E, L2) is a function of energy
and squared angular momentum only.
Notice from eq. (1) that the projected df depends lin-
early on f , but non-linearly on the potential Φ(r). Thus
f will generally be easier to determine from N (R, vk) than
Φ. Moreover, while considerations like those in the last sec-
tion do suggest that, in spherical symmetry and for positive
f (E, L2), both the df and the potential can be determined
from N (R, vk), there is no theoretical proof that this is in
fact true. We only know that in a fixed spherical potential
the df is uniquely determined from N (R, vk) (Dejonghe &
Merritt 1992). For these reasons we have found it useful to
split our problem into two parts (see also Merritt 1993):
(1) We fix the potential Φ, and from the photometric
and kinematic data determine the "best" df f for this po-
tential. Because in practice the surface brightness (sb) pro-
file is much better sampled than the kinematic observations
and also has smaller errors, we treat it separately and de-
termine the stellar luminosity density j(r) at the beginning.
The kinematic data are then used to determine the "best" f
for given j(r) and Φ(r), by approximately solving equation
(1) as a linear integral equation.
(2) We then vary Φ to find that potential which allows
the best fit overall. At present, it is not practical in step (2)
to attempt to determine the potential non-parametrically.
Rather, we choose a parametrized form for Φ, and find the
region in parameter space for which the "best" df as deter-
mined in step (1) reproduces the data adequately.
In view of the modelling of NGC 6703 in Section 5, we
have considered the following family of potentials, includ-
ing a luminous and a dark matter component: The stellar
component is approximated as a Jaffe (1983) sphere, with
scale-radius rJ and total mass MJ , so that
ΦL(r) =
GMJ
rJ
ln
r
r + rJ
.
The dark halo has an asymptotically flat rotation curve,
vc(r) = v0
r
√r2 + r2
c
,
so that its potential is
ΦH (r) = 1
2 v2
0 ln(r2 + r2
c ).
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
(2)
(3)
(4)
The dark halo of the E0 galaxy NGC 6703
5
This is specified by the asymptotic circular velocity v0 and
the core radius rc. Both the luminous and dark halo compo-
nents can be modified when needed, and need not be analytic
functions.
In testing our method below, we use parameters
adapted to NGC 6703. This galaxy is well-fit by a Jaffe
profile (Section 2), so rJ is known. This leaves three free
parameters, the mass MJ or mass -- to -- light ratio Υ of the
stellar component, and the halo parameters rc and v0. If
one assumes that the central kinematics is dominated by
the luminous matter, Υ can be determined. Then only the
two halo parameters rc and v0 are free. The assumption of
maximum stellar Υ is similar to the maximum disk assump-
tion in spiral galaxies.
In any of the potentials specified by eqns. (2) -- (4) we
determine the df by the algorithm described in Section 4.1
below. To assess the significance of the results obtained, we
test the algorithm on Monte Carlo -- generated pseudo data in
Section 4.2. For kinematic data with the spatial extent and
observational errors such as measured for NGC 6703, the
algorithm recovers a smooth spherical df ∼ 70% of the time
to an rms level of ∼ 12%, taken inside three times the radius
of the outermost kinematic data point. In Section 4.3, we
investigate the degree to which the gravitational potential
can be constrained from similar data.
4.1 Recovering f from σ and h4, given Φ
As discussed above, the projected distribution function
N (r, vk) suffices to determine df f (E, L2) uniquely. In prac-
tice, however, only incomplete and noisy data are available
in place of N (r, vk), and contrary to the two-dimensional
function N (r, vk), the observed σ(Ri) and h4(Ri) contain
only one-dimensional information. This suggests that we can
hope to recover only the gross features of f from such kine-
matic data. Indeed, the anisotropy parameter β(r) seems
to be essentially fixed from accurate h4 measurements (e.g.,
Figs. 8,9 in G93). Local fluctuations in the df will be inac-
cessible, but as we will show, smooth dfs can be recovered
with reasonable accuracy from presently available data.
To solve the inversion problem, we first compute a set of
self -- consistent models fk(E, x) for the stellar density j(r),
in the fixed potential Φ(r). The fk(E, x) are models of the
kind discussed in Section 2 and Appendix A; E and x are
the energy and circularity integrals of the motion. Then we
write the df as a sum over these "basis" functions:
akfk(E, x).
(5)
f = Xk=1,K
We do not need to use a doubly infinite, complete set of
basis functions because most of the high -- frequency struc-
ture represented by the higher -- order basis functions in such
a set will be swamped by noise in the observational data.
It is sufficient to choose the number of basis functions, K,
and the fk(E, x) themselves such that the data can be fit-
ted with a mean χ2 ≃ 1 per data point. We have found it
advantageous to use the isotropic model plus tangentially
anisotropic basis models, because with these the anisotropy
of the final composite df (5) can be varied in a more local
way than with radially anisotropic components. Since the
ak can be negative, it is of course no problem to generate a
6 O.E. Gerhard, G. Jeske, R.P. Saglia, R. Bender
(E = Ei, x = xj). Subject to these constraints the ak are
to be determined such that the kinematics predicted from
the df (5) match the observed kinematics in a minimum
χ2 -- sense.
The comparison between model and data is not entirely
straightforward, however. Since the measured (v, σ, h3, h4)
are obtained by fitting to the line profile (the observational
analogue of the projected df), they unfortunately depend
non-linearly on the galaxy's underlying df. Thus we cannot
use the observed v and σ in a linear least squares algorithm
to determine the ak from the data -- they cannot be writ-
ten as moments of f . In the fitting process, they have to be
replaced by quantities that do depend linearly on the df.
Moreover, the error bars for these new quantities generally
depend not only on the observed error bars of v and σ, but
also on the errors and the values of the line profile param-
eters h3 and h4. They must therefore be determined with
some care.
We have investigated several schemes along these lines.
The following seemed to do best in recovering a known un-
derlying spherical df from pseudo data. From the measured
(σ, h4), we compute an approximation to the true velocity
dispersion (second moment) σ2(Ri), by integrating over the
line profile (for negative h4, until it first becomes negative).
We also evaluate a new set of even Gauss-Hermite moments
The velocity profile moments of the basis function mo-
isotropic model with the galaxy's stellar density, in the cur-
rent potential Φ(r).
sn(Ri;eσ) from the data, using fixed, fiducial velocity scales
eσ(Ri) (G93, Appendix B). We have found it convenient to
take for theseeσ(Ri) the velocity dispersions σiso(Ri) of the
dels are transformed to the same velocity scaleseσ(Ri). The
moments σ2(Ri) and sn(Ri;eσ) of the composite df then
depend linearly on the corresponding moments of the basis
function models. For a regularized model they are smooth
functions of R, of which the (noisy) observational moments
are assumed to be a random realisation within the respective
errors.
To determine the best-fitting coefficient ak we minimize
the sum over data points i of all
(8)
(9)
σ,i ≡ w2
χ2
and
χ2
n,i ≡ w2
akσ2
σ(Ri)hσ2(Ri) − Xk=1,K
n(Ri)hsn(Ri;eσ) − Xk=1,K
k(Ri)i2
n (Ri;eσ)i2
aks(k)
for n = 2, 4. These equations make use of the fact that all
the fk are self-consistent models for the same j(r), so that
all surface density factors µk(Ri) = µ(Ri) cancel.
To determine the weights wσ and wn, we have done
Monte Carlo simulations studying the propagation of the
observational errors ∆σ and ∆h4. Based on the results of
these simulations, we have chosen
w−1
w−1
w−1
σ (Ri) = 2.4σ(Ri)[σ(Ri)/σ(Ri)]∆σ(Ri),
2 (Ri) = 0.7∆σ(Ri)/σ(Ri),
4 (Ri) = 0.95α∆h4.
(11)
(10)
(12)
The coefficients are representative in the range of values
taken by the observed error bars and the measured h4. In the
presently used fitting procedure, the additional dependence
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
Figure 3. Anisotropy parameter β for a subset of basis functions
fk used in the self -- consistent Jaffe potential (top) and in a mixed
Jaffe plus halo model (bottom). The top full line in each panel
shows the isotropic model. All other basis models are tangentially
anisotropic. Two values of asymptotic anisotropy as r → ∞ are
used (full and dashed lines). For illustration, the long -- dashed lines
show the β -- profiles of the best -- fitting dfs derived with these
bases from the NGC 6703 data, in both potentials.
radially anisotropic df from the isotropic model plus a set
of tangential basis models.
In Fig. 3 we illustrate the basis models we have used,
by plotting radial profiles of the anisotropy parameter β(r)
for a subset of them. For these basis functions β(r) can be
regarded as a measure of the extent of the df in circularity
x, at the energy E = Φ(r). The figure shows that the basis
resolves three steps in x in the limits r → 0 and r → ∞;
at intermediate energies, it has much finer resolution. The
majority of the number of functions is spent on resolving
the energy dependence, i.e., on placing the main gradient
zones of the functions fk in the (E, x) plane on a relatively
dense grid in energy. The main difference to using power law
components (Fricke 1952, and, e.g., Dejonghe et al. 1996) is
that our basis models are already reasonable in the sense
that they are viable dynamical models for the density in
question, and that the superposition is used only to match
< 20 such func-
tions; this proved sufficient even for analysing pseudo data
of much better quality than we have for NGC 6703.
the kinematics. We have typically used K ∼
Each of the fk(E, x) reproduces the stellar density dis-
tribution j(r); so the ak satisfy
Moreover, the coefficients ak must take values such that
ak = 1.
Xk=1,K
Xk=1,K
akfk(E, x) ≥ 0
(6)
(7)
everywhere in phase-space. In practice, these positivity con-
straints are imposed on a grid in energy and circularity
on h4 and the respective "other" error bar is neglected -- the
tests below show that this leads to satisfactory results. Since
to first order the s2 -- moment thus measures the shift from
σ to σ, the parameter α in eq. (13) will be near α ≃ 2. We
have fixed α by requiring that the distributions of χ2
σ and
χ2
h4 have equal width (see below); this results in α = 1.7.
Finally, we assume that the df underlying the observed
kinematics is smooth to the degree that is compatible with
the measured data. Clearly, unless such an assumption is
made, it is impossible to determine a function of two vari-
ables, f (E, L2), from a small number of data points with
real error bars. One way to ensure that the df is smooth is
to use only a small number of terms in the expansion (5).
However, this is not a good way of smoothing as it biases
the recovered df towards the functional forms of the few
fk that are used in the sum. A better way of smoothing is
the method of regularization, as recently discussed by Mer-
ritt (1993) in a similar context. Regularization has tradition
in other branches of science, and different variants exist; for
references see Press et al. (1986) and Merritt's paper (1993).
In the algorithm here, regularization is implemented by tak-
ing the number of basis functions large enough (typically,
16 − 20) that the data can be modelled in some detail, and
then constraining the second derivatives of the composite
df to be small. That is we also seek to minimize
Λ(f )ij ≡ w2
∂2f
r (E) ×
∂2f
∂E∂x(cid:17)2
∂x2(cid:17)2(cid:21)E=Ei,x=xj
+(cid:16) ∂2f
on a grid of points (Ei, xj) in the energy and circularity in-
tegrals. Here we normalize to the isotropic df to ensure that
fluctuations in the composite df are penalized equally at all
energies; i.e., wr(E) = 1/fiso(E). The constant DE is pro-
portional to the range in potential energy in the respective
model.
To find a regularized spherical df for given kinematic
data in a specified potential, we thus minimize the quan-
tity
∆2 ≡
IXi=1(χ2
σ,i + Xn=2,4
χ2
n,i) + λXi,j
Λ(f )ij
(14)
for given regularization parameter λ, subject to the equality
and inequality constraints (6) and (7). For the actual nu-
merical solution we have used the NETLIB routine LSEI by
Hanson and Haskell (1981). Once the best model is found,
we redetermine its quality by evaluating its deviations from
the actually measured data:
(cid:20)(cid:16)D2
E
∂E2(cid:17)2
+ 2(cid:16)DE
(13)
χ2
σ = I −1
χ2
h4 = I −1
IXi=1hσ(Ri) − σ(f )(Ri)i2
4 (Ri)i2
IXi=1hh4(Ri) − h(f )
/(∆σ)2(Ri),
(15)
/(∆h4)2(Ri).
(16)
The parameters σ(f )(Ri) and h(f )
4 (Ri) are determined by
fitting a Gauss -- Hermite series to the velocity profiles of the
best fitting model; I is the number of kinematic data points.
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
The dark halo of the E0 galaxy NGC 6703
7
4.2 Tests with model data
We have tested this method by applying it to kinematic
datasets generated in the following Monte Carlo -- like way.
First, velocity dispersions and h4-parameters are calculated
from a theoretical df of specified anisotropy in a known
potential, and are interpolated to the radii Ri/rJ at which
observed data points are available for NGC 6703. Error bars
at these Ri are taken to be either the measured error bars
for NGC 6703 (a realistic dataset), or are taken to be inde-
pendent of radius with ∆σ = 3 km s−1 and ∆h4 = 0.01 (an
idealized dataset). Then pseudo data are generated from the
model values of σ and h4 at Ri by adding Gaussian random
variates with 1-σ dispersion corresponding to the respec-
tive ∆σ or ∆h4 error bar at this point. Figs. 4, 5 show
datasets generated in this way from a radially anisotropic
model in a potential of a self-consistent Jaffe sphere with
rJ = 46.′′5, GMJ /rJ = 272 km s−1, and a dark halo with
v0 = 220 km s−1, rc = 56′′. This potential was chosen be-
cause it lies in the middle of the range of acceptable poten-
tials for NGC 6703 (see Sect. 5.2), so that any systematic
errors in our analysis will be similar for this model and for
the galaxy itself.
We have used the regularized inversion algorithm de-
scribed in the last subsection to analyse several such pseudo
data sets, and have determined composite dfs as a func-
tion of the regularization parameter λ. The algorithm was
given 16 basis function models. These included the isotropic
model and a variety of tangentially anisotropic models with
different anisotropy radii and circularity functions but not,
of course, the radially anisotropic model from which the data
are drawn (see Fig. 3 and Appendix A). For each composite
model returned by the algorithm we have determined two
diagnostic quantities. The first is the mean χ2 per σ and h4
data point, χ2
σ+h4, which measures the level at which this
model fits the data from which it was derived. The second
is the rms deviation between the returned df and the true
df of the model from which the data were drawn, in some
specified energy range.
The use of these diagnostic quantities requires some fur-
ther comments. As usual, the number of degrees of free-
dom in such a regularized inversion problem is not well --
determined. For near-zero λ, in the case at hand we can
adjust 16 coefficients ak and an overall mass scale. The to-
tal number of σ and h4 -- data points is 70; thus in this limit
the number of degrees of freedom is 53. For large λ, on the
other hand, the recovered df will be linear in both E and
x and the values of the ak are essentially fixed. Then the
number of degrees of freedom approaches 69. In both cases,
it is of the order of the number of data points. Hence the
use of χ2
σ+h4 per σ and h4 data point instead of a reduced
χ2.
The second diagnostic measuring the accuracy of the re-
covered df must clearly depend on the range in energy over
which it is calculated. Typically, a kinematic measurement
at projected radius Ri contains information about the df
in a range of energy above the energy of the circular orbit
at radius Ri. The precise upper end of this range is model --
specific; it depends on the potential, the kinematic proper-
ties of the df itself, and, through the projection process,
also on the stellar density profile. The use of the outermost
kinematic data points thus contains, explicitly or implic-
8 O.E. Gerhard, G. Jeske, R.P. Saglia, R. Bender
0.16
0.12
0.08
0.04
0
-0.04
-0.08
-0.12
-0.16
0.6
0.4
0.2
0
-0.2
0
10 20
40 50
30
R [arcsec]
60 70
80
0
10 20
40 50
30
r [arcsec]
60 70
80
0.16
0.12
0.08
0.04
0
-0.04
-0.08
-0.12
-0.16
0.6
0.4
0.2
0
-0.2
0
10 20
40 50
30
R [arcsec]
60 70
80
0
10 20
40 50
30
r [arcsec]
60 70
80
Figure 4. Model analysis of pseudo-data generated from a radi-
ally anisotropic model df with the sampling and the measured
error bars of NGC 6703 (see text). The full curves show the true
profiles of projected velocity dispersion, line-of-sight velocity dis-
tribution parameter h4, and anisotropy parameter β of the un-
derlying df. The dashed and dotted lines show the σ, h4, and β
profiles of five regularized composite dfs which were computed
by the method of Sect. 4.1 with λ = 10−4. One of the four curves
in each panel corresponds to the data points actually shown in
this figure; the other three curves derive from statistically identi-
cal kinematic data sets with randomly different values for σ and
h4 within the same (Gaussian) errors.
itly, assumptions on the radial smoothness of these quan-
tities. In a typical elliptical galaxy problem, the df must
be known fairly accurately at around 3Ri for the projected
kinematics at Ri to be securely predicted, and for radially
anisotropic models, even the values of the df near 10Ri can
make some difference. The rms residuals in the recovered
df given below have therefore been calculated in a range of
energies extending from Φ(r = 0) to Φ(r = 3Rm), where the
last kinematic data point in the data sets used is located at
Rm = 1.68rJ .
Figure 6 shows these quantities as a function of the regu-
larization parameter λ for the two pseudo data sets shown in
Figs. 4, 5. For small values of λ, the composite models fit the
Figure 5. Model analysis of pseudo-data generated from the
same radially anisotropic model df as in Fig. 4 but assuming
radially constant error bars ∆σ = 3 km s−1 and ∆h4 = 0.01. The
full curves again show the true σ, h4, and β profiles of the under-
lying df. The dashed lines show the projected σ and h4 profiles
derived by the method of Sect. 4.1 from the pseudo-data shown,
with λ = 6×10−5. The results obtained for these projected quanti-
ties from statistically identical data sets are now nearly identical.
The dotted curves show the uncertainties that remain in the de-
projected quantity β even with such small error bars.
data accurately, but the recovered df is not very accurate
because it contains large spurious oscillations depending on
the particular values of the data points. For large values of
λ, the models are so heavily smoothed that they neither fit
the data well, nor do they represent a good approximation
to the true df. The optimal regime is where the smooth-
ing is large enough to damp out the spurious oscillations,
but still permits resolving the important structures in the
underlying df. For values of λ in this regime the fit to the
data is still satisfactory, and the representation of the df is
optimal. Fig. 6 shows that the rms residuals of the df go
through minima at values of λ ≃ 2 × 10−4 for the pseudo
data in Fig. 4 and λ ≃ 5 × 10−5 for those in Fig. 5 [when
DE = 1; see below eq. (13)].
σ+h4(λ) curves were found to be al-
The shapes of the χ2
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
The dark halo of the E0 galaxy NGC 6703
9
Figure 6. Model results as a function of the regularization pa-
rameter λ, for the two pseudo-data sets shown in Figs. 4 and 5.
The two data sets are flagged by triangles (Fig. 4) and pentagons
(Fig. 5). The upper two curves marked by the open symbols show
the total χ2 per σ and h4 data point of regularized composite
models with 16 basis functions. The lower two curves marked by
the filled symbols show the rms deviation between the recovered
composite model dfs and the true df from which the data sets
were generated. This rms deviation was evaluated on a grid in
energy and angular momentum corresponding to radii < 3 times
the radius of the outermost data point.
ways similar to the upper curves in Fig. 6. The shapes of the
corresponding lower curves in Fig. 6 are more variable. The
resulting optimal values for λ can vary, depending on the
random realisation of the data within the assumed Gaussian
errors, as well as on the distribution function and potential
from which the model values are drawn. We have therefore
investigated 100 realisations of data generated from each of
several model dfs and potentials. Based on these experi-
ments we have fixed optimal values of λ = 1 × 10−4 and
λ = 6 × 10−5 for data with error bars like those in Figs. 4
and 5, respectively, for all models that include a dark halo
component.
We have done similar experiments with a self-consistent
model that is radially anisotropic near the center and tan-
gentially anisotropic in its outer parts, such as might be
relevant in tests for dark matter at large radii. In these tests
we have used 20 basis functions. To match the corresponding
pseudo-data with similar accuracy requires smaller optimal
values of λ (∼ 3 × 10−9), so as to compensate for the larger
derivatives in eq. (13).
σ+h4 and
of the rms residual in the df, as described, for the dynami-
cal models recovered from several such sets of Monte Carlo
data. From the top panel it is seen that our fitting algo-
rithm will match kinematic data in a known potential with
σ+h4 ≤ 1 about 60−70% of the time, and with χ2
χ2
σ+h4 > 1.28
only < 5% of the time. These numbers are similar to those
expected from a χ2 -- distribution with 70 degrees of freedom
Fig. 7 shows the cumulative distributions of χ2
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
Figure 7. Top: The cumulative distribution of the total normal-
ized χ2
σ+h4, for dynamical models recovered from 100 random
Gaussian data sets derived from the underlying true df and the
observational errors in Figs. 4 (solid line) and 5 (dashed). Bot-
tom: Cumulative distribution of residuals between true and re-
covered distribution function, evaluated on a grid extending to
three times the radius of the last data point, for the same 100
datasets from both models. Also shown in the top panel are the
cumulative χ2
σ+h4 -- distributions for data drawn from radially an-
isotropic models in the two potentials corresponding to two of
the extreme solid lines in Fig. 11, with the same error bars as
in Fig. 4 (dot -- dashed lines), and for a self -- consistent model with
more complicated anisotropy structure (dotted line).
(for Gaussian data), which should describe the statistical
deviations of the Monte Carlo data points from the under-
lying true df. If in modelling the data for NGC 6703 a level
of χ2
σ+h4 < 1.28 cannot be reached, the assumed potential is
not correct with 95% confidence.
The bottom panel of Fig. 7 shows the distribution of
residuals in the recovered df for one model. There are three
factors which limit the degree to which a given underlying
df can be recovered. The first is determined by the data,
i.e., the size of the error bars, the sampling, the fact that
there are no measurements at large radii, etc. The second
is the level of detail that can be resolved by the modelling,
given the finite number and the particular form of the basis
functions. The third is the amount of small -- scale gradients
in the model itself. Figs. 6 and 7 show that, if the potential is
known, data like those for NGC 6703 (Fig. 4) allow recover-
< 12% inside
ing reasonably smooth dfs to an rms level of ∼
3Rm about 60 − 70% of the time. Data with much smaller
error bars but the same sampling (Fig. 5) would give an rms
< 8% for the other two halo models shown
level of ∼
< 10% (∼
10 O.E. Gerhard, G. Jeske, R.P. Saglia, R. Bender
in the top panel of Fig. 7). In the radial -- tangential self --
consistent model the df to 3Rm is recovered with an rms
accuracy of 16% and 35% for data with error bars like those
in Figs. 5 and 4, respectively. These comparisons and the
results of Fig. 7 show that the former values are dominated
by the measurement uncertainties rather than the resolu-
tion in the modelling. This conclusion would be different for
highly corrugated true dfs. However, we would not be able
to recover such dfs from realistic data in any case.
The kinematics of the regularized composite dfs de-
rived from pseudo-data with the chosen optimal λ values
are shown in Figs. 4 and 5. (For brevity, such model dfs
obtained with near -- optimal λ will henceforth be denoted as
"best -- estimate models".) The good match to the data points
is apparent. The differences in the intrinsic anisotropy pa-
rameter β between the recovered models and the true model
are larger. For the data set generated with the observed er-
ror bars of NGC 6703 the recovered β values are uncertain
by ∆β ≃ ±0.1, and slightly more at the largest radii where
the observational errors are large. For the pseudo-data with
the small error bars in Fig. 5, this uncertainty is reduced.
4.3 Constraining Φ
So far we have shown how the stellar df can be recovered
from vp -- data in a known spherical potential. In this Section
we return to the discussion of Section 3 and investigate the
degree to which the gravitational potential itself can be con-
strained. Eventually, it will clearly be important to answer
the theoretical question of whether, in principle, the gravita-
tional potential is nearly or even uniquely determined from
the projected df N (r, vk). However, we shall not attempt
to do this here; previous theoretical work suggests that the
range of potentials consistent with ideal data is small (G93,
Merritt 1993, see Section 3). More relevant at the moment
perhaps is the more practical question of how well the po-
tential can be constrained from real observational data with
realistic error bars, finite radial extent, and limited sam-
pling. What is the range of pairs (f, Φ) that correspond to
the same data? How does this range shrink as the data im-
prove?
Here we investigate these questions with a view to the
analysis of the NGC 6703 data below. We again use the
model underlying Figs. 4, 5; this is a radially anisotropic df
in the potential of a self -- consistent Jaffe sphere and a dark
halo with parameters GMJ /rJ = 272 km s−1, rc/rJ = 1.2
and v2
0 /(GMJ /rJ ) = 0.808. As before, random Gaussian
data sets were generated from this model, with the positions
and error bars of the data points (i) as in Fig. 4, (ii) as in
Fig. 5. The last data point is at Rm = 1.68rJ , as for the
NGC 6703 data, well beyond two effective radii. However,
contrary to Section 4.2, the data sets used in the following
were specially selected such that χ2
h4 ≃ 1. (It
turns out that the data points shown in Fig. 4 have less than
5% probability according to Fig. 7.).
σ ≃ 1 and χ2
For these pseudo data we have determined best --
estimate dfs in a number of assumed potentials, includ-
ing the underlying true potential. The potentials were cho-
sen such that (i) they correspond to approximately constant
mass -- to -- light ratio for r ≪ rJ , and (ii) they form a sequence
of varying true circular velocity vc(Rm) at the radius of the
last data point. The sequence, with the slightly different ve-
Figure 8. Goodness of fit χ2
σ+h4 for best -- estimate dfs derived
in a sequence of luminous plus dark matter potentials. The se-
quence of potentials, with rotation curves similar to those shown
in Fig. 11, is here parametrized by the total circular velocity at
the last observational radius, vc(Rm). The dfs were fitted to
pseudo data generated from a model with vc(Rm) = 242 km s−1.
Solid pentagons: data points with error bars as in Fig. 4. Open
pentagons: data points with error bars as in Fig. 5.
Fig. 8 shows the goodness -- of -- fit χ2
locity normalisations appropriate for NGC 6703, is shown in
Fig. 11. For each potential, specified by the selected values
of the parameters rc/rJ and v2
0 /(GMJ /rJ ), we determined
the goodness -- of -- fit χ2
σ+h4 as a function of the velocity scale
GMJ /rJ from the corresponding best -- estimate models. Fi-
nally, we computed the χ2
σ+h4 of the best -- estimate df for
the optimum velocity scale. This optimal velocity scale usu-
ally turns out slightly different for the two pseudo data sets.
σ+h4 of the opti-
mal sequence of best -- estimate models as a function of
vc(Rm), as determined for both model datasets. The un-
derlying true potential has vc(Rm) = 242 km s−1 with the
adopted parameters. Fig. 8 shows that only potentials with
vc(Rm) < 230 km s−1 can be ruled out (have less than 5%
probability according to Fig. 7) from the data with "re-
alistic" error bars. This conclusion is not surprising given
how large these error bars are. However, the surprise is that
also with "idealized" data there remains a range of poten-
tials with larger vc(Rm) than the true value, in which these
data can be fit no worse than in the true potential: formally,
235 km s−1 < vc(Rm) < 285 km s−1 with 95% confidence.
Fig. 9 shows the predicted kinematics and the intrinsic
anisotropy of the four models in Fig. 8 that match the "ideal-
ized" data set with χ2
σ+h4 ≃ 1. The best -- estimate dfs in the
potentials with higher vc(Rm) than that of the true potential
achieve their good fit to the data by the following means:
Inside rJ they compensate the potential's higher circular
speed by a larger radial anisotropy, which leads to slightly
larger h4 -- values (cf. Section 3). This effect is too small to
be detected even with the small error bars. Outside rJ , they
compensate by a radially increasing tangential anisotropy. In
this way, the velocity dispersions near the edge of the model
can be lowered, because the number of high -- energy orbits
coming in from outside is reduced. Such orbits would con-
tribute large line -- of -- sight velocities near their pericentres.
The h4 values in the region concerned are also only barely
affected, because the extra tangential orbits near radius r
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
The dark halo of the E0 galaxy NGC 6703
11
To test this, we have constructed a truly idealized data set
of two times seventy data points with "idealized" small error
bars as in Fig. 5, evenly spaced in radius, and extending to
6rJ . A model with potential differing only in the halo core
radius (66% of the true value) was found to fit even these
data with χ2
σ+h4 = 0.97, while for a model with different
halo core radius and different asymptotic circular velocity
(by 30 km s−1) a satisfactory fit could not be found.
We draw the following conclusions from these experi-
ments:
(i) Velocity profile data with presently achievable er-
ror bars contain useful information on the gravitational po-
tentials of elliptical galaxies. In particular, constant -- M/L
models are relatively easy to rule out once the data extend
beyond 2Re. The examples that we have studied in detail,
tuned to the NGC 6703 data, certainly belong to the less
favourable cases, because the dispersion profile is falling.
(ii) The detailed form of the true circular velocity curve
is much harder to determine. Conspiracies in the df are
possible that minimize the measurable changes in the line
profile parameters. A good way to parametrize the results is
in terms of the circular velocity at the radius of the outer-
most data point. With presently available data, this can be
determined to a precision of about ±50 km s−1.
(iii) Better results can be expected from higher -- quality
data, of the sort one could expect from the new class of 10m
telescopes. However, even with such data some uncertainty
on the detailed circular velocity curve will remain -- regard-
less of whether or not in theory the potential is uniquely
determined from the projected df N (r, vk). Therefore, the
combination of the type of analysis presented here with other
information (e.g., from X-ray data) will give the most pow-
erful results.
5 THE ANISOTROPY AND MASS
DISTRIBUTION OF NGC 6703
5.1 Constant -- M/L model fits
The sb-profile of NGC 6703 is well fitted by a Jaffe mo-
< 15% around R ≃ 25′′,
del. The largest local residuals are ∼
< 10% around R ≃ 40′′, and smaller elsewhere, in particular
∼
at R > 60′′. A non -- parametric inversion of the the surface
brightness profile showed that the deviations from a Jaffe
density law are not significantly larger than those quoted
in sb. In the curve of growth, measuring the total luminos-
ity inside R, the residuals are everywhere less than ∼ 2%
(Fig. 1). Since the potential Φ(r) is determined by the to-
tal mass M (< r), we can thus to good approximation use a
Jaffe model for the gravitational potential of the stars. This
enables us to also compare our kinematic data with a large
set of self-consistent dynamical models from Jeske (1995).
We have fitted all models from this database to the
observational data for NGC 6703, taking the fitted Jaffe
radius rJ = 46.′′5. All data points from Fig. 1 were included
and weighted equally with their individual error bars, except
for two adjustments. (i) The error bars of the three h4-points
near 25" have been set equal to their standard deviation, and
(ii) the error bars of the innermost eleven h4-points have
been set to twice the measured values. These modifications
prevent the total χ2
h4 to be dominated by these points which
0.16
0.12
0.08
0.04
0
-0.04
-0.08
-0.12
-0.16
0.6
0.4
0.2
0
-0.2
0 10 20 30 40 50 60 70 80 90 100
R [arcsec]
0 10 20 30 40 50 60 70 80 90 100
r [arcsec]
Figure 9. Model fits to pseudo velocity dispersion and h4 data
with "idealized" error bars (see text, and Fig. 5). The full line
shows the kinematics of the model from which these pseudo data
were generated. The dashed lines show the kinematics of best --
estimate dfs in a sequence of potentials with increasing halo
circular velocity, all of which fit the data perfectly with χ2
σ+h4 ≃
1. The intrinsic anisotropy of these models (dotted lines) and of
the true model are shown in the bottom panel.
and the lack of higher energy radial orbits from beyond r
nearly compensate.
Clearly, this mechanism will work less well both when
the difference between the true and attempted circular ve-
locity curves increases, and when the radial extent, sam-
pling, and quality of the data points improve. With data
yet better and more extended than those in Fig. 5, some
of the potentials consistent with the presently used model
data could probably be ruled out. Thus it appears that, if
the asymptotic circular velocity is constant, then the value of
that constant can be determined accurately with sufficiently
high quality data.
On the other hand, the potentials we have used are
very simply parametrized functions. There might well exist
more complicated potential functions, whose circular veloc-
ity curves differ in only a restricted range of radii, that would
be impossible to distinguish even with extremely good data.
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
12 O.E. Gerhard, G. Jeske, R.P. Saglia, R. Bender
NGC 6703 Model Fits
5
5
5
5
5
5
5
5
5
5
0
0
0
0
0
0
0
0
0
0
0
5
10
15
-χ2
Figure 10. χ2
σ-diagram from fitting dynamical models to
h4
the kinematic data for NGC 6703. The crosses show fit results for
a variety of anisotropic, self -- consistent models of a Jaffe sphere,
whose density profile is a good approximation for the luminous
matter in NGC 6703. They fall upwards and to the right of a
curved envelope that separates them clearly from a perfect fit,
showing that no self -- consistent model can simultaneously fit both
the dispersion and line profile shape data. The heavy dot is the
best -- estimate self -- consistent model in the stars -- only potential,
obtained with the method of Section 4. The squares show a num-
ber of dynamical models with a dark halo; these are consistent
with the data.
are of no consequence for the halo of NGC 6703. Moreover,
systematic effects may play a role for the innermost data
points (Sect. 2).
h4 -χ2
The velocity scale of each model is matched to opti-
mize the fit to the σ-profile; that to the h4-profile is already
fixed with no extra assumption. The resulting values of χ2
h4 ,
χ2
σ are normalized by the number of fitted data points. In
Fig. 10 we plot a χ2
σ-diagram for all self-consistent mo-
dels from Jeske (1995). The χ2 values are again normalized
to the number of data points. Also plotted on the figure are
best-estimate models in a number of potentials, constructed
with the technique described in Section 4. The optimized
self-consistent model (the heavy dot) lies on the bound-
ing envelope of the self-consistent models; it has χ2
σ = 2.6,
χ2
h4 = 1.3. The squares show the normalized χ2-values for
several models with a dark halo whose rotation curves are
shown by the full lines in Fig. 9. For most of these χ2
σ ≃ 1.3,
χ2
h4 ≃ 0.8.
Fig. 10 shows that no self-consistent model will fit the
data: all self-consistent models are clearly separated by a
curved envelope from the lower left hand corner of the dia-
gram, which corresponds to a perfect fit. Either the velocity
dispersion profile is matched reasonably well, but then the
line profiles cannot be reproduced, or, when the h4-profile
is fitted accurately, the dispersion profile is poorly matched.
The cure for the discrepancy is to raise both σ and h4 at
large radii. Thus, according to Sect. 3 above, we require ex-
tra mass at large r. NGC 6703 must have a dark halo.
5.2 Dynamical models with dark halo
We will now derive constraints on the gravitational poten-
tial of NGC 6703 within the framework of the parametric
mass model of eqs. (2) -- (4). In doing this we have in mind
Figure 11. Rotation curves for a sequence of gravitational po-
tentials (stars plus dark halo) used in the analysis of NGC 6703.
The full lines show rotation curves that are consistent with the
NGC 6703 kinematic data inside the 95% confidence boundary at
λ = 10−5 (open symbols in Fig. 12). The other line styles show
rotation curves inconsistent with the data; among these is the
constant -- M/L model with no dark halo (short -- dashed).
the following working picture: The stellar component is as-
signed a constant mass-to-light ratio Υ, chosen maximally
such that the stars contribute as much mass in the center
as is consistent with the kinematic data. The model for the
halo incorporates a constant density core, and its parame-
ters are chosen such the halo adds mass mainly in the outer
parts of the galaxy if that is necessary. This is similar to
the maximum disk hypothesis in the analysis of disk galaxy
rotation curves. Within this framework we can determine
the maximum stellar mass-to-light ratio, ask whether it is
reasonable, and constrain the halo parameters.
Because determining the two potential parameters rc
and v0 together with the model velocity scale (equivalent
to the stellar mass -- to -- light ratio) is a three -- dimensional
problem, we will proceed in steps. Fig. 11 shows circular
rotation curves for the first sequence of mass models for
NGC 6703 that we have investigated. In all these models the
< Re.
halo contribution becomes significant only outside 20′′
∼
The corresponding halo core radii mostly lie between 1.2rJ
and 1.7rJ , i.e., 55′′ and 80′′. These values are relatively large
because of the falling dispersion curve in NGC 6703. This
implies that we can reliably determine only one of the halo
parameters for this galaxy.
We have chosen the circular velocity vc(Rm) at the ra-
dius Rm of the last kinematic data point as this parameter.
The sequence in Fig. 11 was constructed such as to vary
vc(Rm) and the rotation curve outside ∼ Re, while leaving
the central rotation curve nearly invariant. However, when
the model velocity scales are optimized in the determina-
tion of the best -- estimate dfs, the optimal velocity scale
is found to correlate inversely with vc(Rm). The rotation
curves in Fig. 11 are plotted with their optimal scaling; then
the central rotation curves are no longer identical, but be-
come scaled versions of each other.
In each of the corresponding potentials, we have con-
structed the best -- estimate df with optimal velocity scaling,
as described in Section 4. We first fit a composite df to the
velocity dispersion and line profile shape data for a series of
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
The dark halo of the E0 galaxy NGC 6703
13
Figure 13. The (v0, rc) halo parameter plane. Values of rc are
scaled with respect to rJ = 46′′.5. The luminous plus dark matter
models investigated are shown as plus signs. The contours show
lines of constant χ2
σ+h4 obtained by interpolating between the
model values. Acceptable potentials lie in a band extending from
low v0 and low rc to high v0 and high rc. Models at the upper
left are ruled out because they do not contain enough mass at
large radii. Models at the lower right are ruled out because no
satisfactory fit can be found for any constant value of the stellar
mass -- to -- light ratio.
The fitting results in these potentials are shown by the iso-
lated filled symbols in Fig. 12. Combining with the previous
results these allow us to investigate the range of accept-
able potentials for NGC 6703 more fully than with the one-
dimensional sequence in Fig. 11.
Fig. 13 shows contours of constant χ2
σ+h4 in the (v0, rc) --
plane. The most probable potentials lie in a band extending
from low v0 and low rc to high v0 and high rc. Thus, as al-
ready discussed above, it is not possible to determine both
halo parameters in the NGC 6703 case. However, potentials
in the band of most probable (v0, rc) all have circular veloc-
ities vc(Rm) in the same range of 250 ± 40 km s−1 as before.
In fact, the best -- fitting velocity scales of all models in
Fig. 13 turn out such that the resulting values of vc(Rm) are
in the range [189, 318] km s−1. The fitting procedure tends
to move vc(Rm) into the correct range even when no satis-
factory df can be found. E.g., some models in the lower
right of Fig. 13 with χ2
σ+h4 ≃ 1.5 appear in Fig. 12 at
> 2.4 (not
vc(Rm) = 250 − 300 km s−1, some with χ2
shown) at vc(Rm) = 280 − 340 km s−1. All models shown in
the upper left of Fig. 13 fall near the line defined by the se-
quence of models discussed at the beginning of this section.
The fact that vc(Rm) varies relatively little for a wide range
of luminous matter plus dark halo models suggests that our
results, when expressed in terms of this parameter, are not
sensitive to the choice of halo model in eqs. (3), (4).
σ+h4 ∼
From Fig. 12 we conclude that the true circular ve-
locity of NGC 6703 at 78" is vc(Rm) = 250 ± 40 km s−1
(the formal 95% confidence interval obtained from the filled
symbols, according to Fig. 7). In Section 4.3 we found,
however, that most of this indeterminacy is towards large
circular velocities, at least for a galaxy with a dispersion
curve like that of NGC 6703. By contrast, in potentials with
lower values of vc(Rm) than the true circular velocity, it
quickly becomes impossible to find a satisfactory df. Based
Figure 12. Quality with which the kinematics of NGC 6703
can be fitted in different potentials. The figure shows the av-
erage χ2 per σ -- and h4 data point, of the best -- estimate distri-
bution function fitted to the velocity dispersion and line profile
data, as a function of the assumed potential's circular rotation
velocity at the observed radius of the last kinematic data point.
Filled symbols show best -- estimate models derived with the opti-
mal λ = 10−4 determined in Section 4.2; open symbols represent
models derived with λ = 10−5. The self -- consistent (M/L = const.)
and the vc = const. models are marked separately. The horizon-
tal dashed line shows the 95% confidence boundary derived from
Fig. 7.
values of the unknown velocity scale. From this sequence of
models, we determine the optimal value of the model's ve-
locity scale in this potential, and then recompute the best --
estimate model with this velocity scale. In the following,
when speaking of the best -- estimate df for a given poten-
tial, we will always imply that the velocity scale has been
optimized in this way. In the fitting procedure we have used
a regularization parameter λ = 10−4; this was found appro-
priate in Section 4.2 for the error bars and sampling of the
NGC 6703 data. Compared with the tests in Section 4, we
have included a few extra basis functions (total K = 20) to
resolve the (possibly not real, cf. Section 2) high -- frequency
structure in the center of NGC 6703.
Fig. 12 shows, as a function of vc(Rm), the average χ2
per σ -- and h4 data point of the respective dfs so obtained.
The connected solid symbols represent the sequence of po-
tentials corresponding to Fig. 11. The potential with con-
stant mass -- to -- light ratio appears in the upper left -- hand
corner in Fig. 12; it is inconsistent with the data by a large
margin even for optimum velocity scale (see Fig. 7). The
best -- fitting potential with a completely flat rotation curve
has an optimal value of vc = vc(Rm) = 254 km s−1 and
χ2
σ+h4 = 1.17. Thus it does not provide the best possible fit
but is consistent with the data. Of the stars plus dark halo
models illustrated in Fig. 11, those models in the sequence
with vc(Rm) = 210 − 285 km s−1 have χ2
σ+h4 < 1.28. This
is consistent with the results of Section 4.3, from which we
would expect that the NGC 6703 data can be fit by a range
of gravitational potentials. Models in the sequence outside
this range of vc(Rm) are inconsistent with the kinematic
data at the 95% confidence level (cf. Fig. 7).
In a second step we have analyzed a more complete set
of potentials in a suitable part of the (v0, rc) -- plane (Fig. 13).
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
14 O.E. Gerhard, G. Jeske, R.P. Saglia, R. Bender
on these results, the lower values in the quoted range of
vc(Rm) = 250 ± 40 km s−1 appear to be the more probable
ones.
The open symbols in Fig. 12 show that the range of
potentials consistent with the data is enlarged only slightly
when the df is allowed to be less smooth. These points are
from best -- estimate models derived in the sequence of po-
tentials of Fig. 11 with λ = 10−5 instead of the optimal
λ = 10−4. The resulting curve which connects the λ = 10−5
models in Fig. 12 surrounds the corresponding λ = 10−4
curve. Generally it appeared that, in the models obtained
with λ = 10−5, the df came close to zero more easily and
more often. The last two facts, when taken together, sug-
gest that, in addition to the data themselves, the positivity
constraints on the df play an important role in determining
the boundary of the region in (f, Φ) space that is consistent
with given kinematic data.
Fig. 14 shows best -- estimate models in some of the
potentials consistent with the kinematic data (at λ =
10−4). The three full
lines are models with vc(Rm) =
(218, 231, 242) km s−1, the two dotted lines are models with
vc(Rm) = (253, 277) km s−1. Also shown is the best --
estimate model in the potential of only the stars with con-
stant M/L (long -- dashed), and that in the best potential
with vc = const. = 254 km s−1 (short -- dashed).
the best -- estimate
The dispersion profile of
self-
consistent model falls clearly below the data at both inter-
mediate and large radii, as does its h4-profile. From the dis-
cussion in Section 3, this is a clear sign of extra mass at large
radii. The models including dark halo contributions mainly
differ in the outermost parts of the velocity dispersion pro-
file. As expected, those with the highest velocity dispersions
at large radii correspond to the potentials with the largest
asymptotic circular velocities. This again suggests that with
smaller error bars at large radii and with spatially more ex-
tended data, we will be able to significantly narrow down the
uncertainties in the halo parameters. The model with every-
where constant rotation speed is constrained tightly by the
kinematic data in the central parts, where the ρ ∝∼ r−2 pro-
file is presumably dominated by the stars. It then has some
difficulties both with the h4 -- values at intermediate radii and
the velocity dispersions at large radii.
The lower part of Fig. 14 shows that, in order to match
the observed kinematics in a potential with large circular ve-
locity, the df must become rapidly tangentially anisotropic
at the radii of the last data points and beyond. As discussed
in Section 4, this is different from the better known effect of
increasing the projected dispersions in a potential with not
enough mass at large radii, by making the df more tangen-
tially anisotropic. Here, without the tangential anisotropy,
our models would predict too high values of the velocity
dispersion, because too much mass at large radii is implied.
The tangential anisotropy at radii outside those reached by
the observations reduces the number of orbits that come into
the observed range, orbits which would contribute relatively
large line-of-sight velocities near their turning points. In this
way the velocity dispersions can be reduced down to the ob-
served values. Clearly, more spatially extended data would
reduce this freedom.
From Fig. 14 we conclude that the stellar distribution
function in NGC 6703 is near-isotropic at the centre and
then changes to slightly radially anisotropic at intermedi-
200
160
120
0.16
0.12
0.08
0.04
0
-0.04
-0.08
-0.12
-0.16
0.6
0.4
0.2
0
-0.2
0 10 20 30 40 50 60 70 80 90 100
R [arcsec]
0 10 20 30 40 50 60 70 80 90 100
r [arcsec]
Figure 14. Dynamical models for the kinematics of NGC 6703
in several luminous plus dark matter potentials, compared to
projected velocity dispersion (top panel) and vp-shape param-
eter h4 (middle panel). The bottom panel shows the models' in-
trinsic anisotropy parameter β(r), with the same linestyles: self --
consistent model (stars only; long -- dashed), vc = const. model
(short -- dashed), three models with vc(78′′) in the lower part of
the acceptable range (full), and two models with vc(78′′) in the
upper part of this range (dotted lines).
ate radii (β = 0.3 − 0.4 at 30", β = 0.2 − 0.4 at 60"). It is
not well-constrained near the outer edge of the data, where
formally β =−0.5 − +0.4, depending on the correct poten-
tial in the allowed range. However, the models with large
asymptotic halo circular velocities shown in Fig. 14 appear
less plausible, because they are the models with the most
rapidly increasing velocity dispersions outside R ≃ 60′′. The
same models also show the most rapid increase in tangen-
tial anisotropy at and beyond Rm = 78′′, which again ap-
pears a priori implausible, because it implies rapid changes
in the df just outside the observed range. The combined
signature of both effects is strongly reminiscent of Fig. 9,
where it was clearly an artifact of the limited radial range
of the data. If this assumption is correct, one would again
conclude that lower -- vc(Rm) models in the formal range are
favoured. Fig. 15 shows the recovered df for the potential
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
The dark halo of the E0 galaxy NGC 6703
15
Figure 16. Luminous, dark, and total mass as a function of ra-
dius for the range of acceptable models of NGC 6703, according
to Fig. 12 (short dashed, dotted, and dash -- dashed or full lines, re-
spectively). Mass distributions in which a df with χ2
σ+h4 ≤ 1.12
(including 87% of the distribution in Fig. 7, 1.5σ) are coded by full
lines, those with χ2
σ+h4 ≤ 1.28 (including 95%, 2.0σ) by dash --
dashed lines. The vertical line denotes the position of the last
kinematic data point. At this radius, the luminous mass fails by
at least a factor of 1.6.
Figure 17. The B-band mass -- to -- light ratio of NGC 6703. The
solid and dash -- dashed lines (coding as in Fig. 16) are derived from
the dynamical models that span the range of acceptable vc(Rm)
in Fig. 12. The central mass -- to -- light ratio is Υ = 3.3, that at the
position of the last kinematic data point at 78" (vertical bar) is
in the range of Υ = 5.3 − 10.
5.3 Uncertainties
There are a number of possible sources of systematic error
which would affect the mass -- to -- light ratio derived for NGC
6703. Most of the errors so introduced are probably small
compared to the considerable uncertainty arising from the
kinematic measurement errors and limited radial sampling,
discussed above. One systematic error on the absolute mass-
to-light ratios in NGC 6703 comes from the uncertainty in
the distance, although this does not change the ratio of outer
to central values. A further systematic effect on this ratio
can be introduced by the sky brightness level: If this is in-
creased by 2− 3%, the fitted rJ decreases to 35". To the ex-
tent that the outermost kinematic data point at 78" (which
then moves to larger r/rJ ) is in the flat part of the circu-
Figure 15. Distribution function in the energy -- circularity plane
derived for NGC 6703, in the luminous plus dark matter potential
with vc(Rm) = 231 km s−1 shown as the middle solid line in the
upper panel of Fig. 14. Energy is specified by the value of r(E),
where E = Φ(r), in units of the fitted Jaffe radius rJ , and the
df is given in arbitrary logarithmic units. The last measured
kinematic data point is located at lg r(E)/rJ = 0.23.
with vc(Rm) = 231 km s−1 in Fig. 14. Note, however, that
all models shown in Fig. 14 except the self -- consistent one
are formally consistent with the presently available data.
50 M⊙,
(17)
From the constraints on the circular velocity vc(Rm) =
50 kpc
210−290 km s−1 the range in mass inside Rm = 13.5 h−1
is
M (< Rm) = 1.6 − 2.6 × 1011h−1
where h50 ≡ H0/50km/s/Mpc. The total mass in stars in-
side this radius is 8− 9× 1010 M⊙, assuming constant mass-
to-light ratio Υ and a maximum stellar mass model, and
taking an average value from the models consistent with the
kinematic data. The radial run of the luminous, dark, and
total mass is shown in Fig. 16 for the models that span the
allowed range according to Fig. 12. After dividing by the lu-
minosity LB(r) for the stars, the mass-to-light ratios shown
in Fig. 17 result. Between the centre and the last data point
r = 78′′ ≃ 2.6Re, the mass-to-light ratio of NGC 6703 rises
by a factor of 1.6 − 3.
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
16 O.E. Gerhard, G. Jeske, R.P. Saglia, R. Bender
lar rotation curve, the inferred M/L changes only to second
order because the luminosity inside 78" remains essentially
unchanged. The same is true if the sky value is decreased by
1%, in which case the fitted rJ increases to 54′′.
In the previous analysis, we have ignored a possible
small rotation in the outer parts of NGC 6703 (perhaps
> 50′′, but the errors are large; cf.
∼ 20 − 30 km s−1 at R ∼
Fig. 1). The simplest possible estimate of the effect of this
rotation on the derived masses is to replace σ ≃ 105 km s−1
in this radial range by (σ2 + v2)1/2 ≃ 110 km s−1. This gives
a factor of 1.1, neglecting changes in the model structure
that would result because the central kinematics remain un-
changed.
Next we consider the possibility that NGC 6703 may
contain a face-on extended disk (see de Vaucouleurs, de Vau-
couleurs and Corwin 1976). From an R1/4-law plus disk de-
composition, we estimate that the contribution of such a
disk in the region where we model the kinematics could be
up to 10-20%. In this case we expect the velocity dispersion
to be decreased and the h4 coefficient to become more pos-
itive where the disk contributes significant light (Dehnen &
Gerhard 1994; see NGC 4660 as an example in BSG), most
likely in the outer parts.
Similarly, it is conceivable that NGC 6703 is in reality
slightly triaxial and is seen from a special direction so as to
appear E0. The likelihood of this is the smaller, the more
triaxial the intrinsic axial ratios are; thus slightly triaxial
shapes are the most plausible ones. Again this will imply
some extra loop orbits seen nearly face-on, similarly increa-
sing h4 and decreasing σ.
In both cases, we therefore expect the spherical com-
ponent in NGC 6703 to have lower h4 and larger σ than
the measured values. A similar analysis of such kinematics
would, according to the discussion in Section 3, lead to a mo-
del with greater tangential anisotropy at large radii, with
the mass distribution less affected. Recall that decreasing
the mass at large r in a spherical model lowers both σ and
h4.
6 DISCUSSION AND CONCLUSIONS
This study is part of an observational and theoretical pro-
gram aimed at understanding the mass distribution and or-
bital structure in elliptical galaxies. In the following we first
discuss general results on potential and anisotropy determi-
nation, and then proceed to the specific case of NGC 6703.
6.1 Velocity profiles and anisotropy and mass
The analysis of the vps of simple dynamical models in Sec-
tion 3 has broadly confirmed the conclusions of G93. At
large radii, where the luminosity profile falls rapidly, the vps
are dominated by the stars at tangent point. Then radially
(tangentially) anisotropic dfs can be recognized by more
peaked (more flat-topped) vps with more positive (nega-
tive) h4 than for the isotropic case. Increasing β at constant
potential thus lowers σ and increases h4. On the other hand,
increasing the mass of the system at large r at constant
anisotropy, increases both the projected dispersion and h4.
This suggests that by modelling σ and h4 both mass M (r)
and anisotropy β(r) can be constrained.
In practical applications, such an analysis is compli-
cated by a number of factors. Radial orbits at large radii
may lead to increased central velocity dispersions and flat --
topped central vps (already pointed out by Dejonghe 1987).
The former effect can be compensated by a decrease in the
stellar mass -- to -- light ratio. The latter is independent of this,
but can be compensated by changes to the distribution func-
tion in the inner parts of the galaxy (as in a number of cases
studied in Sections 4, 5). A more serious uncertainty is in-
troduced by the possibility of significant gradients in the
orbit population across the radii of interest. For example, a
population of high energy radial orbits with pericentres in a
limited radial range may mimic tangential anisotropy there.
In many cases it will be possible to exclude such a population
of orbits by its effects on the vps at exterior radii, i.e., by si-
multaneously analysing a number of observed vps. Yet this
is least possible precisely at the largest observed radii, where
mass determination is most interesting. Thus this chain of
argument suggests (correctly, see below) that the largest un-
certainty in determining masses and anisotropies in ellipti-
cals from vp -- data is the finite radial extent of these data.
To analyze realistic data we have constructed an algo-
rithm by which the distribution function and potential of
a spherical galaxy can be constrained directly from its ob-
served σ and h4 -- profiles. To assess the significance of the
results obtained, we have tested the algorithm on Monte
Carlo -- generated data sets tuned to the spatial extent, sam-
pling, and observational errors as measured for NGC 6703.
From such data, the present version of the algorithm recov-
ers a smooth spherical df, ∼ 70% of the time, to an rms
level of better than ∼ 12% inside three times the radius of
the outermost kinematic data point.
We have used this algorithm to study quantitatively
the degree to which the gravitational potential can be de-
termined from such data. Our main conclusion is that veloc-
ity profile data with presently achievable error bars already
constrain the gravitational potentials of elliptical galaxies
significantly. In particular, constant -- M/L models are rela-
tively easy to rule out once the data extend beyond 2Re.
The examples that we have studied in detail, tuned to the
NGC 6703 data, certainly belong to the less favourable cases,
because in this galaxy the dispersion profile is falling.
A good way to parametrize the results is in terms of the
true circular velocity vc(Rm) at the radius of the outermost
data point, Rm. With presently available data, vc(Rm) can
< 50 km s−1. This
be determined to a precision of about ± ∼
will improve when high -- quality data at several Re become
available, of the kind expected from the new class of 10m
telescopes. Apart from the fact that smaller error bars will
decrease the formally allowed range in vc(Rm), tests show
that this range often includes high -- vc(Rm) models which be-
come rapidly tangentially anisotropic just outside the data
boundary. These (not very plausible) models can be elimi-
nated with data extending to larger radii.
On the other hand, the detailed form of the true circu-
lar velocity curve is much harder to determine than vc(Rm).
Conspiracies in the df are possible that minimize the mea-
surable changes in the line profile parameters. Our tests
showed that two potentials differing by just the value of the
halo core radius could not be distinguished even with very
good data out to 6Re. Thus some uncertainty will remain in
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
practise, regardless of whether or not in theory the potential
is uniquely determined from the projected df N (r, vk).
A similar picture holds for the related determination
of the anisotropy of the df. For the present error bars in
the data, β(r) is relatively well -- determined out to about
half the limiting radius of the observations. Near the edge of
the data, uncertainties can be large depending on the grav-
itational potential (recall that in a fixed spherical potential,
the df is uniquely determined by the complete projected df
N (r, vk)). Again the unknown nature of the orbits beyond
the last data point has a large part in this uncertainty.
Because the largest uncertainties in determining masses
and anisotropies from vps occur near the outer radial limit
of these data, the combination of the type of analysis pre-
sented here with other information (e.g., from X-ray data)
will be particularly powerful.
6.2 The dark halo of NGC 6703
Fig. 10 shows that no self-consistent model will fit the
kinematic data for NGC 6703. Our non -- parametric best --
estimate self -- consistent model is inconsistent with the data
at the > 99% -- level (Figs. 12, 7). With self-consistent mo-
dels, either the velocity dispersion profile is matched reason-
ably well, but then the line profiles cannot be reproduced,
or, when the h4-profile is fitted accurately, the dispersion
profile is poorly matched. The cure for the discrepancy is to
raise both σ and h4 at large radii. Thus, as discussed above,
we require extra mass at large r. NGC 6703 must have a
dark halo.
We have next derived constraints on the parameters of
this halo as follows. The luminous component is assigned a
constant mass-to-light ratio Υ, chosen maximally such that
the stars contribute as much mass in the center as is consis-
tent with the kinematic data. Our parametric model for the
halo incorporates a constant density core, and its param-
eters (core radius rc and asymptotically constant circular
velocity v0) are chosen such the halo adds mass mainly in
the outer parts of the galaxy if that is necessary. We call
these models maximum stellar mass models (analogous to
the maximum disk hypothesis in the analysis of disk galaxy
rotation curves).
50 kpc is 9 × 1010h−1
We find that maximum stellar mass models in which
the luminous component provides nearly all the mass in the
centre fit the data well. In these models, the total luminous
mass inside the limiting observational radius Rm = 78′′ =
13.5 h−1
50 M⊙, corresponding to a central
B -- band mass -- to -- light ratio Υ = 3.3h50 M⊙/ L⊙. According
to Worthey's (1994) models, this is a rather low value for the
stellar population of an elliptical galaxy and would point to a
relatively low age (5 Gyrs) and/or low metallicity (less than
solar). However, the galaxy has a color (B − V )0 = 0.93 and
a central line index M g2 = 0.280 (Faber et al. 1989) which
are typical for ellipticals of similar velocity dispersion.
A larger value of H0 could increase the M/L value and
alleviate the demands on the stellar populations. However,
the distance used here (36 Mpc) includes a correction for
the large inferred peculiar velocity of the galaxy. If we had
used a distance based on the larger radial velocity in the
CMB frame, our derived M/L would be even lower. It is
also implausible that the low central value of M/L stems
from the contribution of a young stellar population in a disk
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
The dark halo of the E0 galaxy NGC 6703
17
component, which we estimate cannot be larger than 20% of
the total light (see above). Thus we conclude that the dark
halo in NGC 6703 is unlikely to have higher central densi-
ties than inferred from our maximum stellar mass models,
because otherwise the M/L of the stellar component would
be reduced to implausibly small values.
In a recent preprint, Rix et al. (1997) have analyzed
the velocity profiles of the E0 galaxy NGC 2434 with a
linear orbit superposition method. This galaxy provides an
interesting contrast to NGC 6703 because it has an essen-
tially flat dispersion profile. Its kinematics are likewise in-
consistent with a constant -- M/L potential, but are well -- fit
by a model with vc = const. This can be interpreted as a
maximum stellar mass model in the sense defined above, in
which the luminous component with maximal Υ contributes
most of the mass inside Re. The kinematics of NGC 2434
are also well -- fit by a range of specific, cosmologically moti-
vated mass models which, if applicable, would imply lower
Υ and significant dark mass inside Re. In NGC 6703, a mo-
del with vc = const is formally consistent with the present
data (within 2σ), but it is not a very plausible fit at large R
and requires large anisotropy gradients between 40" and 70".
It will be interesting to see whether future studies confirm
differences between the shapes of the true circular velocity
curves of elliptical galaxies.
Because of the falling dispersion curve in NGC 6703,
we can determine only one of the halo parameters (The halo
parameters of the most probable potentials lie in a band
extending from low v0 and low rc to high v0 and high rc.)
However, the circular velocity vc(Rm) at the data boundary
is relatively well -- determined for all these models. Thus we
find (Fig. 12) that the true circular velocity of NGC 6703 at
78" is vc(Rm) = 250± 40 km s−1 (formal 95% confidence in-
terval). Tests on pseudo data have shown that this range of-
ten includes high -- vc(Rm) models which become rapidly tan-
gentially anisotropic just outside the data boundary. Such
models may not be very plausible, so the lower values in the
quoted range of vc(Rm) = 250± 40 km s−1 may be the more
probable ones.
50 kpc the total mass en-
closed is M (< Rm) = 1.6 − 2.6 × 1011h−1
50 M⊙, and the inte-
grated mass -- to -- light ratio out to this radius is Υ = 5.3− 10,
corresponding to a rise from the center by at least a factor
of 1.6. We have already noted that NGC 6703 is an un-
favourable case because of its falling dispersion curve. The
fact that relatively small variations in Υ can nonetheless
be detected shows the power of the method. Note that a
scheme based on the analysis of the line of sight velocity
dispersions alone (Binney, Davies, Illingworth 1990, van der
Marel 1991) would conclude that constant mass -- to -- light ra-
tio models can provide good fits.
Thus, at Rm = 78′′ = 13.5 h−1
The stellar distribution function in NGC 6703 is near-
isotropic at the centre and then changes to slightly radi-
ally anisotropic at intermediate radii (β = 0.3 − 0.4 at 30",
β = 0.2−0.4 at 60"). It is not well-constrained near the outer
edge of the data, where formally β =−0.5− +0.4, depending
on the correct potential in the allowed range. Models near
the lower end of this range may be consistent with the data
only because of the limited radial extent of the measure-
ments.
18 O.E. Gerhard, G. Jeske, R.P. Saglia, R. Bender
6.3 Conclusions
In summary, we have shown that the mass distribution M (r)
and anisotropy structure β(r) for spherical galaxies can both
be constrained from vp and velocity dispersion measure-
ments. NGC 6703 must have a dark halo, contributing about
equal mass at 2.6Re as do the stars. The circular velocity at
the last kinematic data point (78") is constrained to lie in the
range 250 ± 40 km s−1 at 95% confidence. The anisotropy of
the stellar orbits changes from near-isotropic at the center to
slightly radially anisotropic at intermediate radii, and may
be either radially or tangentially anisotropic at 78". With
more extended and more accurate data it will be possible to
considerably narrow down these uncertainties.
If the results for this galaxy are typical, they suggest
that also in elliptical galaxies the stellar mass dominates at
small radii, and the dark matter begins to dominate at radii
around 10 kpc. It is important to obtain extended kinematic
data and do a similar analysis for a number of elliptical ga-
laxies. When we know the systematics and the spread in the
circular velocity curves and anisotropy profiles for a sam-
ple of ellipticals, we will have an important new means for
testing the currently popular formation theories.
ACKNOWLEDGMENTS
We thank U. Hopp for providing us with a CCD frame of
NGC 6703, and David Merritt for helpful discussions on reg-
ularization methods. We thank the referee for his rapid and
constructive comments, especially on the revised version,
and the editorial staff for their relentless efforts to secure his
reports. We acknowledge financial support by the Deutsche
Forschungsgemeinschaft under SFB 328 and SFB 375 and by
the Schweizerischer Nationalfonds under grants 21-40464.94
and 20-43218.95. OEG also acknowledges a Heisenberg fel-
lowship while at Heidelberg.
REFERENCES
Awaki H., et al. , 1994, PASJ 46, L65
Arnaboldi M., Freeman K.C., Hui X., Capaccioli M., Ford H.,
1994, Messenger 76, 40
Bender R., 1990, A&A 229, 441
Bender R., Saglia, R., Gerhard, O.E., 1994, MNRAS 269, 785
(BSG)
Binney J.J., 1978, MNRAS 183, 501
Binney, J.J., Davies, R.L., Illingworth, G.D., 1990, ApJ 361, 78
Binney J.J., Mamon G.A., 1982, MNRAS 200, 361
Carollo C.M., de Zeeuw P.T., van der Marel R.P., Danziger I.J.,
Qian E.E., 1995, ApJL 441, L25
Dehnen W., Gerhard O.E., 1994, MNRAS 268, 1019
Dejonghe H., 1987, MNRAS 224, 13
Dejonghe H., de Bruyne V., Vauterin P., Zeilinger W.W., 1996,
A&A 306, 363
Dejonghe H., Merritt D., 1992, ApJ 391, 531
de Vaucouleurs G., de Vaucouleurs A., Corwin H.G., 1976, Second
Reference Catalogue of Bright Galaxies, Univ. Texas Press,
Austin
Faber S.M., Wegner G., Burstein D., Davies R.L., Dressler A.,
Lynden-Bell D., Terlevich R.J., 1989, ApJS 69, 763
Franx M., van Gorkom J.H., de Zeeuw T., 1994, ApJ 436, 642
Fricke W., 1952, Astr. Nachr. 280, 193
Gerhard O.E., 1991, MNRAS 250, 812 (G91)
Gerhard O.E., 1993, MNRAS 265, 213 (G93)
Grillmair C.J., Freeman K.C., Bicknell G.V., Carter D., Couch
W.J., Sommer-Larsen J., Taylor K., 1994, ApJ 422, L9
Hanson R.J., Haskell K.H., 1981, Math. Programm. 21, 98.
Jaffe W., 1983, MNRAS 202, 995
Jeske, G., 1995, PhD Thesis, University of Heidelberg
Jeske, G., Gerhard, O.E., Bender, R., Saglia, R.P., 1996, Proc.
IAU 171, eds. Bender, R., Davies, R.L., p. 397
Kim D.-W., Fabbiano G., 1995, ApJ 441, 182
Kochanek C.S., Keeton C.R., 1997, in The Nature of Elliptical
Galaxies, 2nd Stromlo Symposium, eds. Arnaboldi M., Da
Costa G.S., Saha P., 1997, ASP 116, 21.
Maoz D., Rix H.-W., 1993, ApJ 416, 435
Merritt D., 1985, AJ 90, 1027
Merritt D., 1993, ApJ 413, 79
Osipkov L.P., 1979, Pis'ma Astr. Zh. 5, 77
Press, W.H., Flannery, B.P., Teukolsky, S.A., Vetterling, W.T.,
Numerical Recipes (Cambridge: Cambridge University Press)
Rix H.-W., de Zeeuw P.T., Carollo C.M., Cretton N., van der
Marel R.P., 1997, preprint
Saglia R.P. et al. , 1993, ApJ 403, 567
Saglia, R.P., Bender, R., Gerhard, O.E., Jeske, G., 1997a, in Dark
and Visible Matter in Galaxies and Cosmological Implica-
tions, eds. Persic M., Salucci P., ASP 117, 113.
Saglia R.P. et al. , 1997b, ApJS 109, 79
van der Marel, R., 1991, MNRAS 253, 710
van der Marel R.P., Franx M., 1993, ApJ 407, 525
Worthey, G. 1994, ApJS, 95, 107
APPENDIX A: LIBRARY OF ANISOTROPIC
SPHERICAL DISTRIBUTION FUNCTIONS
To understand the connection between anisotropy struc-
ture and observable line profile shapes we have constructed
a number of spherical distribution functions of the quasi --
separable form (Gerhard 1991, G91)
f (E, L) = g(E) h(x),
(18)
where the variable x depends on both energy and angular
momentum:
L
(19)
x =
L0 + Lc(E)
.
L0 is an angular momentum constant, or equivalently, an
anisotropy radius times a characteristic velocity; Lc(E) is
the angular momentum of the circular orbit at energy E. dfs
of the form (18) have the following properties: (i) the circu-
larity function h(x) has the effect of shifting stars between
orbits of different angular momenta on surfaces of constant
energy, while g(E) controls the distribution of stars between
energy surfaces. (ii) For the most bound stars L < Lc(E) ≪
L0; thus the model becomes isotropic in the centre unless
L0 = 0. (iii) For loosely bound stars L ∼ Lc(E) ≫ L0,
i.e., the angular momentum distribution becomes a func-
tion of circularity L/Lc(E) which is one-to-one related to
orbital eccentricity. Outside the anisotropy radius, the df
(18) therefore corresponds to an energy-independent orbit
distribution with constant anisotropy, radial or tangential.
In these models h(x) is an assigned function built into
the model to achieve the desired anisotropy (orbit distribu-
tion). Radially biased distribution functions correspond to
circularity functions h(x) decreasing with x; for example
h(x) = hα(x) ≡(cid:0)1 − x2(cid:1)α
.
(20)
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
The dark halo of the E0 galaxy NGC 6703
19
µ(E) =
1
2
+
1
π
arctan(cid:20) k
2
2
E2 − E
E2
(cid:21).
(23)
The parameters E and k determine the orbital energy near
which the anisotropy transition occurs, and the width of the
transition. A similar function µ(E) was used for models in
halo potentials. Figure 18 shows the intrinsic and projected
properties of a number of dfs of this kind, constructed in
the self -- consistent potential of a Jaffe sphere. Notice the
wide variety of kinematical profiles that can be constructed
in this way.
APPENDIX B: TRANSFORMING TO LINEAR
KINEMATIC DATA
In velocity line profile measurements, the depth of an ab-
sorption line in a spectral resolution element is assumed to
be proportional to the number of stars with line-of-sight ve-
locities corresponding to this wavelength interval. The line-
of-sight velocity distribution measured from the line profiles
is a discretized function linearly related to the underlying
df [cf. eq. (1)]. This linearity is lost when line profile mea-
surements are represented by the quantities v, σ, h3, h4 --
these quantities are obtained by a least-square fit of a Gauss-
Hermite series to the observed line profile.
To re-express the observed kinematics in terms of quan-
tities that depend linearly on f we proceed as follows. Con-
sistent with the assumption of spherical symmetry, we set
the mean streaming velocity v and all odd velocity profile
moments to zero. Next, we obtain an estimate for the veloc-
ity dispersion (second moment) σ2, by integrating over the
line profile specified by (σ, h4); for negative h4, until it first
becomes zero. For small h4, the linear correction formula
σ = σ(1 + √6h4) holds (van der Marel & Franx 1993), this
correction results in σ > σ for peaked profiles with h4 > 0.
The numerical correction from integrating over the velocity
profile also has this property (BSG).
From the measured h4(Ri), we compute new even
(24)
(25)
hj (Ri) Hj(x) exp(−x2/2)
(h0 = 1, h1 = h2 = h3 = 0) with x ≡ vk/σ(Ri) as
Nn sn Hn(ex) exp(−ex2/2),
Gauss-Hermite moments sn(Ri;eσ) by expanding the series
L(Ri, vk) = L0 Xj=0,4
L(Ri, vk) = Xn=0,2,4,...
where ex ≡ vk/eσ(Ri). Here Hn are Hermite polynomials,
the Nn are normalization constants (G93), and eσ(Ri) are
practice, we have found it convenient to take foreσ(Ri) the
galaxy being analyzed. The sn(Ri;eσ) are estimates for the
sn(Ri;eσ) = (cid:0)2n−1n!(cid:1)−1/2
Gauss-Hermite moments related to the true velocity profile
by
velocity dispersions σiso(Ri) of the isotropic model in the
given potential Φ(r) with the same stellar density as the
fiducial scaling velocities generally different from σ(Ri). In
(26)
×Z ∞
−∞
For a theoretical model, we obtain corresponding mo-
dvk Hn(vk/eσ) exp(−v2
k/2eσ2)L(Ri, vk).
Figure 18. Anisotropy parameter β(r), projected velocity disper-
sion σp, and line profile shape parameter h4 for several families
of anisotropic dfs. Left: scale -- free radially anisotropic (dotted)
and tangentially anisotropic (dashed lines) in the potential of a
self -- consistent Jaffe sphere. For these models, circularity func-
tions of the type (20) and (21) were used with different α and
c, respectively. The isotropic model is shown for reference (solid
line). Right: Families of models with anisotropy changing from
radial to tangential (dotted) or from tangential to radial (dashed
lines). These were constructed from the same circularity functions
and a weight factor as in eq. (23).
The family can also be used to construct tangentially aniso-
tropic models, such as
(21)
hc,α(x) = c + (1 − c)(cid:2)1 − (1 − x2)α(cid:3) .
In these tangentially anisotropic models one cannot choose
h(0) = 0 unless L0 = 0, otherwise the density at r = 0 would
be zero. Of course, other forms for h(x) are possible, such
as Gaussians.
Given the assigned function h(x), the integral equation
for ρ(r) in terms of f (E, L2) is solved for the derived func-
tion g(E); see G91 and Jeske (1995). Fig. 3 shows line profile
shape parameters for representative dfs constructed in this
way. Fig. 3 shows the anisotropy profiles of two sets of tan-
gentially anisotropic models. Here the density of stars has
been taken to be that of a Jaffe sphere, and the potential
in which the stars orbit is either the self -- consistent poten-
tial or one of the mixed stars plus halo potentials used in
Sections 4, 5. Sequences like that in Fig. 3 are used as basis
functions in the non -- parametric analysis in Section 5.
Models whose anisotropy changes from radial to tan-
gential or vice versa were constructed by linearly combining
the above circularity functions with energy -- dependent co-
efficients. In this way one obtains dfs of the more general
form
f (E, L) = g(E) h(E, x).
(22)
For self -- consistent Jaffe models we used energy -- dependent
coefficients µ(E) of the following form:
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
20 O.E. Gerhard, G. Jeske, R.P. Saglia, R. Bender
ments by inserting the projected df N (Ri, vk) from (1) into
the new sn-moments of the composite model are linear in
the ak, i.e.,
eq. (26) instead of L(Ri, vk), using the sameeσ(Ri). Clearly,
sn(Ri;eσ) = Xk=1,K
n (Ri;eσ)
with the s(k)
n corresponding to the respective fk.
aks(k)
(27)
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
|
astro-ph/9608098 | 1 | 9608 | 1996-08-15T11:22:04 | Depolarization of the cosmic microwave background by a primordial magnetic field and its effect upon temperature anisotropy | [
"astro-ph"
] | We estimate the depolarizing effect of a primordial magnetic field upon the cosmic microwave background radiation due to differential Faraday rotation across the last scattering surface. The degree of linear polarization of the CMB is significantly reduced at frequencies around and below 30 GHz $(B_* /10^{-2}{\rm Gauss})^{1/2}$, where $B_*$ is the value of the primordial field at recombination. The depolarizing mechanism reduces the damping of anisotropies due to photon diffusion on small angular scales. The $l\approx 1000$ multipoles of the CMB temperature anisotropy correlation function in a standard cold dark matter cosmology increase by up to 7.5\% at frequencies where depolarization is significant. | astro-ph | astro-ph |
Depolarization of the cosmic microwave background
by a primordial magnetic field
and its effect upon temperature anisotropy
Diego D. Harari,a,b,1 Justin D. Haywarda,c,2 and Matias Zaldarriagad,3
a D.A.R.C., Observatoire de Paris - Meudon
5 Place Jules Janssen, 92195 Meudon, France
b Departamento de F´ısica, Facultad de Ciencias Exactas y Naturales
Universidad de Buenos Aires, Ciudad Universitaria - Pab. 1
1428 Buenos Aires, Argentina
c Department of Applied Mathematics and Theoretical Physics
University of Cambridge, Cambridge CB3 9EW, UK
d Department of Physics, MIT, Cambridge, MA 02139 USA
Abstract
We estimate the depolarizing effect of a primordial magnetic field upon
the cosmic microwave background radiation due to differential Faraday rota-
tion across the last scattering surface. The degree of linear polarization of
the CMB is significantly reduced at frequencies around and below 30 GHz
(B∗/10−2Gauss)1/2, where B∗ is the value of the primordial field at recom-
bination. The depolarizing mechanism reduces the damping of anisotropies
due to photon diffusion on small angular scales. The l ≈ 1000 multipoles of
the CMB temperature anisotropy correlation function in a standard cold dark
matter cosmology increase by up to 7.5% at frequencies where depolarization
is significant.
PACS numbers: 98.80.-k, 98.70.Vc, 98.80.Es
1Email address: [email protected]
2Email address: [email protected]
3Email address: [email protected]
1
Introduction
The cosmic microwave background radiation (CMB) is expected to have acquired a
small degree of linear polarization through Thomson scattering [1], which polarizes the
radiation if there is a quadrupole anisotropy in its distribution function [2]. Typically,
the CMB degree of linear polarization is expected to be more than ten times smaller
than the relative temperature anisotropy on comparable angular scales, at least within
a standard ionization history of the Universe. The CMB has not yet been observed
to be polarized, the upper limit on its degree of linear polarization on large angular
scales being P < 6 × 10−5 [3]. When measured, the CMB polarization will provide
a wealth of information about the early Universe, additional to that revealed by the
CMB anisotropy.
The polarization properties of the CMB may prove particularly valuable to either
constrain or detect an hypothetical primordial magnetic field [4, 5]. A cosmological
magnetic field could leave significant imprints upon the CMB polarization through
the effect of Faraday rotation. After traversing a distance L in a direction q within
an homogeneous magnetic field ~B, linearly polarized radiation has its plane of polar-
ization rotated an angle
ϕ =
e3nexe ~B · q
8π2m2c2 λ2L .
(1)
ne is the total number-density of electrons and xe its ionized fraction. λ is the wave-
length of the radiation, m is the electron mass, and c is the speed of light. We work
in Heaviside-Lorentz electromagnetic units (α = e2/4π ≈ 1/137 is the fine structure
constant if we take ¯h = c = 1).
Faraday rotation of synchrotron emission by distant galaxies serves, for instance,
to estimate the value of galactic and extragalactic magnetic fields [6]. Faraday ro-
tation acts also as a depolarizing mechanism. If an extended source emits polarized
radiation, the total outcome may become significantly depolarized by a magnetic
field, after the radiation emanating from points at different depths within the source
experience different amounts of Faraday rotation. This process affects significantly
the radio emission of galaxies and quasars [7].
In this paper we analyze the depolarizing effect exerted by a primordial magnetic
field upon the CMB across the last scattering surface. We consider a Robertson-
Walker universe with scalar, energy-density fluctuations, and assume a standard
thermal history. We make use of an analytic approach [8], based on a recent re-
finement and extension [9] of the tight-coupling approximation [10], that highlights
the physical process responsible for the CMB polarization and its dependence upon
various cosmological parameters, while still yielding reasonably accurate results. The
polarization of the CMB is proportional to the width of the last scattering surface
(LSS), the interval of time during which most of the CMB photons that we observe
today last-scattered off free electrons. A primordial magnetic field could prevent the
1
polarization from growing across the full width of the LSS. We shall see that the effect
is controlled by the dimensionless and time-independent parameter
F ≡
3
2πe
B
ν2 ≈ 0.7(
B∗
10−3Gauss(cid:17)(cid:16)10GHz
ν0
(cid:17)2
.
(2)
The coefficient F represents the average Faraday rotation (in radians) between Thom-
son scatterings [4]. ν0 is the CMB frequency observed today. B∗ = B(t∗) is the
strength of the primordial magnetic field at a redshift z∗ = 1000, around the time
of decoupling of matter and radiation. Current bounds suggest that a magnetic field
pervading cosmological distances, if it exists, should have a present strength below
B0 ≈ 10−9 Gauss [6]. It is conceivable that the large scale magnetic fields observed
in galaxies and clusters have their origin in a primordial field, and several theoretical
speculations exist about its possible origin [11]. A primordial magnetic field is ex-
pected to scale as B(t) = B(t0)a2(t0)/a2(t), where a(t) is the Robertson-Walker scale
factor. Thus, a primordial field with strength B∗ = 10−3 Gauss at recombination
would have a present strength
B0 =
B∗
(1 + z∗)2 ≈ 10−9 Gauss (
B∗
10−3 Gauss
)
.
(3)
A primordial magnetic field may significantly depolarize the CMB right before
its decoupling from matter. The effect is sensitive to the strength of the magnetic
field at recombination, not to its present strength. A value of B∗ somewhat larger
than 10−3 Gauss is not ruled out. Compatibility with big-bang nucleosynthesis, for
instance, places an upper bound that, extrapolated to the time of recombination, is
at most B∗ = 0.1 Gauss [12]. Recent proposals for either detecting or constraining a
primordial field at recombination were suggested in [5, 13]. In [5], Faraday rotation of
the CMB polarization was analyzed in the limit of small rotation angles, concluding
that a measurement of the effect could provide evidence for magnetic fields of order
B∗ ≈ 10−3 Gauss at recombination. In [13] the change in the photon-baryon sound
speed in the presence of a magnetic field of order B∗ = 0.2 Gauss was claimed to
distort the structure of the acoustic peaks in the CMB anisotropy power spectrum at
a level detectable by currently planned CMB experiments.
We shall entertain in our discussions the possibility that the strength of the pri-
mordial magnetic field at recombination be somewhat larger than B∗ = 10−3 Gauss.
We will show that currently planned CMB experiments might be sensitive to the effect
of depolarization upon the temperature anisotropy power spectrum on small angu-
lar scales if B∗ is around or larger than 0.01 Gauss, while experiments at somewhat
lower frequencies would be sensitive to primordial fields of strength around B∗ ≈ 10−3
Gauss.
The impact of depolarization upon anisotropy comes about as a consequence of
the polarization-dependence of Thomson scattering, which feeds back polarization
2
into anisotropy [8, 14, 15]. The dominant effect is a reduction in the exponential
damping due to photon diffusion, which results in an increase of the anisotropy at
those frequencies for which depolarization is significant. We shall perform an analytic
estimate of the effect, based on the tight coupling approximation. In order to make
more quantitative and specific predictions about the impact and potential measura-
bility of the effect of depolarization upon temperature anisotropy, we shall also use
a recently developed numerical code [16] to integrate the Boltzmann equations in
a standard cold dark matter model. We shall see that the temperature anisotropy
correlation function multipoles at l ≈ 1000 increase by up to 7.5% at frequencies
where depolarization is significant. We conclude that a primordial magnetic field of
strength around 10−2 Gauss at recombination is worth of membership in the list of
multiple cosmological parameters that one may attempt to determine through CMB
anisotropy measurements on small angular scales [17].
The paper is structured as follows. In section II, we write down and describe the
radiative transfer equations for the total and polarized photon-distribution function
in the presence of a single Fourier mode of the scalar metric fluctuations. We include
the term describing Faraday rotation by a primordial magnetic field. We solve these
equations in the tight coupling approximation, and find the dependence of the degree
of polarization upon the frequency of the CMB photons in the presence of a primordial
magnetic field. In section III we discuss the effects of the depolarizing mechanism
upon the anisotropy of the CMB on small angular scales, both analytically as well
as numerically. We discuss the possibility that the effect be detected by currently
planned CMB experiments. Section IV is the discussion and conclusion.
2 Depolarization by a magnetic field
2.1 Boltzmann equations
We begin by considering the radiative transfer equations for a single Fourier mode
of the temperature and polarization fluctuations in a Robertson-Walker spatially-flat
Universe with scalar (energy-density) metric fluctuations, described in terms of the
gauge-invariant gravitational potentials Ψ and Φ. We follow the notation and for-
malism of Ref.
[8]. The total temperature fluctuation is denoted by ∆T , while the
fluctuation in the Stokes parameters Q and U are denoted by ∆Q and ∆U respec-
tively. The degree of linear polarization is given by ∆P = (∆2
U )1/2. All three
quantities are expanded in Legendre polynomials as ∆X = Pl(2l + 1)∆XlPl(µ), where
µ = cos θ = ~k · q/~k is the cosine of the angle between the wave vector of a given
Fourier mode ~k, and the direction of photon propagation q. The evolution equations
for the Fourier mode of wave vector ~k of the gauge-invariant temperature and polar-
ization fluctuations [8, 9, 18, 19], including the Faraday rotation effect of a primordial
Q + ∆2
3
magnetic field [5], read
∆T + ikµ(∆T + Ψ) = − Φ − κ[∆T − ∆T0 − µVb +
1
2
P2(µ)SP ]
∆Q + ikµ∆Q = − κ[∆Q −
1
2
(1 − P2(µ))SP ] + 2ωB∆U
∆U + ikµ∆U = − κ∆U − 2ωB∆Q .
We have defined
SP ≡ −∆T2 − ∆Q2 + ∆Q0
(4)
(5)
(6)
(7)
which acts as the effective source term for the polarization. Vb is the bulk velocity of
the baryons, which verifies the continuity equation
Vb = −
a
a
Vb − ikΨ +
κ
R
(3∆T1 − Vb)
.
(8)
An overdot means derivative with respect to the conformal time τ = R dta0/a, with
a(t) the scale factor of the spatially flat Robertson-Walker metric, and a0 = a(t0) its
value at the present time. R ≡ 3ρb/4ργ coincides with the scale factor a(t) normalized
to 3/4 at the time of equal baryon and radiation densities.
κ = xeneσT a/a0 is the
Thomson scattering rate, or differential optical depth, with ne the electron number
density, xe its ionized fraction, and σT the Thomson scattering cross-section. Finally,
ωB is the Faraday rotation rate [5]
dϕ
dτ
=
e3nexe ~B · q
8π2m2ν2
a
a0
ωB ≡
(9)
If there were axial symmetry around ~k and no Faraday rotation, one could always
choose a basis for the Stokes parameters such that U = 0. A magnetic field with
arbitrary orientation breaks the axial symmetry, and Faraday rotation mixes Q and
U.
2.2 Tight coupling approximation
We now solve the equations (4,5,6) in the tight - coupling approximation, which
amounts to an expansion in powers of kτC, where τC ≡ κ−1 is the average conformal
time between collisions.
At times earlier than decoupling, Thomson scattering is very efficient, and the
mean free path of the photons is very short. The lowest order tight-coupling expres-
sion constitutes in that case an excellent approximation. It implies that the photon
distribution function is isotropic in the baryon's rest frame, and hence the polariza-
tion vanishes [8]. To first order in kτC there is a small quadrupole anisotropy, and
4
thus a small polarization. As decoupling of matter and radiation proceeds, the tight-
coupling approximation breaks down. Still, for wavelengths longer than the width
of the last scattering surface, it provides a very accurate approximation to the exact
result.
In the absence of a magnetic field (ωB = 0), the tight - coupling solutions, to first
order in kτC, are such that [8]
∆U = 0 ; ∆Q =
3
4
SP sin2 θ
SP = −
5
2
∆T2 =
4
3
ikτC∆T1 = −
4
3
τC ∆0
,
(10)
(11)
where we defined ∆0 ≡ ∆T0 + Φ. Notice that ∆Q0 = −5∆Q2 = − 5
2SP , while
all multipoles with l ≥ 3 vanish to first order in kτC. All quantities of interest can be
expressed, in the tight-coupling approximation, in terms of ∆0, which in turn verifies
the equation of a forced and damped harmonic oscillator [9]
4 ∆T2 = 1
∆0 + h
R
1 + R
+
16
45
k2τC
(1 + R)i ∆0 +
k2
3(1 + R)
∆0 =
k2
3(1 + R)
[Φ − (1 + R)Ψ]
,
(12)
where we have neglected O(R2) corrections.
Now consider the effect of the magnetic field (ωB 6= 0), assumed spatially homo-
geneous over the scale of a perturbation with wave-vector ~k. Faraday rotation breaks
the axial symmetry around the direction of the wave-vector. The depolarizing effect
of Faraday rotation depends not only upon the angle between the magnetic field and
the direction in which the radiation propagates, but also upon the angle between the
magnetic field and the wavevector ~k. Nevertheless, we shall only be interested in
the stochastic superposition of all Fourier modes of the density fluctuations, with a
Gaussian spectrum that has no privileged direction. Average quantities thus depend
only upon the angle between the line of sight and the direction of the magnetic field,
but not upon the angle between the magnetic field and the wavevector ~k, which is
integrated away. For simplicity of the calculation, when computing the evolution of
perturbations with wave-vector ~k we shall consider a magnetic field with no compo-
nent perpendicular to ~k. This choice also satisfies the condition of axial symmetry
around ~k, under which eqs. (4,5,6) for ∆T , ∆Q and ∆U were derived. We shall later
use the result of this calculation for the stochastic superposition of all Fourier modes
with arbitrary orientation relative to the magnetic field. This simplification will result
at most in an underestimate of the net depolarizing effect, since the case ~B k ~k is
that for which depolarization is less effective, the magnetic field being perpendicular
to the direction in which polarization is maximum.
To first order in kτC the tight-coupling solutions in the presence of a homogeneous
5
magnetic field ~B k ~k are such that:
∆U = −F cos θ∆Q ; ∆Q =
3
4
SP sin2 θ
(1 + F 2 cos2 θ)
where we have defined the coefficient F as
F cos θ ≡ 2ωBτC
and so
F =
e3
4π2m2σT
B
ν2 ≈ 0.7(cid:16)
B∗
10−3Gauss(cid:17)(cid:16)10 GHz
ν0
(cid:17)2
.
(13)
(14)
(15)
The coefficient F represents the average Faraday rotation between collisions, since
2ωB is the Faraday rotation rate and τC = κ−1 is the photons mean free path (in con-
formal time units). When calculating the evolution of each mode ~k we have assumed
that the strength of the primordial magnetic field scales as B(t) = B(t∗)a2(t∗)/a2(t),
which is justified by flux conservation and because the Universe behaves as a good
conductor [6]. Since the frequency also redshifts as ν = ν0a(t0)/a(t), the parameter F
is time-independent. ν0 is the frequency of the CMB photons at present time, while
B∗ is the strength of the magnetic field at a redshift z∗ = 1000, around recombina-
tion. Within a standard thermal history, with no early reionization, depolarization
is only significant across the LSS, and it thus depends only upon the value of the
primordial magnetic field around the time of recombination. Notice that Faraday
rotation between collisions becomes considerably large, paving the way to an efficient
depolarizing mechanism, at frequencies around and below νd defined such that
F ≡ (cid:16)νd
ν0(cid:17)2
so that
νd ≈ 8.4 GHz (cid:16)
B∗
10−3Gauss(cid:17)1/2
≈ 27 GHz (cid:16)
B∗
0.01 Gauss(cid:17)1/2
.
(16)
(17)
From eqs. (13) we can read the values of ∆Q0 and ∆Q2. They reduce to eqs. (10)
with O(F 2) corrections for small F , while they vanish as F −1 for large F . We write
them as:
∆Q0 =
d0(F )SP
; ∆Q2 = −
1
2
1
10
d2(F )SP
.
(18)
The coefficients d0, d2 are defined so that di ≈ 1 + O(F 2) for small F , while di →
O(1/F ) as F → ∞, and represent the effect of depolarization. They read:
d0(F ) =
3
2harctan(F )
F
(1 +
d2(F ) =
15
4 harctan(F )
F
(1 +
4
F 2 +
6
1
F 2 ) −
3
F 4 ) −
1
F 2i
3
F 2 −
(19)
(20)
3
F 4i
.
In terms of the combination
5
6
d ≡
(d0 +
d2
5
) =
15
8 harctan(F )
F
(1 +
2
F 2 +
1
F 4 ) −
5
3F 2 −
1
F 4i
and using the definition of SP , we find the relation
∆T2 = −SP (1 −
3
5
d)
and from the equation for ∆T in the tight coupling limit we get
SP =
4
3(3 − 2d)
ikτC∆T1 = −
4
3(3 − 2d)
τC ∆0
.
(21)
(22)
(23)
Notice that d ≈ 1 − F 2/7 if F << 1 while d → 15
16πF −1 for large F . We stress
here again that in a general case the depolarizing coefficient d depends upon the
angle between ~k and ~B. The net anisotropy and polarization being the outcome of
the stochastic superposition of all Fourier modes of the density-fluctuations, with
a spectrum that has no privileged direction, the average depolarizing factor, after
superposition of all wavevectors in arbitrary orientations with respect to the magnetic
field, depends only upon F . The average depolarizing factor might slightly differ from
that calculated with ~k k ~B, which at most underestimates the average effect.
Equations (13,22,23) condense the main effects of a magnetic field upon polariza-
8 ∆T2 sin2 θ.
tion. When there is no magnetic field (F = 0, d = 1) ∆U = 0 and ∆Q = − 15
A magnetic field generates ∆U , through Faraday rotation, and reduces ∆Q. In the
limit of very large F (large Faraday rotation between collisions) the polarization van-
ishes. The quadrupole anisotropy ∆T2 is also reduced by the depolarizing effect of
the magnetic field, by a factor 5/6 in the large F limit, because of the feedback of
∆Q upon the anisotropy or, in other words, because of the polarization dependence
of Thomson scattering. The dipole ∆T1 and monopole ∆T0 are affected by the mag-
netic field only through its incidence upon the damping mechanism due to photon
diffusion for small wavelengths, that we shall discuss in detail in section III. Indeed,
the equation for ∆0 = ∆T0 + Φ, neglecting O(R2) contributions, now reads:
∆0 + h
R
1 + R
+
16
90
(5 − 3d)
(3 − 2d)
k2τC
(1 + R)i ∆0 +
k2
3(1 + R)
∆0 =
k2
3(1 + R)
[Φ − (1 + R)Ψ]
.
(24)
The damping term is reduced by a factor 5/6 at frequencies such that d << 1, for
which depolarization is significant.
We have assumed that the magnetic field is spatially homogeneous. We can expect
corrections to our result if the field is inhomogeneous over scales smaller than τC
at any given time around decoupling.
Indeed, if the field reverses its direction N
times along a photon path during a time τC, Faraday rotation will not accumulate
7
as assumed above. In that case depolarization would start to be significant only at
those frequencies such that Faraday rotation is large over the scale on which the
magnetic field reverses its direction. The frequencies at which depolarization starts
to be significant would thus be reduced by a factor 1/√N .
2.3 Frequency-dependence of the degree of polarization
The anisotropy and polarization observed at present time can be evaluated using the
formal solutions of eqs. (4,5,6)
∆T (τ0) = R τ0
∆Q(τ0) = Z τ0
0
2 P2(µ)SP (τ )]
0 dτ eikµ(τ −τ0)g(τ )[∆T0(τ ) + µVb(τ ) − 1
0 dτ eikµ(τ −τ0)e−κ(τ0,τ )( Ψ − Φ)
+R τ0
dτ eikµ(τ −τ0)g(τ ){
1
2
[1 − P2(µ)]SP (τ ) + F ∆U (τ )}
where
∆U (τ0) = −Z τ0
0
dτ eikµ(τ −τ0)g(τ )F ∆Q(τ )
g(τ ) ≡ κe−κ(τ0,τ )
(25)
(26)
(27)
(28)
is the visibility function. It represents the probability that a photon observed at τ0
last-scattered within dτ of a given τ . For a standard thermal history, with no signifi-
cant early reionization after recombination, g(z) is well approximated by a Gaussian
centered at a redshift of about z ≈ 1000 and width ∆z ≈ 80 [20]. In conformal time,
we shall denote the center and width of the Gaussian which approximately describes
the process of decoupling by τD and ∆τD respectively.
The visibility function being strongly peaked around the time of decoupling, the
first integral in eq. (25) for the anisotropy is well approximated, at least for wave-
lengths longer than the width of the last scattering surface, by its instantaneous
recombination limit. In that case it reduces to the tight-coupling expression of its
integrand evaluated at time τ = τD [9]. This first integral is dominated by its first
two terms, proportional to the monopole ∆T0 and the baryon velocity Vb (in turn
proportional to ∆T1) respectively. The quadrupole term SP gives a negligible contri-
bution for long wavelengths, but becomes relatively significant on small scales. The
second integral in eq. (25) corresponds to the anisotropies induced by time-dependent
potentials after the time of last-scattering.
Eqs. (26,27) for the polarization can be approximated replacing the integrand by
its tight-coupling expression. Then
∆Q(τ0) =
3
4
sin2 θ
(1 + F 2 cos2 θ) Z τ0
0
dτ eik cos θ(τ −τ0)g(τ )SP (τ )
(29)
8
while the total polarization, ∆P = (∆2
∆U (τ0) = −F cos θ∆Q(τ0)
U )1/2 reads
Q + ∆2
∆P (τ0) = √1 + F 2 cos2 θ∆Q(τ0)
(30)
(31)
Evaluation of the time integral in eq. (29) requires a more detailed knowledge of
the time-dependence of the integrand than in the case of the anisotropy.
Indeed,
the tight-coupling expression (23) for the quadrupole term SP being proportional to
the mean free path τC, which varies rapidly during decoupling, the instantaneous
recombination approximation becomes inappropriate. The induced polarization is,
indeed, proportional to the width of the last scattering surface. Adapting the method
of [8] to include also the effect of the primordial magnetic field, we write down the
equation satisfied by SP when all other quantities are already approximated by their
first-order tight-coupling expressions:
SP +
3
10
(3 − 2d) κSP =
2
5
ik∆T1
(32)
Neglect of SP returns the tight-coupling result of eq. (23). Instead, the formal solution
to equation (32)
SP (τ ) =
2
5
ik Z τ ′
0
dτ ′∆T1e−
3
10 κ(τ,τ ′)(3−2d)
(33)
tracks down the time-dependence of SP through the decoupling process with better
accuracy.
For wavelengths longer than the width of the LSS we can neglect the time variation
of ∆T1 and that of eik cos θ(τ −τ0) around decoupling. We also approximate the visibility
function by a Gaussian, which justifies the approximation κ(τ0, τ ) ≈ −κ(τ0,τ )
[21].
Then
∆τD
3
10 xκ(3−2d)
e−
,
(34)
SP (τ ) ≈
2
5
ik∆T1(τD)∆τDe
3
10 κ(τ0,τ )(3−2d) Z ∞
1
dx
x
where the integration variable has been changed to x = κ(τ0,τ )
approximations,
κ(τ0,τ ′). Thus, within these
R τ0
0 dτ g(τ )SP (τ ) = − 2
= 4
5 ik∆T1(τD)∆τD R ∞
1+6d ik∆T1(τD)∆τD[ln( 10
0 dκe−
1+6d
10 κEi(− 3
3 ) − ln(3 − 2d)]
10 (3 − 2d)κ)
(35)
Finally, the total polarization induced at an angle θ with respect to the wavector
~k, reads
∆P (τ0) =
3
(1 + 6d)
[ln(
10
3
) − ln(3 − 2d)]
sin2 θeik cos θ(τD −τ0)
√1 + F 2 cos2 θ
ik∆T1(τD)∆τD .
(36)
9
It can also be written as follows, in terms of the polarization that would be induced
if there were no magnetic field (or equivalently, in terms of the polarization at fre-
quencies large enough such that the depolarizing effect is negligible):
where we have defined the depolarizing factor as
∆P (θ, F ) = D(θ, F )∆P (B = 0)
with
D(θ, F ) =
1
√1 + F 2 cos2 θ
f (F )
f (F ) =
7
1 + 6dh1 −
ln(3 − 2d)
ln(10/3) i
(37)
(38)
(39)
Eq. (38), together with the defining eqs. (15,16) and (21) for F and d, summarize
the main result of this section. Notice that
f → 1
as F → 0 ;
f → 0.61 as F → ∞
(40)
The polarization observed at present times depends upon the angle between the
line of sight and the orientation of the magnetic field at the time of decoupling. There
will be no depolarization if the magnetic field is perpendicular to the line of sight.
The magnetic field is likely to change orientation randomly over scales longer than
the Hubble radius at the time of decoupling, which subtends an angle of order one
degree in the sky, so that after averaging over many regions separated by more than
a few degrees, we can always expect a net average depolarizing effect. To roughly
estimate its order of magnitude we could assume an average component of ~B parallel
to the observation direction of order B/√2 and define an average ¯D as
¯D =
1
q1 + F 2/2
f (F )
.
(41)
Fig. 1 displays the depolarizing factor ¯D as a function of the CMB frequency ν0. We
have plotted it for three different values of the magnetic field B∗ to help visualize
the relevant frequency range, but notice that since depolarization depends only upon
F = (νd/ν0)2, the plot for an arbitrary value of B∗ is identical to that corresponding
to another value of the magnetic field after an appropriate scaling of the frequency
units, proportional to the square root of the magnetic field.
At low frequencies, those for which the effect is large, the average depolarizing
factor scales as
√2
F ≈ 0.85(cid:16) ν0
νd(cid:17)2
¯D ≈ 0.6
if
ν0 << νd
.
At comparatively large frequencies instead
¯D ≈ 1 − 0.36F 2 = 1 − 0.36(cid:16)νd
ν0(cid:17)4
if
ν0 >> νd
.
(42)
(43)
10
3 Effects upon the anisotropy
Depolarization by a primordial magnetic field has significant and potentially measur-
able effects upon the anisotropy of the CMB on small angular scales.
Indeed, the
polarization properties of the CMB feed back into its anisotropy, as evidenced in eq.
(4), due to the polarization dependence of Thomson scattering. The dominant effect
of polarization upon anisotropy derives from its impact upon the photon diffusion
length[8, 14, 15], which damps anisotropies on small angular scales [9, 22, 23]. It was
shown in [14], through numerical integration of the Boltzmann equations, that neglect
of the polarization properties of the CMB leads to an overestimate of its anisotropy
on small angular scales as large as 10%. We thus expect depolarization by a primor-
dial magnetic field to introduce a significant frequency-dependent distortion of the
CMB anisotropy power spectrum. Notice that a different (frequency-independent)
distortion of the CMB anisotropy power spectrum by a primordial magnetic field,
due to its impact upon the photon-baryon fluid sound speed, was recently discussed
in [13].
3.1 Reduced diffusion-damping
Photon diffusion damps anisotropies on small angular scales [9, 22, 23]. The effect
is described by the term proportional to k2τC ∆0 in eq. (24). The photon diffusion
length depends upon the degree of polarization of the CMB [8, 14, 15]. Thus, the
photon-diffusion length is different at frequencies where the depolarizing effect is
significant.
The damping of anisotropies on small angular scales due to photon diffusion can be
found, now including the full R-dependence, by solving the tight-coupling equations
to second order, assuming solutions of the form
∆X(τ ) = ∆X eiωτ
(44)
for X = T , Q, and U, and similarly for the baryon velocity Vb. One then finds that
ω =
k
q3(1 + R)
+ iγ
with the photon-diffusion damping length-scale determined by
γ(d) ≡
k2
k2
D
=
k2τC
6(1 + R)(cid:16) 8
15
(5 − 3d)
(3 − 2d)
+
R2
1 + R(cid:17)
(45)
(46)
The depolarizing effect of the magnetic field reduces the viscous damping of anisotro-
pies. In the case of small R, such that R2 terms can be neglected, the damping factor
11
γ is smaller by a factor 5/6 at those frequencies for which the depolarizing effect is
large.
We make now an analytic estimate of the effect of the frequency-dependence of
the photon diffusion length, in the presence of a primordial magnetic field, upon the
CMB anisotropy power spectrum. The temperature anisotropy correlation function
is typically expanded in Legendre polynomials as
C(θ) =< ∆T (n1)∆T (n2) >n1·n2=cos θ=
1
4π
∞
Xl=0
(2l + 1)ClPl(cos θ)
.
(47)
The multipole coefficients of the anisotropy power spectrum are given by
Cl = (4π)2 Z k2dkP (k)∆Tl(k, τ0)2
(48)
with P (k) the power-spectrum of the scalar fluctuations, assumed scale-invariant in
the sCDM model. The largest contribution to a given multipole Cl comes from those
wavelenghts such that l = k(τ0−τD), where τ0 is the conformal time at present and τD
the conformal time at decoupling. The average damping factor due to photon diffusion
upon the Cl's is given by an integral of e−2γ times the visibility function across the
last scattering surface [9]. It depends upon cosmological parameters, notably R, and
upon the recombination history. Approximately, and for a standard cold dark matter
model, we can take 2γ(d = 1) ≈ (l/1500)2. The relative change in the Cl's due to
the change in the photon-diffusion length, as we move down from frequencies where
depolarization is insignificant (d = 1) to lower frequencies (d << 1), is then given by
∆Cl =
Cl(d)
Cl(d = 1) − 1 ≈ exp(cid:16)(l/1500)2(1 − d)
(6 − 4d)
(cid:17) − 1
.
(49)
In figure 2 we plot ∆Cl (expressed as a percentage) at l = 1000 as a function of
frequency, for three different values of the magnetic field B∗ = 0.001, 0.01 and 0.1
Gauss. Once again, the graph for an arbitrary value of B∗ can be read from any of
these with an appropriate scaling of the frequency units. We have chosen to display
the effect at l = 1000, that will be accessible by the recently funded CMB satellite
experiments, MAP [24] and COBRAS/SAMBA [25].
3.2 Reduced quadrupole contribution
The depolarizing effect of a primordial magnetic field also changes the strength of the
quadrupole term SP around decoupling, and thus its incidence upon the anisotropy
of the CMB on small angular scales.
Indeed, the quadrupole anisotropy and the
polarization of the CMB at the time of recombination contribute to the presently
observed anisotropy through the following term of eq. (25)
∆SP (τ0) ≡ −
1
2
P2(cos θ)Z τ0
0
dτ eik cos θ(τ −τ0)g(τ )SP (τ )
.
(50)
12
This term is negligible for long wavelengths, those that dominate the lowest multipoles
of the present anisotropy, but becomes non-negligible on small angular scales (large
multipoles). Indeed, in the tight coupling approximation SP ∝ τC ∆0 and thus, barring
a very strong time-dependence of the scalar potential, the contribution of SP is well
below that of the monopole term, except for small wavelengths.
The depolarizing effect of a magnetic field modifies the value of SP around de-
coupling, compared to what it would have been if there were no magnetic field, and
then
∆SP (F ) = f (F )∆SP (B = 0)
(51)
with f as defined in eq.
(39). Recall that f ≈ 1 if ν >> νd while f ≈ 0.6 if
ν << νd. Thus, at frequencies such that the depolarizing effect of the magnetic field is
significant, the partial contribution of the quadrupole term SP to the total anisotropy
is reduced by a factor 0.6 compared to that at frequencies where depolarization is
unimportant. On small angular scales this could represent a decrease of the anisotropy
by a few per cent. The effect is opposite to that of the change in diffusion damping,
but is likely to be less significant on small angular scales.
3.3 Numerical estimate of the effect upon the anisotropy
In order to accurately ascertain the net effect of the depolarizing mechanism upon
the CMB anisotropy and to make definite quantitative predictions within a standard
cosmological model, we turn now to the numerical integration of the Boltzmann equa-
tions (4,5,6). We use the recently developed code CMBFAST[16], that integrates the
sources over the photon past light cone. Its starting point are the formal solutions
(25,26,27), where the geometrical and dynamical contributions are separately handled
to improve efficiency. As in our analytic estimates, when computing the evolution of
each Fourier mode we introduce in the code the Faraday rotation term with the angu-
lar dependence corresponding to the case where ~B has no component perpendicular
to ~k.
Figures 3 and 4 summarize the numerical calculation of the effect of depolarization
upon temperature anisotropy, in a standard cold dark matter model (sCDM).
The quantity plotted in Fig. 3 is l(l + 1)Cl, for the sCDM model without a
magnetic field and with a magnetic field and at frequencies such that F = 1, 4, 9,
corresponding to ν0 = νd, νd/2 and νd/3 respectively. Fig. 3 clearly shows that the
CMB anisotropy on small angular scales increases at frequencies where depolarization
is significant. This result indicates that the reduction in diffusion damping due to
depolarization is the dominant effect among the two opposite effects discussed in the
previous sections.
Fig. 4 displays the same results but expressed in terms of ∆Cl, the percentual
increase in Cl relative to the case without magnetic field. The monotonic curves
13
in the same figure, included for comparison purposes, correspond to the analytic
estimate of the effect of reduced diffusion damping, eq.(49). As expected, the effect is
larger on smaller angular scales (larger l). The numerical result approximately follows
the analytic estimate of the effect of the reduction in diffusion damping. The total
effect, however, does not increase monotonically with l. This can be understood as
a consequence of the nature of the subdominant quadrupole contribution SP , which
oscillates out of phase with the Cl's [8] (remember that SP ∝ ∆0), and is reduced by
depolarization through a factor f (F ).
It is also clear from Fig. 4 that the analytic result for the change in diffusion
damping due to depolarization overestimates the total effect at high l. This is because
the actual damping in the Cl spectra has two contributions, one from Silk damping
and the other due to cancellations in the integral across the last scattering surface
produced by the oscillations in the exponential and sources in equation (25). Only
Silk damping is reduced by the magnetic field, and that is why equation (49) slightly
overestimates the net effect.
The analytic and numeric calculations are in very good agreement around l ≈
1000. The frequency-dependence of ∆Cl at l = 1000 plotted in Fig. 2 fits very well
the analogous result after the full numerical integration of the Boltzmann equations.
We conclude that the depolarizing effect of the magnetic field results in an increase
of the anisotropy correlation function multipoles of up to 7.5% (for sufficiently low
frequencies) on small angular scales (l ≈ 1000), those that will be accessible by future
CMB satellite experiments such as MAP and COBRAS/SAMBA. The frequencies at
which the effect is significant, however, depend on the strength and coherence length
of the primordial magnetic field at the time of recombination.
Depolarization depending upon frequency, the effect might be difficult to separate
from foreground contamination. The relative change of the Cl's at l = 1000 is larger
than 2% on frequencies below 30 GHz (accessible to the first two channels in MAP),
if B∗ = 0.02 Gauss or larger. The first two channels in COBRAS/SAMBA being
at 31.5 and 53 GHz, the signal would reach a 2% level within this range only if B∗
is around or larger than 0.1 Gauss. However, COBRAS/SAMBA might reach out
to larger values of l, where ∆Cl is larger, and might thus have a sensitivity to the
depolarizing effect of B∗ comparable to MAP. In any case, both experiments will be
sensitive to a magnetic field around B∗=0.1 Gauss, and would thus at least be able
to place a direct constraint on B∗ comparable or better than the one obtained from
extrapolation of the nucleosynthesis bound.
Experiments searching CMB anisotropy and polarization at smaller frequencies,
which currently operate down to 5 GHz [26, 27], may play a significant role to detect
the depolarizing effect of a primordial magnetic field.
14
4 Conclusion
The CMB is expected to have a small degree of linear polarization. Several esti-
mates were made for the predicted polarization, both in the context of anisotropic
cosmological models [1, 28], as well as in isotropic and homogeneous cosmologies per-
turbed with either energy-density fluctuations or gravitational waves [18, 21, 29]. The
polarization of the CMB remains undetected, its upper limit being P < 6 × 10−5 [3].
A primordial magnetic field depolarizes the CMB radiation on those frequencies
that experience a significant amount of Faraday rotation around the time of decou-
pling.
[8]
to estimate the depolarizing effect of a primordial magnetic field across the last-
scattering surface, assuming a standard ionization history. The result is expressed by
eqs. (38,42,43) and is represented in Fig. 1. The CMB becomes significantly depo-
larized at frequencies around and below 30 GHz (B∗/0.01 Gauss)1/2, below which the
degree of polarization decreases quadratically with frequency. B∗ is the value of the
primordial field at a redshift z∗ = 1000, around recombination, likely to be 106 times
larger than an hypothetical cosmological magnetic field at present times.
In this paper, we have applied the analytic method developed in Ref.
The average depolarizing factor depends only upon the parameter F , as defined by
eq. (15), which represents the average Faraday rotation between collisions. We have
calculated the depolarizing factor d(F ), as given by equation (21), in the particular
case of a wavevector ~k k ~B. In a general case, the factor d depends upon the angle
between ~k and ~B. This dependence integrates away in average quantities, after the
stochastic superposition of all Fourier modes of the density fluctuations. The value
derived here for d is at most an underestimate of the average depolarizing effect,
which would eventually start to be significant at slightly larger frequencies. Our
derivation also assumed that Faraday rotation accumulates over the width of the last
scattering surface. If the primordial magnetic field is very entangled over that scale,
the depolarizing effect starts to be significant at smaller frequencies.
The depolarizing mechanism has a significant effect upon the anisotropy of the
CMB on small angular scales. On those angular scales and at frequencies such that
the depolarizing effect is large, the damping of anisotropies by photon diffusion is
reduced, which results in a significant increase of the anisotropy at a fixed angular
scale. Besides, depolarization reduces the contribution of the intrinsic quadrupole
anisotropy. Figure 2 displays the estimate for the percentual change of the anisotropy
power spectrum at l = 1000 due to the reduction in diffusion damping, as a function
of frequency and for different values of the primordial magnetic field at recombination.
We conclude that a primordial magnetic field increases the anisotropy of the CMB
by up to 7.5% at l ≈ 1000 in a standard CDM cosmology. The asymptotic strength of
the effect is independent of the intensity of the magnetic field, but the frequencies at
which it starts to be significant are those around and below 30 GHz (B∗/0.01Gauss)1/2.
15
Measurements of anisotropy and polarization at sufficiently low frequencies could
probe primordial magnetic fields in an interesting range.
Acknowledgements
DH and JH are grateful to Nathalie Deruelle and the DARC at Meudon for hospitality
while working on this project. The work of DH was partially supported by an EEC,
DG-XII Grant No. CT94-0004, and by CONICET. JH is grateful to the Royal Society
for a grant.
References
[1] M. Rees, Astrophys. J. 153,L1 (1968).
[2] S. Chandrasekhar, Radiative Transfer, Oxford University Press, Oxford, Eng-
land (1950).
[3] P. M. Lubin and G.F. Smoot, Astrophys. J. 273, L1 (1983).
[4] E. Milaneschi and R. Fabbri, Astron. Astrophys. 151, 7 (1985).
[5] A. Kosowsky and A. Loeb, "Faraday rotation of microwave background polar-
ization by a primordial magnetic field", astro-ph/9601055.
[6] For reviews on extragalactic magnetic fields see H. Asseo and H. Sol, Phys. Rep.
148, 307 (1987); P.P. Kronberg, Rep. Prog. Phys. 325, 382 (1994).
[7] B. J. Burn, Mon. Not. R. Astr. Soc. 133, 67 (1966).
[8] M. Zaldarriaga and D. Harari, Phys. Rev. D52, 3276 (1995).
[9] W. Hu and N. Sugiyama, Phys. Rev. D 51, 2599 (1995); Astrophys. J. 445, 521
(1995).
[10] P.J.E. Peebles and J.T. Yu, Astrophys. J. 162, 815 (1970).
[11] M. Turner and L. Widrow, Phys. Rev. D37, 2743 (1988); T. Vachaspati, Phys.
Lett. B265, 258 (1991); B. Cheng and A. Olinto, Phys. Rev. D50, 2421 (1994);
M. Gasperini, M. Giovannini and G. Veneziano, Phys. Rev. Lett. 75, 3796
(1995); D. Lemoine and M. Lemoine, Phys. Rev. D52, 1955 (1995); F.D. Mazz-
itelli and F.M. Spedalieri, Phys. Rev. D52, 6694 (1995).
[12] D. Grasso and H.R. Rubinstein, Astroparticle Phys. 3, 95 (1995); B.L. Cheng,
A. Olinto, D.N. Schramm and J.W. Truran, preprint astro-ph/9602055.
16
[13] J. Adams, U.H. Danielson, D. Grasso and H. Rubinstein, "Distortion of the
acoustic peaks in the CMBR due to a primordial magnetic field", astro-
ph/9607043.
[14] W. Hu, D. Scott, N. Sugiyama and M. White, Phys. Rev. D52, 5498 (1995).
[15] N. Kaiser, Mon. Not. R. Astron. Soc. 202, 1169 (1983).
[16] U. Seljak and M. Zaldarriaga, "A line of sight integration approach to CMB
anisotropies", astro-ph/9603033.
[17] G. Jungman, M. Kamionkowski, A. Kosowsky and D.N. Spergel, Phys. Rev.
Lett. 76, 1007 (1996). See also astro-ph/9605147 for a discussion in terms of the
parameters of the MAP and COBRAS/SAMBA planned experiments.
[18] J.R. Bond and G. Efstathiou, Astrophys. J. 285, L45 (1984).
[19] A. Kosowsky, Ann. Phys. 246, 49 (1996).
[20] B.T.J. Jones and R.F.G. Wyse, Astron. Astrophys., 149, 144 (1985).
[21] A. Polnarev, Sov. Astron. 29, 607 (1985).
[22] J. Silk, Astrophys. J. 151, 459 (1968).
[23] P.J.E. Peebles, "The Large Scale Structure of the Universe", Princeton Univer-
sity Press (1980).
[24] Visit the MAP home page at http://map.gsfc.nasa.gov.
[25] Visit the COBRAS/SAMBA homepage at http://astro.estec.esa.nl:80/SA-
general/Projects/Cobras/cobras.html.
[26] E. Fomalont, B. Partridge et al., Astrophys. J. 404, 8 (1993).
[27] Visit the CMB Jodrell Bank experiment home page at http://www.jb.man.ac.-
uk/sjm/cmb jb.html.
[28] G.P. Nanos, Astrophys. J. 232,341 (1979); M. Basko and A. Polnarev, Sov.
Astron. 24, 3 (1979); J. Negroponte and J. Silk, Phys. Rev. Lett. 44,1433 (1980);
B.W. Tolman and R.A. Matzner, Proc. R. Soc. London Ser. A 392, 1803 (1984);
B.W. Tolman, Astrophys. J. 290,1 (1985).
[29] J.R. Bond and G. Efstathiou, Monthly Not. R. Astron. Soc. 226, 655 (1987); E.
Milaneschi and R. Valdarnini, Astron. Astrophys. 162,5 (1986);R. Crittenden,
R.L. Davis and P.J. Steinhardt, Astrophys. J. 417, L13 (1993); D.D Harari and
M. Zaldarriaga, Phys. Lett. B 319, 96 (1993); R.A. Frewin, A.G. Polnarev, P.
17
Coles, Monthly Not. R. Astron. Soc. 266, L21 (1994); K. L. Ng and K.W. Ng,
Phys. Rev. D 51, 364 (1995) and Astrophys. J. 445, 521 (1995); D. Coulson,
R.G. Crittenden and N.G. Turok, Phys. Rev. Lett. 73, 2390 (1994) and Phys.
Rev. D52, 5402 (1995).
18
B
=0.001G
*
B*B*
=0.01G
=0.1G
1
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
_
D
0
0
10
20
40
50
60
30
0 (GHz)
ν
Figure 1: The average depolarizing factor ¯D as a function of the CMB frequency ν0.
The corresponding figure for an arbitrary value of B∗ is identical to any of these after
a scaling of the frequency units, proportional to B1/2
.
∗
19
8
7
6
5
∆ C
(%)
l
4
3
2
1
0
B*=0.001G
B*=0.01G
B*
=0.1G
(l = 1000)
0
10
20
40
50
60
30
0 (GHz)
ν
Figure 2: Analytic estimate of the percentual change due to reduction in diffusion
damping of the l = 1000 anisotropy correlation function multipoles as a function
of the CMB frequency for different strengths of the primordial magnetic field at
recombination. The corresponding figure for arbitrary B∗ can be obtained from any
of these after a scaling of the frequency units, proportional to B1/2
∗
20
π
2
/
0
1
0
1
x
l
C
)
1
+
l
(
l
4
3.5
3
2.5
2
1.5
1
0.5
0
0
sCDM
F=1
F=4
F=9
200
400
600
800
1000
1200
1400
1600
l
Figure 3: Numerical integration for the multipoles of the anisotropy correlation func-
tion in a standard CDM model without a primordial magnetic field (F = 0), and
with F = 1, 4, 9, which correspond to ν0 = νd, νd/2, νd/3 respectively, with νd ≈
27 GHz (B∗/0.01Gauss)1/2.
21
F=1
F=4
F=9
16
14
12
10
8
6
4
2
0
∆ C
(%)
l
-2
0
200
400
600
1000
1200
1400
1600
800
l
Figure 4: Numerical result for the percentual change of Cl as a function of l relative
to its value without magnetic field in a standard CDM model. The monotonic curves
also shown for comparison purposes correspond to the analytic estimate of the effect
of reduced diffusion damping.
22
|
astro-ph/9408020 | 1 | 9408 | 1994-08-05T18:02:42 | Are BL Lacertae Objects Beamed QSO Remnants? | [
"astro-ph"
] | This paper considers the hypothesis that BL Lacertae objects (BLLs) are the beamed remnants of Quasi Stellar Objects. The hypothesis explains why BLLs do not undergo the strong evolution seen in other active galactic nuclei since it naturally predicts that the space density of BLLs should increase with cosmic time, as shown by recent observations. Numerical models reproduce, with reasonable parameters, the known redshift and magnitude counts of BL Lac objects. It is assumed that radio-quiet as well as radio-loud quasars are capable of generating jets but that jets are snuffed in young radio-quiet objects and only emerge in aged ones. I argue that the observations allow this assumption. | astro-ph | astro-ph | Are BL Lacertae Objects Beamed QSO Remnants?
E.F. Borra
Centre d'Optique, Photonique et Laser, Observatoire Astronomique du Mont
Département de Physique, Université Laval, Québec, Canada G1K 7P4
Mégantic,
SUBJECT HEADINGS: BL Lacertae objects -- galaxies: jets -- galaxies: nuclei
-- radio sources: galaxies
To appear in: COMMENTS ON ASTROPHYSICS Vol. 17 No 5
4
9
g
u
A
5
0
2
0
8
0
4
9
/
h
p
-
o
r
t
s
a
ABSTRACT
2
This paper considers the hypothesis that BL Lacertae objects (BLLs) are the beamed
remnants of Quasi Stellar Objects. The hypothesis explains why BLLs do not undergo the
strong evolution seen in other active galactic nuclei since it naturally predicts that the space
density of BLLs should increase with cosmic time, as shown by recent observations.
Numerical models reproduce, with reasonable parameters, the known redshift and
magnitude counts of BL Lac objects. It is assumed that radio-quiet as well as radio-loud
quasars are capable of generating jets but that jets are snuffed in young radio-quiet objects
and only emerge in aged ones. I argue that the observations allow this assumption.
3
1. INTRODUCTION
It is increasingly accepted that the objects that populate the high-energy extragalactic
"zoo" are Active Galactic Nuclei (AGNs), that they are manifestations of the same basic
phenomenon and that they all will eventually be unified within energetic, geometrical or
evolutionary schemes. Among AGNs, the BL Lacertae objects (hereafter referred to as
BLLs) are particularly puzzling "beasts", since they exhibit extreme characteristics among
the AGNs. A consensus has emerged that these characteristics are due to relativistic
beaming1. Over the past several years, evidence has gradually built up showing a puzzling
behavior that perhaps differentiates them the most from other AGNs: They do not show
the large decrease in comoving density with cosmic time exhibited by other AGNs. On the
contrary, Morris et al.2 recently found evidence of negative evolution, namely that the
space density of BLLs actually increases with cosmic time.
Borra 3 has proposed that BL Lacertae objects are the beamed remnants of Quasi
Stellar Objects (QSOs).This paper was probably not noticed at that time by workers active
in AGN research because the hypothesis was used in a cosmological test rather than
presented as the theme of the paper. What prompted this hypothesis was that it naturally
explains why BLLs do not undergo the strong evolution seen in other AGNs; indeed
negative evolution is a natural consequence of it. Borra 3 made the assumption that all
Quasi Stellar Objects (QSOs), radio-quiet (RQQ) as well as radio-loud (RLQ), are capable
of generating jets but that jets are snuffed in young RQQs and only emerge in aged
objects. This assumption was necessary to obtain good agreement between the space
densities of the objects concerned. Prima facie, this assumption runs counter the prevailing
belief that radio-quiet and radio-loud quasars are distinct objects hosted in different types
of galaxies (spirals versus ellipticals) located in different environments (poor versus rich
clusters of galaxies). On the other hand, radio-loud and radio-quiet QSOs have remarkably
4
similar spectra and powers, so that it is natural to assume that the same basic mechanisms
are at work and that, for some reason, energy in the radio region is either absorbed or not
generated in the RQQs. Section III further addresses this issue.
In this paper, I reexamine the hypothesis in the light of 10 years of advances
in our knowledge of AGNs. I assume that all QSOs have jets but that the jet is quenched in
the young RQQs. As the QSO ages, the quenching mechanism weakens allowing the jet to
emerge. I shall also argue that the observational evidence on the environmental statistics of
the RLQs and RQQs does not rule out the hypothesis.
2. NUMERICAL MODELS
The basic hypothesis assumes snuffing that decreases in strength with age but does
not identify a specific snuffing model. Let us simply assume that all AGNs are capable of
generating a jet and that material surrounding the nucleus is somehow responsible for
snuffing it in radio-quiet QSOs. This material may be in a leaky shell, or torus,
surrounding the central powerhouse, with holes or weak spots that allow the emergence of
a jet only if it is aimed at one of them. With a toroidal geometry, the jet can be directed
along directions not necessarily aligned with the axis of symmetry of the torus. We shall
not make any assumption on the snuffing mechanism itself but assume that the material
interferes somehow with the mechanism that either accelerates the material or delivers it to
the outer extended lobes. It seems reasonable to assume that the quantity of material
surrounding the nucleus eventually decreases as the QSO ages, so that it becomes easier
for a jet to emerge from an older quasar.
One could in principle evolve the observed luminosity function of QSOs from high
to low redshifts, compute the rate of formation of remnants and compare the predictions to
the observed luminosity function of the local BL Lac objects. However, in the absence of
a quantitative theory, this is not a particularly fruitful approach, and we shall instead
5
simply find a local luminosity of the QRs that is compatible with the known evolution of
the luminosity function of QSOs and the local luminosity function of BLLs. We then shall
see whether relativistic beaming can give a luminosity function that reproduces the
observations.
Urry & Shafer 4 have studied the effects of relativistic beaming on a luminosity
function. Following 4, we shall assume a power law for the isotropic unbeamed
component of the luminosity function of the local QSO remnants (QR):
F QR(L,z=0) dL = KL-B dL for L1<L< L2
= 0
for L<L1 or L<L2
.
(1)
As Urry & Shafer 4 point out, the upper and lower luminosity cutoffs, although
unrealistic, approximate the fact the luminosity function must turn down at some
luminosity. They also show that relativistic beaming changes the original power law
luminosity function in a way that can be approximated by two power laws, the high
luminosity end having index ~ B, and the low luminosity end having a flattened exponent
~ (p+1)/p, where p depends on the shape of the spectrum, the structure of the jet and the
frequencies being compared; it usually is in the range 3<p<5. We keep the notation of 4
and the reader is referred to that paper for more details.
If QSOs are long lived, the local density of their remnants should be essentially
equal to the density of quasars at z = 2 and
M urF QR(M,z=0.0)dM =
M lr
M uqF QSO(M,z=2.0)dM ,
M lq
(2)
6
where F QR(M,z) and F QSO(M,z) are the luminosity functions of QSO remnants and
QSOs. This approximates the actual situation since QSOs are probably continuously
formed rather than in a burst at z= 2.0 as implied by Equation 2. The Mu and Ml upper
and lower magnitude cut-offs take into account the fact that the luminosity functions must
turn down at some magnitudes and that very faint QSOs probably do not evolve
significantly to our epoch. The unbeamed luminosity function is further restricted by the
requirement that local QSO remnants are inconspicuous and therefore much less luminous
than QSOs. Therefore we must have, at least for MV <-22
F QR(M,z=0) << F QSO(M,z=0).
(3)
We now feel free to experiment with luminosity functions of the form given by Equation 7
in 4 subject only to the conditions set by Equations 2 and 3.
The redshift distribution given in 2 has two peculiarities: It shows a very sharp
drop at z < 0.2 and a more gradual, but still rapid, decrease for z > 0.2. A narrow
luminosity function having the appropriate mean luminosity, combined with negative
evolution of the space density of the objects can reproduce this. Such a luminosity
function naturally arises from a beamed population of low-luminosity galactic nuclei
having a power-law luminosity function with a steep index and small luminosity range.
The small luminosity range may be real and approximate a sharply peaked luminosity
function (e.g., Gaussian-like). One also may justify the low-luminosity cut-off by the fact
that a faint object no longer looks like a BL Lac object since there is a strong contribution
to the spectrum from the host galaxy; while the high luminosity drop-off approximates the
fact that the luminosity function must turn down at some luminosity. It is also important
7
that the unbeamed objects be faint since too high a mean luminosity would shift the peak
of the redshift distribution to high redshifts.
We obtain a good fit to the observations with a model having an unbeamed power-
law luminosity function with index B= 3.0 and -11 < MV > -13. This luminosity function
respects the requirements set by Equation 3, that the unbeamed component be
inconspicuous. We obtain the normalization factor K from Equation 2, where we use the
luminosity function F QSO(M,z) determined in 5 for q0=0.5 and H0=50 Km/sec/Mpc. We
use Mlq=-22 and Muq=-30, thus assuming that QSOs fainter than MB=-22 do not evolve
enough to z =0 to contribute significantly to the local population of QRs. Although the
luminosity functions in 5 are truncated at MB=-23 they show strong evolution at MB = -23,
and are compatible with significant evolution at MB=-22. There are sources of uncertainty
in K, arising from the determination of the luminosity function of QSOs, from
uncertainties in the cosmological parameters and from the lower magnitude cutoff. We use
a relativistic Lorentz factor g = 10, which is a reasonable average value 6, and p= 4 for the
exponent in Equation 1, a value appropriate for our problem. The beamed luminosity
function is computed assuming viewing angles 0.0º < q
< 5º. The beamed luminosity
function has a steep power law from -26<MV<-24 joined to a shallow one for -24<MV<-
21.5, compatible with the known luminosities of BL Lac objects. The MV<-21.5 low-
luminosity cutoff justifies truncating the viewing angle at 5º for the luminosity would then
be less than the luminosity of the host galaxy (BLLs appear in bright host galaxies) and the
object would no longer be identified as a BLL. We compute the redshift distribution from
the usual cosmological relation given by
dN
dW dz
=
m1F BL(M,z)dV
dz
m0
dm ,
(4)
where F BL(M,z) is the beamed differential luminosity function (per unit magnitude), W
the
surface area and dV(H0,q0,z) the cosmological volume element. The z dependence of the
luminosity function is given by
8
F BL (M,z) = F BL (M,z=0)(1+z)b ,
(5)
following the density evolution law determined in 2. We shall use b = -5.5,
the best fit
values determined by them for density evolution. The sample in 2 is X-ray selected and,
in principle, it would be desirable to base our computations fully on the X-ray luminosity
function of QSOs. In practice, because one needs the luminosity functions of QSOs at z =
0 and z = 2.0 (Equations 2 and 3) and because they are much better known in the optical,
especially at low fluxes and high redshifts, it is preferable to use optical luminosity
functions. For our computations, we shall use the mV distribution given in 2 for their
sample and a value of ml = 20 in Equation 4 as indicated from their mV magnitude
distribution. We use H0=50 Km/sec/Mpc and q0= 0.5. The model described gives the
redshift distribution of Figure 1 (dashed line), where the histogram gives the
observations in 2. We can see that the model fits well the main features of the redshift
distribution in 2 and it is reasonable to assume that the observations sample the computed
parent distribution of Figure 1. The computed z-distribution rises sharply, peaks at the
same redshift and rapidly drops to undetectability for z>0.8.
We have plotted in Figure 2 the integrated counts as a function of magnitude
obtained from
dN
dW dm
=
z1F BL(M,z)dV
dz
z0
dz ,
(6)
9
where the integrand is the same as in Equation 6 and all symbols and constants are the
same. We have plotted in the same figure the estimates of the optical integrated counts of
BLLs obtained in 6; the arrows indicate either lower or upper limits. We can see that the
theoretical counts also are in agreement with the observations.
Our model is a simple one since it considers the contribution of the beam alone,
assumes that all sources have the same value of g , and uses a power-law luminosity
function. It does not take into account that, in the optical, there also is a contribution from
the light of the host galaxy. Padovani and Urry 6 have considered this case and find that it
does not change the basic conclusions of the simple model. Their modified model, that
uses two additional free parameters, still predicts a double power law with slope ~B above
the break but with slope steeper than (p+1)/p below the break. They have also extended
their basic model 7 to more general luminosity functions and the case where there is a
distribution of Lorentz factors g . The uncertainties in the observations and some of the
assumptions that are made in the computations do not justify the introduction of additional
free parameters into our models nor the consideration of other types of luminosity
functions.
1 0
3. DO RADIO-QUIET QUASARS HAVE SNUFFED JETS?
a) Snuffing and energy distributions
RLQs and RQQs have remarkably similar spectral energy distributions and powers
at all wavelengths except radio wavelengths; it is therefore logical to assume that they
actually are similar objects and that the remarkable differences in their radio fluxes are due
to absorption of the radio power or snuffing of the radio emission mechanism. Another
attractive model invokes relativistic beaming and different viewing angles to explain the
difference but fails (8, 9, 10). A number of absorption processes have been considered
(e.g., synchrotron self-absorption, free-free absorption) but fail to account for the bimodal
distribution of radiopower in quasars 8. The remaining viable hypothesis, also suggested
by 9, is that some mechanism prevents a relativistic jet from fully developing and
generating the relativistic electrons responsible for the radio emission. Snuffing naturally
explains the remarkable similarity of the 2 types of quasars at all wavelengths but the radio
region, as well as the bimodality of the radio emission. It accounts for the observations of
11 who find that lobe dominated radio quasars have spectra that show deep minima in the
millimeter region, indicating thus that their radio emission is not a continuation of the
optical-infrared spectrum but that they contain normal RQQ cores with additional emission
from a morphologically unrelated radio source.
b) Changes in the number ratios (RLQ/RQQ) with z and magnitude
Under our hypothesis radio-jets emerge as the RQQs age, leading to the prediction
that the fraction of radio-loud objects should increase with age, hence decreasing z.
Therefore one could, in principle, use radio observations of optically selected QSOs to
1 1
verify these predictions. However, an observational test is difficult because we simply do
not know how the snuffing factor varies with age; hence we cannot quantitatively predict
how the ratio RLQ/RQQ varies with redshift. Furthermore, if the jets emerge when the
objects are faint, they will not longer be identified as quasars, rendering the comparison
meaningless. The redshift distribution of the X-ray selected BLLs gives some clue, since
the counts peak at z~0.3, first indicating the cosmic time at which most radio jets emerge
and second that jets emerge mostly among very evolved and very faint objects that would
not be recognized as quasars because of their very low luminosities. Comparison to the
observations must therefore be tempered by these realizations. On the other hand, it is
legitimate to assume that, although jets emerge very late in the evolution of most RQQs,
the emergence of the jets may occur earlier in some objects and there may therefore be a
tendency for the fraction of RLQs to increase with cosmic time. Whether this should be
detectable and is detected with the surveys presently available is another matter that we
shall now consider. This preamble should therefore make it clear that our hypothesis
would stand even if the test of this subsection fails.
There are very few surveys of optically selected QSOs. The most recent and largest
ones are by 10 and 12. They confirm the bimodality of the radio emission in QSOs.
Visnovsky et al. 12, who also use the data in 10 find strong evidence that in optically
selected QSOs the radio-loud fraction increases with decreasing redshift, in agreement
with our prediction but also that the ratio decreases with fainter limiting magnitude (albeit
with weaker statistical significance). Under our hypothesis, the ratio loud/quiet should
increase with decreasing redshift in optically selected samples, because at a given
magnitude, as the universe ages, one collects an increasing number of older, originally
brighter QSOs.
Because quasars get fainter as they age, one would expect, prima facie, that the
ratio should also increase with faintness; however, this prediction may be too simple-
minded. If brighter sources have a higher probability of developing a jet either simply
1 2
because the jet is stronger and can get through a snuffing shield more easily or because
the snuffing shield is ablated faster by the stronger radiation from the brighter core, this
will cause the RLQ/RQQ ratio to increase with brightness for middle-aged objects. This
picture finds some support from 12 where correlation between radio and optical
luminosity for the radio-weak sources is found, as though more luminous objects have a
stronger embryonic radio-jet trying to break through and giving some radio flux prior to
snuffing. Optically selected QSOs are found in surveys that are magnitude limited, the
absolute flux decreasing with z. They therefore include only relatively bright (typically MB
<-23), presumably middle-aged objects, missing thus the very aged objects that have all
developed jets, because they are too faint. Visnovsky et al. 12, mostly discuss QSOs with
MB<-23 and 0.1<z< 3. It is possible that in the range of brightness and redshifts
discussed in 12, the brightness effect dominates over the aging effect thus explaining why
the ratio RLQ/RQQ increases with brightness.
I do not claim that this discussion proves that RQQs have snuffed jet but
only that the observations do not rule this hypothesis. Section 3b is particularly speculative
but , on the other hand, it could be dropped without affecting our discussion. The stronger
argument in the favor of the snuffing hypothesis is by absentia of an alternative.
4. DISCUSSION
Prima Facie, the weakness of this work comes from the assumption that RLQs and
RQQs are basically the same objects, which runs counter the presently held belief that
these are different objects, residing in different environments (RLQs in elliptical galaxies
and rich clusters, RQQs in disk galaxies and poor clusters). However, first, this is not a
firmly established fact since the statistics are poor and the earthbound observations have
too marginal a resolution to unambiguously reveal the morphologies of the host galaxies.
The uncertainties in the properties of the host galaxies have been studied 13, reaching the
1 3
conclusion that, although they favor the hypothesis that RLQs are in elliptical hosts and
RQQs in disks, the uncertainties are considerable. A second consideration comes from the
fact that the meager observational data are obtained at z < 0.6 and one has no information
about the morphologies of the host galaxies of RLQs and RQQs at higher redshifts.
Third, recent spectroscopic and morphological observations coming from an HST key
project carried out by Butcher, Dressler, Oemler and Gunn (reported in 14) of 2 clusters at
z = 0.4 find a much greater proportion of disk galaxies in those clusters than there are at
low z. They also find that they often appear a bit disturbed or irregular in morphology.
Dressler concludes that some mechanism, presumably interactions and mergers, is
responsible for the disappearance of the spiral galaxies in clusters since z = 0.6. This is
relevant to this work since it allows us to assume that disk galaxies hosting RQQs at high z
have changed morphology to ellipticals at low z. In conclusion, the observational evidence
therefore allows us two assumptions: First that at z 2 any Hubble type can host RQQs as
well as RLQs; even though RLQs may favor ellipticals and RQQs disk galaxies at low
redshifts. One may of course wonder why it should be so; one can only speculate that this
is due to a younger universe and younger environments. Second, the HST observations
also allow another plausible assumption; that the host spirals at high z have now become
ellipticals. The HST observations makes this assumption the most plausible one.
Until recently, it was generally believed that the hosts of BLLs were exclusively
elliptical galaxies. However, there is now conclusive evidence that at least some BLLs are
in disk galaxies. For example, Abraham, McHardy, & Crawford 15 have carried out an
optical imaging survey of BLL host galaxies, concluding that out of 6 classifiable hosts 2
are disk systems. One may be skeptical of this classification since there have been
previous controversies regarding the nature of the hosts of some BLL (e.g. 16 , and
references therein). However, an unpublished high-quality HST image of the BLL
PKS1413+135 (McHardy, private communication) unambiguously shows it to be a in a
disk galaxy. As 15 points out, the discovery of disk hosts is in disagreement with the
1 4
standard model that assumes that all BLLs are hosted in elliptical galaxies . This is
however allowed by our hypothesis.
There have been proposals to relate BLLs to RLQs (2, 17 ). With respect to our
hypothesis, these have the advantage of not having to assume a snuffing mechanism since
they relate only radio-loud objects. However, because the ratio RQQ/RLQ is about 10 at z
=2, it is more difficult to reproduce the observed space density of BLLs at low z and
obtain the sort of good agreement with observations seen in Figures 1 and 2. It is of
course possible to reduce the gap by assuming parent quasars having Mv >-22 , by
decreasing the relativistic g factor or by taking the lower limits to the space densities of
BLLs from 2. However, this fine tuning must be contrasted with the agreement that we
obtained with more reasonable estimates. Finally, the assumption that RLQs are
exclusively in elliptical galaxies and are the progenitors of BLLs is in conflict with the
presence of some BLLs in disk galaxies.
There obviously are details that our basic hypothesis does not explain. For
example, one may want to explain the relation between OVV quasars (radio quasars with
flat radio spectra and highly polarized QSOs) and BLLs and their spectral differences (e.g.
18, 19 ). This is beyond the scope of this work, but it is reasonable to assume that the
spectral differences are due to different evolutionary ages or environments . Assuming that
BL Lacs are more evolved objects than quasars, it is not surprising that there are spectral
and morphological differences between them and similar objects they may be evolutionary
related to. It is the reverse that would be surprising. In particular, our hypothesis is not in
conflict with the assumption that the parent population of BL Lacs is a subset of FR I. We
can assume that although strong and young RLQs are FR II, the weaker and more evolved
BL Lacs are FRI. Perhaps it is power that decides whether an object is FR I or FR II.
Finally, it is interesting to notice that the leaky-shell, leaky-torus model may also
explain why BLLs have very weak broad emission lines. If the shell contains enough
absorbing material (e.g., dust), and if it surrounds the broad-line region, it can greatly
1 5
absorb the flux from the region that forms the broad lines. Although the jet punches a hole
through the shell, the hole may be large enough to let the jet through but too small to allow
a significant flux from the emission-line region to give strong broad emission lines. In a
classical QSO, the viewing angle would be such that the observer sees the nucleus and
broad-line region through a large hole or at a small angle with respect to the axis of
symmetry of the torus. Old age may explain why narrow lines are weak. While a young
QSO should have a healthy narrow-line region, since it is in a young galactic environment
with plenty of gas, a BLL does not have a strong narrow line region since gas is depleted
in its older environment.
5. CONCLUSION
We have reexamined, after ten years of advances in AGN research, the hypothesis
advanced by Borra 3 that BL Lacertae Objects are the beamed remnants of Quasi Stellar
Objects. This hypothesis explains why BLLs do not undergo the strong evolution seen in
other AGNs. It naturally accounts for the observations of 2 that show that the space
density of BL Lacertae objects increases with cosmic time; in the opposite sense of the
evolution of other Active Galactic Nuclei. Constrained by the luminosity function of QSOs
at z= 2 and the requirement that their remnants at z < 1 be faint, we find that a power-law
luminosity function that is relativistically beamed can give redshift and magnitude counts
compatible with the observations. Our successful model has free parameters that were
adjusted to yield a good fit to the observations, and the solution obtained may not be
unique. On the other hand, the values used for these free parameters are all reasonable and
compatible with the observations.
To obtain agreement between the observed space densities, we make the additional
assumption that all QSOs, radio-quiet as well as radio-loud, are capable of generating jets,
1 6
but that jets are snuffed in the young RQQs and only emerge in old age, thus implicitly
assuming a unification of RLQs and RQQs.
The meager existing statistics of the classification of the host galaxies of BLLs seem
to indicate that they favor elliptical (60%) rather than disk galaxies (30%). Prima facie, one
would have to conclude that this is only allowed by our hypothesis either if elliptical can
form RQQs as well as RLQs at z 2 or if a large fraction of QSO-hosting spirals can
evolve into elliptical. Recent HST observations seem to favor the later.
Although this paper does not prove that BLLs are quasar remnants, it shows that
the hypothesis is compatible with our present knowledge of the statistics of QSOs and BL
Lac objects, its outstanding success being its natural prediction that the space density of
BLLs should increase with cosmic time, and that it should be taken seriously. It does not
say anything regarding the unbeamed component of BL Lac objects, except that they
should be optically inconspicuous objects and have FR I morphology.
Aknowledgments
1 7
This research was supported by the Natural Sciences and Engineering Research
council of Canada and the FCAR program of the province of Québec. I wish to thank Dr.
I. McHardy for sending me an unpublished HST image of PKS1413+135.
1 8
REFERENCES
1 Blandford, R., & Rees, M.J., 1978, in Proc. Pittsburgh Conf. on BL Lac Objects, ed.
A.N. Wolfe (Pittsburgh: University of Pittsburgh Press), p. 328
2 Morris, S.L., Stocke, J.T., Gioia, I.M, Schild, R.E., Wolter, A., Maccacaro, T., &
Della Ceca, R. 1991, ApJ 389, 49
3 Borra E.F, 1983 ApJ 273, L55
4 Urry, C. M., & Shafer, R.A. 1984, ApJ 280,569
5 Hartwick, F.D.A, & Schade, D. 1990, ARA&A 28, 437
6 Padovani, P., & Urry, C.M. 1991 ApJ 368, 373
7 Urry, C. M., & Padovani, P.1991, ApJ 371, 60
8 Condon, J.J., O'Dell, S.L., Puschell, J.J., & Stein, W.A.. 1981, ApJ. 246, 624
9 Strittmatter, P.A., Hill, P., Pauliny-Toth, I.I.K., Stepe, H., & Witzel, A. 1980, A&A
88, L12
10 Kellerman, K.I., Sramek, R., Schmidt, M., Shaffer, D.B., & Green, R. 1989, AJ 98,
1195
11 Antonucci, R., Barvainis, R. & Alloin, D. 1990, ApJ 353,416
12 Visnovsky, K.L., Impey, C.D., Foltz, C.B., Hewett, P.C., Weymann, R.J., &
Morris, S.L. 1992, ApJ 391,560
13 Abraham, R.G., Crawford, C.S., & McHardy, I.M. 1992, ApJ 401, 474
14 Dressler, A. 1992 preprint of an invited talk at the Milan conference on observational
cosmology.
15 Abraham, R.G., McHardy, I.M., &Crawford, C.S. 1991, MNRAS 252, 482
16 Romanishin, W. 1992, ApJ 401, L65
17 Vagnetti, F., Giallongo, E. &Cavaliere, A. 1991, ApJ 368, 366
18 Guilbert, P.W. Fabian, A.C. & McCray, R. 1983, ApJ 266, 466
19 Worrall, D.M. & Wilkes, B.J. 1990, ApJ 360, 396
1 9
FIGURE CAPTIONS
2 0
r
e
b
m
u
N
1 0
8
6
4
2
0
0
0 . 2
0 . 4
0 . 6
0 . 8
1
Redshift
Figure 1: The dashed line gives the redshift distribution obtained from the relativistically
beamed model described in the text. The counts are in bins of D z = 0.05. The histogram
shows the observations of 2.
2 1
1
-
g
a
m
2
-
g
e
d
)
V
<
N
(
g
o
L
- 1
- 1 . 5
- 2
- 2 . 5
- 3
- 3 . 5
- 4
- 4 . 5
1 2
1 4
1 6
1 8
2 0
2 2
2 4
2 6
V
Figure 2: Integrated counts (per square degree) from the relativistically beamed model
described in the text. The observed counts are from 6; and the symbols have the same
meaning as in that reference. The arrows indicate upper or lower limits.
|
astro-ph/9704202 | 1 | 9704 | 1997-04-21T15:17:14 | Using the Fundamental Plane to Estimate the Total Binding Mass in A2626 | [
"astro-ph"
] | We use fundamental plane (FP) distance estimates to the components of the double cluster A2626 (cz~17,500 km/s) to constrain cluster kinematics and estimate total binding mass. The FP coefficients for a sample of 24 early type and S0 cluster members (alpha=1.30+/-0.36 and beta=0.31+/-0.06) are consistent with others reported in the literature. We examine the Mg_b distributions within both subclusters and find them to be indistinguishable. Lacking evidence for stellar population differences, we interpret the FP zeropoint offset (\log(D_B/D_A)=-0.037+/-0.046, where D_{cl} is distance to subcluster cl) as a measure of the distance difference. This measurement is consistent with the subclusters being at the same distance, and it rules out the Hubble flow hypothesis (distances proportional to velocity) with 99% confidence; analysis of the subcluster galaxy magnitude distributions rules out Hubble flow at 93% confidence. Both results favor a kinematic model where the subclusters are bound and infalling. We estimate the total cluster binding mass by modelling the subcluster merger as radial infall. The minimum possible total binding mass is 1.65 times higher than the sum of the standard virial masses, a difference statistically significant at the ~3sigma level. We discuss explanations for the inconsistency including (1) biases in the standard virial mass estimator, (2) biases in our radial infall mass estimate, and (3) mass beyond the virialized cluster region; if the standard virial mass is significantly in error, the cluster has an unusually high mass to light ratio (~1000h). Because observational signatures of departures from radial infall are absent, we explore the implications of mass beyond the virialized, core regions. (abridged) | astro-ph | astro-ph |
USING FUNDAMENTAL PLANE DISTANCES to ESTIMATE
the TOTAL BINDING MASS in ABELL 26261
Departments of Physics and Astronomy, University of Michigan, Ann Arbor, MI 48109
Joseph J. Mohr
Department of Physics and Astronomy, Dartmouth College, Hanover, NH 03755
Gary Wegner
Accepted for publication in The Astronomical Journal
ABSTRACT
We use fundamental plane (FP) distance estimates to the components of the double
cluster A2626 (cz ∼ 17, 500 km/s) to constrain cluster kinematics and estimate total
binding mass. We employ deep R band CCD photometry, multi–object spectroscopy,
and software designed to account for seeing effects to measure the FP parameters Re,
σ, and hµei for 24 known early type and S0 cluster members. The FP coefficients from
this sample (α = 1.30 ± 0.36 and β = 0.31 ± 0.06) are consistent with others reported
in the literature.
We examine the Mgb equivalent width distributions within both subclusters and
find them to be indistinguishable. Lacking evidence for stellar population differences,
we interpret the FP zeropoint offset between the two subclusters as a measure of the
distance difference. We find log(DB/DA) = −0.037 ± 0.046, where Dcl is the distance
to subcluster cl. This measurement is consistent with the subclusters being at the
same distance, and it rules out the Hubble flow hypothesis (distances proportional to
velocity) with 99% confidence; analysis of the subcluster galaxy magnitude distributions
rules out Hubble flow at 93% confidence. Both results favor a kinematic model where
the subclusters are bound and infalling.
We estimate the total cluster binding mass by modelling the subcluster merger as
radial infall. The projected separation, the line of sight velocity difference and the line
of sight separation constrain the cluster mass; the minimum possible total binding mass
is 1.65 times higher than the sum of the standard virial masses, a difference statistically
significant at the ∼ 3σ level. We discuss explanations for the inconsistency including
(1) biases in the standard virial mass estimator, (2) biases in our radial infall mass
estimate, and (3) mass beyond the virialized cluster region; if the standard virial mass
is significantly in error, the cluster has an unusually high mass–to–light ratio (∼ 1000h).
Because observational signatures of departures from radial infall are absent, we explore
the implications of mass beyond the virialized, core regions.
1Observations reported here were obtained at the MDM Observatory, a facility jointly operated by the University of Michigan, Dartmouth
College, and the Massachusetts Institute of Technology.
1
1.
Introduction
Galaxy cluster masses are clearly of cosmological
significance, and examples abound: (1) the differences
in cluster evolution as a function of mass provide con-
straints on the power spectrum (Lacey & Cole 1993,
1994), (2) comparisons of the amounts of luminous
and non–luminous matter in clusters provide clues
to the nature of dark matter and the efficiency of
galaxy formation (David, Jones & Forman 1995), and
(3) the cluster mass–to–light ratio and baryon frac-
tion constrain the cosmological density parameter Ω0
(e.g. Ramella, Geller & Huchra 1989, White et al.
1993, Mohr et al. 1996).
Cluster mass studies tend to focus on the central
regions where virial equilibrium is a more accurate
approximation, and the weak lensing signals are eas-
iest to detect. Here we apply a method to measure
the mass beyond the virialized region in the double
cluster A2626. We utilize recent improvements in the
understanding of the fundamental plane (FP) to con-
strain the total cluster binding mass by revisiting the
subcluster radial infall model (Beers, Geller & Huchra
1982). Specifically, measuring FP distances (with
an accuracy of 15% to 25%) to a reasonably large
sample of cluster galaxies can yield a very accurate
(sub)cluster distance (Lucey & Carter 1988, Lynden–
Bell et al. 1988). For two bound subclusters, the
line of sight separation, the line of sight velocity dif-
ference, and the projected separation provide enough
information to estimate the total cluster binding mass
in cases where a radial infall model is appropriate.
Although A2626 is rather distant, it is well suited
for this analysis because simply demonstrating that
the two subclusters are bound and infalling has inter-
esting implications. A straightforward application of
the radial infall model, using the (rest frame) line of
sight velocity difference and the virial mass estimates
within each subcluster, rules out bound and infalling
merger models (Mohr, Geller & Wegner 1996; here-
after MGW96); the virial masses are not large enough
to produce the apparent velocity difference. Thus, FP
distances used to determine whether or not the sub-
clusters in A2626 are bound and infalling serve as an
independent test of the virial estimators, or alterna-
tively, as a means of measuring the mass beyond the
virialized region of the cluster.
Possible pitfalls to this approach include (1) dis-
tance biasing FP zeropoint (or shape) differences be-
tween subclusters, (2) deviations from radial infall,
2
and (3) significant interactions between the virial-
ized regions of the two subclusters. FP variation
with cluster environment is an area of active research
(e.g. Guzman et al. 1992, Worthey, Trager, & Faber
1996, Jørgensen, Franx & Kjaergaard 1996; hereafter
JFK96); current indications are that distance biases
are smaller than 5% for ellipticals within clusters and
are related to observable variations in the stellar pop-
ulations within cluster galaxies. As discussed below,
subcluster pairs suffering from pitfall numbers 2 and 3
can be identified through corresponding observational
signatures. We argue that the radial infall model and
FP distances provide a promising method for deter-
mining cluster masses on scales larger than the viri-
alized region in low redshift clusters; if coupled with
estimates of cluster light at larger radii, these mass
measures should suffice to determine whether mass
to light ratios vary significantly between the virial-
ized core and the surrounding infall region. For higher
redshift cluster pairs, our approach could be used to
quantify the level of environmentally induced funda-
mental plane zeropoint variations.
We describe the double cluster A2626, the FP ob-
In Section 3,
servations, and the reductions in §2.
we derive FP coefficients, discuss constraints on the
stellar populations in each subcluster, and derive con-
straints on their relative distances. Section 4 con-
tains a discussion of the radial infall model applied
to A2626. Conclusions are summarized in Section 5.
Throughout the paper we use H0 = 100h km/s/Mpc.
2. Galaxy Sample and Data
Below we describe Abell 2626 (Abell 1958), sum-
marize previous observational results detailed else-
where (MGW96), present the new observations, and
describe the reductions and analysis in detail.
2.1. Abell 2626
A redshift survey with the Decaspec (Fabricant &
Hertz 1990) mounted on the MDM Hiltner 2.4 m tele-
scope in Fall 1993 and 1994 revealed that Abell 2626
(hczi ∼ 17, 500 km/s, richness class 0) is composed of
at least two systems with mean line of sight velocities
which differ by ∼ 2, 500 km/s (MGW96). As is clear
from Fig. 1 and Table 1, there are two main compo-
nents with cluster–like dispersions: group A centered
at a velocity of hczi = 16, 533 ± 141 km/s and group
B centered at a velocity of hczi = 19, 164 ± 138 km/s
(90% statistical confidence limits). As discussed in
MGW96, the 11 galaxies centered at hczi = 21, 173 ±
119 km/s are most likely part of a background, low
density structure because (1) the dispersion is low
(σ = 200+119
−52 km/s) and (2) a large fraction of these
galaxies are gas rich- 73% have emission lines com-
pared to 39% (33%) for group A (B). If the velocity
of the low density structure corresponds to Hubble
flow then its distance from groups A and B is from
20h−1 Mpc to 50h−1Mpc.
The group A and B masses, velocity dispersions,
central densities and X–ray properties all differ. As
noted in Table 1, the virial mass of subgroup A is
roughly twice that of B, and the line of sight veloc-
ity dispersions are σA = 658+111
−81 km/s and σB =
415+117
−72 km/s (90% statistical confidence limits). The
absence of observed X–ray emission from group B can
be used to place a limit on the central gas density
in group B (MGW96).
In particular, the Einstein
imaging proportional counter (IPC) image of the re-
gion reaches roughly 50 times fainter than the peak
in the X–ray emission from group A. The group A
central gas density is ∼ 3× 10−3 cm−3, so the central
gas density in group B must be ≤ 5 × 10−4 cm−3.
Although the contrast in the projected galaxy densi-
ties of groups A and B is difficult to quantify, the
smoothed distribution (Figure 3 in MGW96) indi-
cates that the central density in group B is ∼ 7 times
less than the central density in group A. Thus, the ra-
tio of gas to galaxy density may be the same in both
groups.
As discussed below, if groups A and B are on a ra-
dial infall trajectory, then the minimum implied grav-
itational mass is more than the sum of their virial
masses. However, it is also possible that the velocity
difference is just due to Hubble flow. The magni-
tude distributions of the confirmed members of both
groups indicate that the merger hypothesis is favored
over Hubble flow at 93% confidence (MGW96). To
further investigate this issue we measure fundamen-
tal plane distances to each cluster and determine the
line of sight distance difference directly. To do so
we first use R band CCD photometry to identify a
sample of early type galaxies with redshift confirmed
membership.
2.2. Photometry
On November 21, 1995, we used a thinned, Tek-
tronix 10242 CCD mounted on the MDM Hiltner
2.4 m telescope to image galaxies in A2626 and J8
(Jackson 1982). The 24µ pixels are 0.275 ′′ on a side.
The seeing varied during the night between 0.85 ′′ and
1.05 ′′. Because the night was non–photometric, we
determine the photometric zeropoint of each R band
image externally. For the A2626 galaxies we use pho-
tometric images taken previously (MGW96) on the
MDM 1.3 m to zeropoint our images. Specifically, we
use aperture photometry of isolated stars in the 2.4 m
and 1.3 m frames to determine the 2.4 m zeropoint.
The 1.3 m images are reduced to the Kron–Cousins
system using Landolt (1992) standards. For J8 we
rely primarily on published photoelectric galaxy aper-
ture photometry (Colless et al. 1993) to determine
the R band zeropoint but also use stellar aperture
photometry from a photometric 1.3 m image for one
galaxy (Saglia et al. 1997). In addition to our primary
imaging run, we acquired additional Hiltner 2.4m im-
ages through the generosity of Paul Schechter; the
reductions for these data were similar.
The R band galactic extinction is 0.062 mag toward
A2626 (NASA/IPAC Extragalactic Database, Savage
& Mathis 1979) and 0.185 mag toward J8 (Saglia
et al. 1997). We approximate the k correction as
kR = zg where zg is the galaxy redshift (Frei & Gunn
1994). We account for the peculiar velocity compo-
nent of the line of sight velocity when applying the
cosmological dimming correction; specifically, two of
the factors of (1 + z) in the cosmological dimming
are due to relativistic effects and two are due to the
change in the geometry of the universe between emis-
sion and detection. Writing zg = zd + zp, where zd
and zp are the Hubble flow and peculiar velocity com-
ponents of the redshift, the cosmological dimming is
written C = +5 log{1 + zd} + 5 log{1 + zg}; in terms
of the observed and peculiar redshift, the correction
is C = +10log{(1 + zg)p1 − zp/(1 + zg)}. Thus, C
is model dependent. As an example, if groups A and
B are at the same distance and the peculiar velocity
correction were incorrectly taken to be zp = 0, the
fundamental plane distances would be biased so that
group B would appear closer than group A by ∼ 1%.
We bias subtract, flat field (using twilight flats),
and clean the images using standard IRAF tasks. For
each galaxy, we calculate sky subtracted radial pro-
files and uncertainties. Uncertainties have a Pois-
son component added in quadrature with an assumed
1% flat fielding uncertainty. We fit these profiles to
summed R1/4laws and exponential disks,
Ib(R) = Ib0e−7.67[(R/Rb)1/4−1]
(2-1)
3
Id(R) = Id0e−R/Rd
where Ib0 (Rb) and Id0 (Rd) are the characteristic
surface brightnesses (scale lengths). Our fitting pro-
cedure accounts for the effects of PSF smoothing and
pixel extent as described in Saglia et al. 1993b.
More specifically, we determine the parameters of
the R1/4and disk components by minimizing the χ2
difference between the seeing convolved model and
the observed profile. The expected surface brightness
I(R, dR) in the bin of width 2dR and radius R is
I(R, dR) =
[F (R + dR) − F (R − dR)]
4πRdR
(2-2)
where F (R) is the seeing convolved integral flux
within radius R. The integral flux follows from the
bulge, disk and PSF parameters. Specifically,
0
F (R) = RZ ∞
dkJ1(kR)p(k)(cid:16) Ib(k) + Id(k)(cid:17) (2-3)
where J1 is a Bessel function, and p(k), Ib(k), and
Id(k) are the Fourier transforms of the PSF, the bulge,
and the exponential disk. Following Saglia et al.
1993b, we use a PSF of the form
p(k) = e−(kb)γ
(2-4)
and find it to be a good fit to stars on the 2.4 m images
with 1.4 ≤ γ ≤ 1.6. We use the analytic approxima-
tion to the Fourier transform of the R1/4law (Saglia
et al. 1993b), and the exact form of the transform of
the disk component (e.g. Bracewell 1986).
The fitting procedure has three steps:
• Fit the PSF parameters b and γ using radial
profiles of ∼ 4 stars within each 2.4 m image;
best fit parameters minimize the χ2 difference
between the observed and model profile.
• Feed initial guesses for the bulge and disk pa-
rameters into a simplex minimization routine
(Press et al. 1988) which minimizes the χ2 dif-
ference between the seeing convolved theoretical
and observed profiles. At each iteration the full
Fourier integral is solved for each point in the
radial profile; with an HP–735 the minimization
takes several minutes.
In addition to the un-
certainties in the observed profile, a theoretical
uncertainty is introduced to account for imper-
fections in the smoothing corrections (e.g. the
exact form of the PSF). Specifically, the uncer-
tainty for each profile point has a component
which is proportional to the fractional effect of
the smoothing on the model profile. This has
the positive effect of making the fit results less
sensitive to the inclusion/exclusion of the heav-
ily smoothed inner points in the profile and to
the details of the smoothing operation. We in-
clude a theoretical uncertainty which is 10% of
the fractional effect of the smoothing on the
model profile. This is consistent with a phi-
losophy that we can correct for seeing effects at
the 90% level.
• We fit a bulge only model to the profile and
compare the minimum χ2 for this model to the
bulge plus disk χ2. For galaxies where χ2 is only
minimally improved by the inclusion of the disk
component we use the bulge only fit. Once the
fitting is complete, we use the best fit parame-
ters to calculate Re– the half luminosity radius,
µe– the surface brightness at Re, and hµei– the
mean surface brightness within Re.
We test our fitting routines by (1) fitting artificial
galaxy images, (2) comparing results from multiple
images of the same galaxy, and (3) comparing our re-
sults to those of Saglia et al. (1997) in the cluster
J8. We create artificial images by adding a galaxy
described by an R1/4 profile to a flat background of
103 cts/pixel. We then smooth the artificial image
and introduce Poisson noise. By using delta func-
tions instead of R1/4 profiles we produce stars which
are then used to measure the PSF parameters b and
γ. The process of fitting artificial galaxy images tests
the accuracy of our fitting in that case where the
true galaxy profile is well described by an R1/4 pro-
file. We find that the parameters Re and hµei are
constrained to an accuracy of ∆ log(Re) ∼ 0.02 and
∆hµei ∼ 0.07 for Re as small as 50% the FWHM
of the PSF. As expected, the accuracy of the com-
bination log(Re) − 0.33 ∗ hµei (which appears in the
equation describing the fundamental plane) is con-
strained to ∼ 0.01 because the errors in Re and hµei
are correlated.
We also compare results from subarcsecond seeing
images of seven A2626 galaxies to results from poorer
quality images. The mean offset between the param-
eters in the good and poor seeing images of the same
galaxies are consistent with zero, and the scatter is
small. Specifically, ∆hµei = −0.0262 ± 0.0311 with
4
an RMS of 0.0823, ∆ log(Re) = −0.0073±0.0074 with
an RMS of 0.0196, and ∆ (log(Re) − 0.33 ∗ hµei) =
0.0013 ± 0.0037 with an RMS of 0.0097.
We observed five galaxies in the EFAR cluster J8
(Wegner et al. 1996) for which we have photometric
zeropoints. These observations provide the strongest
test of our methods because the data, reductions and
analyses differ significantly. There are no significant
offsets between our values and EFAR values (Saglia et
al. 1997); specifically, ∆hµei = −0.0766±0.1223 with
an RMS of 0.2735, ∆ log(Re) = −0.0257±0.0326 with
an RMS of 0.0729, and ∆ (log(Re) − 0.33 ∗ hµei) =
−0.0006±0.0112 with an RMS of 0.0249. These three
tests of our photometric methods indicate that our
procedure is accurate and makes only a small con-
tribution to the observed scatter in the fundamental
plane (see below).
2.3. Spectroscopy
On November 22 and 23, 1995, we used the De-
caspec (Fabricant & Hertz 1990) plus MkIII spectro-
graph mounted on the MDM Hiltner 2.4 m telescope
to obtain spectra of four galaxy fields in Abell 2626,
one field in J8 (Wegner et al. 1996), and multiple
spectra of five HD stellar velocity standards. The De-
caspec fibers subtend 2.3′′ in the focal plane. We used
a grism blazed at 5,700 A with 600 l/mm, yielding a
typical PSF of 5.3 A at 2.18 A per pixel and coverage
from 4,300 A to 6,500A. Four or five thirty minute
exposures were obtained for each of the A2626 fields;
shorter exposures were obtained for the brighter J8
field. We extracted the spectra, using arc lamp ex-
posures taken on either side of each object exposure
to dispersion correct. We then combined the spec-
tra from the four to five exposures, excluding cosmic
rays through sigma clipping of large, positive devia-
tions. With the Decaspec there are four sky spectra
for each object spectrum. We combined the sky spec-
tra for each object, scaled the sky spectrum using the
flux in the 5577 A sky line, and then removed the
sky contribution from the object spectrum. The final
spectra range in signal to noise (S/N) from S/N=20
to 75 per pixel at 5,300 A.
We obtained spectra of more galaxies, stellar ve-
locity and Lick linestrength standards (Worthey et
al. 1994) using the Modspec mounted on the MDM
Hiltner 2.4m telescope in Nov and Dec '96. We used
a 1,200 l/mm grating, the 200 mm camera, a thinned
CCD with 24 µm pixels, and a 1.7′′ wide slit; stan-
dard longslit reductions led to spectra with 2 A reso-
lution, 1 A/pixel sampling, and coverage from 4,800A
to 5,800A. Galaxy exposure times ranged from 20 min
to 2.5 hrs, and the spectra have S/N≥ 30 per pixel.
The higher resolution of this setup allowed us to accu-
rately measure galaxy dispersions as low as 100 km/s.
For each run, galaxy dispersions and velocities are
extracted using the stellar templates from that run
and the cross correlation peak fitting available in fx-
cor (IRAF.RV); we use the approximate wavelength
region 4,800–5,800A which contains the strong Mg
features. The correlation peak width is transformed
to an intrinsic galaxy dispersion using translation ta-
bles created for each stellar template; the transla-
tion tables are produced by cross correlating the stel-
lar templates against Gaussian broadened versions of
themselves. The final galaxy dispersion is the mean
of the dispersions from each stellar template in the
run. Finally, an aperture correction
log(cid:18) σcor
σobs(cid:19) = 0.038 log(cid:18) rap
rn
ap
θn
e
θe(cid:19)
(2-5)
is applied. This aperture correction accounts for the
falling dispersion profile in early type galaxies; we
normalize to an aperture of radius rn
ap = 1.7′′ and a
galaxy with θn
e = 20′′. For the longslit spectra we use
rap = 1.025pxy/π, where x and y are the width and
length of the rectangular aperture (Jørgensen, Franx
& Kjaergaard 1995, Baggley 1996).
We estimate the uncertainties in our dispersions
using Monte Carlo techniques. For each galaxy spec-
trum we broaden a stellar template to mimic the mea-
sured dispersion, scale the template to the observed
cts/pix in the region around 5,200 A, and add the sky
at the observed level. We sample this artificial galaxy
spectrum 100 times,
introducing the Poisson noise
and sky subtracting, and then measure the dispersion.
We use the RMS of the measured dispersions around
the input value as a measure of the velocity dispersion
uncertainty (see Table 2). Not surprisingly, the sim-
ulations indicate that averaging the dispersions mea-
sured by cross correlating against multiple templates
does not improve the accuracy. However, these sim-
ulations consider only Poisson noise effects; we aver-
age the multiple dispersion measurements from each
galaxy spectrum to reduce template mismatch sys-
tematics.
We have multiple spectra for 21 galaxies. The dis-
tribution of dispersion differences scaled by the un-
certainties for the 28 comparison pairs has an RMS
of 0.97, indicating that the Monte Carlo uncertain-
5
ties are a reasonably good estimate of our true dis-
persion uncertainties. The mean scaled difference
(h(σMod − σDeca)/ǫσi where ǫσ is the uncertainty in
the difference) between Modspec and Decaspec mea-
surements for 23 comparison pairs has a mean of
−0.11 ± 0.19 with an RMS scatter about this mean
of 0.89; the variance weighted, mean logarithmic dif-
ference between Modspec and Decaspec dispersions
is −0.001 ± 0.013. Because there is no evidence for
a systematic difference between dispersions measured
with the Decaspec and the Modspec, we apply no cor-
rection. We note that the Modspec was used in mea-
suring dispersions for 10 of the 16 galaxies in group A
and 5 of the 8 galaxies in group B; so any systematic
offset between Decaspec and Modspec measurements
would add to the FP scatter within each group, but
would not bias the estimate of the distance difference.
Table 2 contains a list of the galaxies and their
properties; for the galaxies with multiple measure-
ments, we use the variance weighted average disper-
sion. The columns of Table 2 contain the galaxy tag
(first letter corresponds to group membership), the
coordinates, the redshift cz, the velocity dispersion σ
and uncertainty ǫσ, the R band apparent magnitude
MR, the mean surface brightness hµei within the half
light radius, the half light radius θe, the M gb equiva-
lent width, and the bulge luminosity fraction FB.
For 6 galaxies in J8 we compare our measured dis-
persions with those of Saglia et al. (1997). There is
evidence of an offset; the distribution of (σEF AR −
σobs)/ǫσ has a mean of −0.80 ± 0.31 (where ǫσ is the
uncertainty in the dispersion difference). The aver-
age EFAR velocity dispersion is 15 ± 7 km/s lower
than ours. This offset does not affect our estimates
of the relative distances to the two subclusters, but
would complicate efforts to bring our distances onto
the EFAR system.
We place constraints on possible differences in the
stellar populations of the two subclusters by using
IRAF scripts to measure M gb linestrengths in each
galaxy spectrum. The rest wavelengths given in
Burstein et al. (1984) for the feature and continuum
bands are corrected using the redshifts in Table 2. We
broaden our spectra to the nominal 8.6 A resolution
of the Lick system (Worthey & Ottaviani 1997) be-
fore making measurements. We tranform to the Lick
system using an expression of the form
M gLICK
b
= af (σ)M gobs
b + cap,
(2-6)
where a is a scale factor, f (σ) is a velocity disper-
6
sion correction we determine by broadening our Lick
standards, and cap is an aperture correction of the
same form as in Eq. 2-5 but with a coefficient of
0.050 instead of 0.038. We determine a separately
for each run; observations of Lick standards indicate
a = 1.060±0.016 for the Nov '96 and a = 1.086±0.028
for the Dec '96 data. We use overlapping galaxy
observations to measure a = 1.066 ± 0.034 for the
Nov '95 data. The aperture correction is similar to
that used by Jørgensen (1997), but generalized to the
form used for the velocity dispersions above (Baggley
1996).
We test the accuracy of our M gb measurements us-
ing multiple galaxy observations. Twenty three com-
parison sets from 18 different galaxies yield an RMS
of 0.051 in ∆ log M gb. Thus, we estimate the uncer-
tainty of a single observation is approximately 9%.
The linestrengths listed in Table 2 are averages of
multiple observations where appropriate. Finally, we
compare our linestrengths to preliminary EFAR M gb
linestrengths in 6 J8 cluster galaxies; the EFAR values
are lower: ∆ log M gb = −0.0733± 0.014 with an RMS
of 0.039. The linestrengths in Table 2 are corrected
to the EFAR system.
3. Fundamental Plane Analysis
In this section we, determine the coefficients and
zeropoint of the FP within the two clusters (§3.1),
discuss the evidence for stellar population variations
(§3.2), and then interpret the observed FP zeropoint
offset (§3.3).
3.1. Determining FP Coefficients
Even if the assumption that elliptical and S0 galax-
ies within both clusters are similarly distributed within
the FP is valid, the FP coefficients and cluster dis-
tances can be biased by differences in selection (e.g.
Lynden–Bell et al. 1988, Baggley 1996). Our galaxy
sample is drawn from a list of known elliptical and
S0 cluster members sorted by central aperture mag-
nitude (aperture is 2′′× 2′′ square); the parameters of
our final sample indicate that central aperture mag-
nitude is correlated with log θe − 0.3 hµei, a quantity
similar to the combination which appears in the FP.
Thus, our selection is similar to selection by isopho-
tal diameter Dn (Dressler et al. 1987). We seek to
minimally bias the coefficients by (1) combining the
samples from both clusters into a single FP fit and
(2) weighting galaxies within each cluster by the sam-
ple completeness. Finally, we examine the variation
of the zeropoint offset between the two clusters as a
function of variations in the FP coefficients (§3.3).
Our galaxy sample extends over 1.5 mag in cen-
tral aperture magnitude. We divide this range into
four bins and calculate the completeness within each
separately for the two clusters. The completeness in
each bin is simply the number of galaxies within our
final FP sample divided by the estimated total popu-
lation of elliptical and S0 cluster members. We esti-
mate the total cluster population within each bin to
be Ncl = Ntot ∗ fcl, where Ntot is the total number
of elliptical and S0 galaxies identified in the R band
survey of the cluster, and fcl is the fraction of all el-
liptical and S0 galaxies with redshifts that lie within
the velocity range of cluster cl.
Weighting by the inverse of the completeness ap-
pears to be ill advised for the faintest galaxies within
each cluster. For these two galaxies the completeness
is small enough (weight is 13.2 (17) for the faintest in
group A (B)), that weighting by the incompleteness
would lead to them dominating the fit. We exclude
the faintest galaxy from each sample when determin-
ing the FP coefficients, but include them when cal-
culating the best estimate of the cluster zeropoint.
Finally, we normalize the weights within each cluster
so that the ratio of weighting for cluster A:B is 15:7,
the ratio of the number of galaxies in each.
We fit the galactic properties to a FP of the form
log θe = α log σ + β hµei + γ, where α and β are
the (cluster invariant) coefficients of the plane and
γ is a distance dependent zeropoint (Faber et al.
1987, Djorgovski & Davis 1987, Jørgensen, Franx
& Kjaergaard 1993, JFK96). We combine galaxies
from both groups in a single fit by using deviations
around median values of distance dependent quanti-
ties (Baggley 1996). The corrected quantities hµei
and log σ are distance independent, so we transform
log θe alone; specifically, we remove an estimate of the
zeropoint within each subcluster by expressing log θe
as the variation around the γf it which minimizes the
cluster scatter in γ
log θe → log θe − γf it
(3-1)
Note that the transformation depends on the FP co-
efficients, so it must be reapplied during each fit iter-
ation.
We determine the best fit FP coefficients by mini-
mizing the sum of the absolute value of the weighted,
orthogonal deviations from the plane (JFK96, Bag-
gley 1996). The best fit values are α = 1.30 and
β = 0.31; the scatter in log Re is 0.090 (23% distance
uncertainty). Our FP has the form
log θe = (1.30 ± 0.36) log σ + (0.31 ± 0.06)hµei + γ.
(3-2)
The uncertainties are determined using a bootstrap
resampling procedure (JFK96, Baggley 1996) to re-
sample the galaxy list and fit the FP 104 times; the
uncertainty intervals are half the width of the 68%
confidence region. For comparison, the FP coeffi-
cients in the case where all galaxies are given equal
weight are α = 1.27 and β = 0.27. Our best fit FP
coefficients are statistically consistent with the val-
ues α = 1.24 ± 0.07 and β = 0.33 ± 0.01 determined
in a sample of 207 galaxies with Gunn r photome-
try (JFK96); as discussed by JFK96, their coefficients
are consistent with previous studies (e.g. Faber et al.
1987, Djorgovski & Davis 1987, Bender, Burstein &
Faber 1992, Guzman, Lucey & Bower 1993, Saglia,
Bender & Dressler 1993, Jørgensen, Franx & Kjaer-
gaard 1993). Fig. 2 contains plots of the group A
(filled) and B (hollow) galaxies along two projections
of the FP; the lines in each figure represent the FP
for groups A (solid) and B (dashed).
3.2. Constraints on the Stellar Populations
We compare the distribution of M gb equivalent
widths within the two subclusters to place constraints
on stellar population variations which could introduce
offsets in the cluster FP zeropoints, biasing the rela-
tive distances to the two clusters (e.g. JFK96). Fig. 3
contains a plot of log (M gb) versus log σ for the galax-
ies in groups A (solid points) and B (hollow points).
The lines represent best fit models for groups A (solid)
and B (dashed).
Rather than allowing both the slope and zeropoint
to vary, we constrain the slope to be that found by
Jørgensen (1997) for a sample of ∼ 300 galaxies.
Specifically, the combined correlations of M g2 with
M gb and M g2 with σ imply
log M gb = 0.307 log σ − b,
(3-3)
where the zeropoint for the large galaxy sample is
b = 0.034 (Jørgensen 1997). By minimizing orthogo-
nal deviations from the line, we find b = 0.054± 0.016
for sample A and b = 0.042 ± 0.022 for sample B.
The zeropoint of each cluster is statistically consistent
with the zeropoint from the Jørgensen sample. We
7
can measure the offset between the two clusters some-
what more accurately; the zeropoint difference be-
tween the two clusters is ∆b = bA−bB = 0.012±0.019,
statistically consistent with no offset.
Trends in M g2 have been noted as a function of
radius from the center of the Coma cluster (Guzman
et al. 1992) and as a function of cluster velocity dis-
persion for a sample of 11 clusters (JFK96; Jørgensen
1997). Another study indicates that elliptical galax-
ies in low velocity dispersion environments have both
higher metallicities and younger stellar populations
than galaxies in higher velocity dispersion environ-
ments (Rose et al.
1994). The velocity disper-
sion difference between clusters A (σ = 658 km/s)
and B (σ = 415 km/s) should introduce an offset
of ∆b ∼ −0.016 into the M gb–σ correlation. This
expectation is inconsistent with our observations at
the 1.5σ level; we would require additional galaxy
linestrengths to resolve an offset of that magnitude.
3.3. FP Zeropoint Differences
We determine the zeropoint of each group by using
the median γ; the 16 galaxies in group A yield γA =
−8.4083 with an RMS around this value of 0.0879,
and the 8 galaxies in group B yield γB = −8.3632
with an RMS of 0.1149. The homogeneous Malmquist
bias has a marginal effect on the relative distances to
these two clusters (Lynden–Bell et al. 1988); we cal-
culate δγA = −0.0033 and δγB = −0.0114, resulting
in an increase in ∆γ of ∼ 0.0081. Thus, the zeropoint
difference between the two clusters is
∆γ = γA − γB = −0.037 ± 0.046,
(3-4)
where we have assumed the uncertainty in the zero-
point is the scatter about that value divided by √N ,
where N is the number of galaxies (note that this is
a conservative estimate of the true uncertainties be-
cause the scatter is sensitive to outliers). As discussed
in §2.2, if we assume groups A and B are at the same
distance rather than at distances proportional to their
mean redshifts, the cosmological dimming corrections
change so that ∆γ increases by ∼ 0.004.
In general, zeropoint offsets can be caused by stel-
lar population or metallicity induced M/LR differ-
ences and/or distance differences. Because there is no
compelling evidence for stellar population differences,
we use the observed zeropoint offset and uncertainty
as relative distance constraints. The quantity γ (in
Eqn. 3-2) can be written γx = log (Γ/Dx) where Dx
is the distance of galaxy x and Γ follows from the
standard form implied by the FP
Re = Γσα hIei−2.5β .
(3-5)
Assuming the intrinsic FP zeropoint log Γ is the same
for all galaxies, the relative distances to two galaxies
x and y is
log(cid:18) Dx
Dy(cid:19) = γy − γx.
(3-6)
Fig. 4 contains a plot of the distances to the galaxies
in groups A (solid) and B (hollow) relative to the me-
dian distance to group A versus the galaxy redshift.
Also marked with vertical lines are the (homogeneous
Malmquist bias corrected) median distances to groups
A (DA; solid line) and B (log (DB/DA) = −0.037;
dashed line); the horizontal lines mark the mean ve-
locities for groups A (solid) and B (dashed). The error
bars on the dashed, vertical line represents the rela-
tive distance uncertainty. The large star marks the
position of cluster B if its distance relative to cluster A
were reflective of pure Hubble flow (∆D ∝ ∆hvi); the
Hubble flow hypothesis implies log (DB/DA) = 0.065.
The zeropoint difference is consistent with zero,
and it is sufficient to rule out Hubble flow at 2.2σ;
therefore, the FP measurements suggest that cluster
B is closer to us than cluster A (in §5.1 we return to
this issue in more detail). Naturally, the zeropoint
offset depends on the coefficients of the FP, which are
only constrained to within ∼20% of the best fit values.
Using the distribution of γA − γB from the 104 boot-
strap resampling fits, we find that in 99.05% of the
simulations the zeropoint offset between the two clus-
ters is less than the Hubble flow offset, rejecting the
Hubble flow hypothesis. This probability corresponds
to a +2.35σ deviation in a Gaussian, similar to the
2.2σ estimate above which we derived from the scat-
ter about the median distance in each cluster. The
bootstrap probability is superior because it includes
variations in the FP coefficients (and is less sensitive
to outliers than the scatter estimator).
4. Radial Infall Model as a Mass Constraint
The radial infall model has been applied to many
close galaxy cluster pairs to determine whether or not
the pairs will merge (e.g. Beers, Geller & Huchra
1982, Beers et al. 1991, Beers et al. 1992, Colless &
Dunn 1995, Scodeggio et al. 1995). In the following
we use the measured line of sight separation and this
model to estimate the total binding mass required to
8
explain the observed infall velocity. We then compare
this mass to virial estimates of the cluster binding
mass.
4.1. Radial Infall Model
For two isolated, bound objects within an expand-
ing universe, the dynamics can be parametrized in the
standard way
(4-1)
l = (cid:0)GMGB2(cid:1)1/3
B (cid:19)1/3
v = (cid:18) GMG
t = B(η − sin η)
(1 − cos η)
sin η
(1 − cos η)
where η is a development angle, l is the separation,
v is the relative velocity, t is the age of the universe,
B is an undetermined coefficient, and MG is the sum
of the individual masses (e.g. Peebles 1993). When
we observe two clusters on a merger trajectory at the
present epoch, we measure the projected separation
of the clusters ∆l⊥ and a line of sight velocity dif-
ference ∆vLOS. Modelling the collision as a radial
infall, we parametrize the merger in terms of the an-
gle φ between the merger trajectory and the line of
sight (∆vLOS = ∆v cos φ and ∆l⊥ = ∆l sin φ).
With four free parameters in this model (η, B, φ
and MG) and three observables (the age of the uni-
verse tH , the projected separation ∆l⊥ and the line
of sight velocity difference ∆vLOS), the total cluster
mass and age of the universe are a function of the an-
gle φ and the present epoch value of the development
angle ηH .
MG = 2.3 × 1014M⊙(cid:0)(1 + cos ηH ) cos2 φ sin φ(cid:1)−1
1000 km/s(cid:19)2(cid:18) ∆l⊥
1 Mpc(cid:19)
(cid:18) ∆vLOS
(4-2)
tH = 9.8 × 108 yrs (cid:18) (ηH − sin ηH ) sin ηH
cot φ(cid:19)
(1 − cos ηH )2
1 Mpc(cid:19)(cid:18) 1000 km/s
∆vLOS (cid:19)
(cid:18) ∆l⊥
Using an approximate age of the universe to deter-
mine ηH , we arrive at the total cluster mass function
MG(φ). Fortunately the dependence of MG on tH is
rather weak (see Fig. 5).
If the line of sight sepa-
ration between the subclusters ∆lLOS is measured, φ
is determined, yielding the total system mass (which
9
we term the cluster merger mass MG). MG can then
be compared to other mass estimates, which typically
probe the cores of clusters.
Departures from radial infall and the physical ex-
tent of the clusters can bias the merger mass. Gen-
erally speaking, the large mass and rarity of galaxy
clusters make them good candidates for this analysis.
We plan to study the accuracy of merger masses us-
ing numerical simulations of cluster evolution within
"realistic" environments.
Note that there are scenarios in which the merger
mass would be substantially biased; it is possible to
recognize these scenarios observationally. In the first
scenario, the two merging clusters are at small sep-
interactions among the cluster
aration (≤ 1 Mpc);
components, significant overlap of their mass profiles,
and small departures from pure radial infall combine
to bias the inferred merger velocity and the implied
merger mass. Fortunately, the signatures of cluster
mergers have been extensively studied (e.g. Geller &
Beers 1982, Dressler & Shectman 1988, Jones & For-
man 1992, Mohr, Fabricant & Geller 1993, Pearce,
Thomas & Couchman 1994, Mohr et al. 1995, Buote
& Tsai 1995), and typically there are merger clues in
the galaxy and gas distributions.
In the second scenario, a third object of compara-
ble mass to the two merging clusters is close enough
to invalidate the two body analysis. In this case, the
observed distribution of nearby galaxy clusters should
be enough to determine whether or not the radial in-
fall model is appropriate.
4.2. Merger Mass in A2626
To calculate MG we require (see Eqn. 4-2) the
observed rest frame velocity difference and the pro-
jected separation. The rest frame velocity difference
is ∆v = 2, 486 ± 112 km/s. The projected separation
between the centers of mass is assumed to be the dis-
tance between the bright, central elliptical in group A
(its position is consistent with the peak in the X–ray
emission) and the centroid of the 30 group B members
with measured redshifts (using the centroid of group
A members slightly increases the projected separa-
tion and the merger mass). At a cluster distance of
l = 175h−1 Mpc, the projected separation between
the two groups is ∆l⊥ = (0.707 ± 0.052)h−1 Mpc.
The uncertainty reflects the statistical uncertainties
associated with centroiding group B. The fractional
uncertainty in the merger mass MG due to ∆vLOS
and ∆l⊥ is 12%.
Fig. 5 displays the ratio of the cluster merger mass
MG to the sum of the virial masses of groups A and
B (MA + MB = 9.1 × 1014h−1M⊙; see Table 1 and
MGW96) versus the line of sight separation between
the subclusters, ∆lLOS = ∆l⊥/ tan φ. The dotted
line corresponds to tH = 18 Gyr and the solid line
corresponds to tH = 13 Gyr. The FP distances to
galaxies within groups A and B yield a ratio of the
group B distance to the group A distance (lB/lA)
which can be used to calculate ∆lLOS. Specifically,
we estimate the line of sight separation between the
subclusters as ∆lLOS ∼ 175h−1(1− lB/lA) Mpc. The
sample of 24 FP distances to groups A and B yields
∆lLOS = (14 ± 18)h−1 Mpc.
Because distance uncertainties in this case are
large, our method does not yield tight mass con-
straints in Abell 2626; however, if groups A and B are
merging (99% confidence from FP distances and 93%
confidence from galaxy magnitude distributions), it is
clear that the virial masses of groups A and B under-
estimate the total system binding mass. Specifically,
the sum of the masses of groups A and B must be
at least 1.5 × 1015M⊙, compared to the virial sum of
9.1+3.1
−2.7 × 1014M⊙ (MGW96), a factor of 1.65 higher.
The virial mass range corresponds to 90% statistical
confidence limits, where it is assumed the mass uncer-
tainties are dominated by the velocity dispersion un-
certainties (Heisler, Tremaine & Bahcall 1985). The
virial sum is inconsistent with the minimum merger
mass at the ∼ 3σ level.
5. Discussion
The study of double clusters with the fundamental
plane holds promise for significant progress in two ar-
eas. First, the study of appropriately isolated nearby
double clusters should lead to new constraints on the
total cluster binding mass. These constraints will not
only provide an independent test of the virial, hy-
drostatic, and weak lensing mass estimators, but will
also provide information about the mass (and typical
M/L) outside the virialized cluster region. Second,
the study of more distant double clusters (at z ≥ 0.1)
provides a means of directly probing for environmen-
tally driven biases in FP distance estimates.
5.1. Fundamental Plane Analysis
Here we apply this approach to the double cluster
A2626 at cz ∼ 17, 500 km/s. Because of the cluster
10
distance, it is critical to make seeing corrections while
extracting θe and hµei. Our method is similar to the
one described by Saglia et al. (1993b), but includes
a seeing–correction uncertainty which makes the pho-
tometric parameters less sensitive to the in/exclusion
of central points in the profile. We also note that
the cosmological dimming correction to the surface
brightness has a peculiar velocity dependence which
may be important in cases where many FP distance
estimates are combined to produce a single, more ac-
curate cluster distance (§2.2).
Through numerous cross checks we demonstrate
that our estimates of the combination log(θe)− β hµei
which appears in the FP are sufficiently accurate that
their errors make no significant contribution to the FP
scatter; using a Monte Carlo approach we estimate
the velocity dispersion uncertainties (§2.3). These pa-
rameters for the 8 galaxies in group B and the 16 in
group A are listed in Table 2.
We combine galaxies from both clusters in a single
fit, weighting to account for incompleteness (§3.1).
Our best fit FP coefficients (α = 1.30 ± 0.36 and
β = 0.31± 0.06) are statistically consistent with mea-
surements in other clusters (see Fig. 2). The RMS
scatter about this plane in ∆ log Re is 0.09, corre-
sponding to 23% distance uncertainties per galaxy.
We examine the distribution of M gb equivalent
widths with the galaxy spectra from both subclus-
ters (§3.2). We find that there is no evidence for sig-
nificant differences in the M gb–σ relation (see Fig.
3). Our result does not contradict the recently
noted correlation between cluster environment and
linestrength (JFK96, Jørgensen 1997); the expected
offset in M gb (given the velocity dispersion differences
between groups A and B) is too small to detect with
our data.
The difference in the fundamental plane zeropoints
for the two clusters is γA − γB = −0.037± 0.046, con-
sistent with no offset (§3.3). Under the assumption
that this zeropoint difference is indicative of distance
differences, log (DB/DA) = −0.037±0.046, where the
uncertainty follows from scatter around the median
distance in each cluster with the best fit FP coeffi-
cients. We use the 104 bootstrap resampling simula-
tions to measure the variation of the distance offset
as FP coefficients vary. The relative distance con-
straint is robust enough to reject the Hubble flow hy-
pothesis (log (DB/DA) = 0.065) in 99% of the simu-
lations. Under the Hubble flow hypothesis, the two
clusters are not interacting gravitationally, and the
velocity difference reflects pure Hubble flow. An anal-
ysis of the R band magnitude distributions of con-
firmed members of both clusters rules out Hubble flow
with 93% confidence (MGW96).
If the subclusters were on an outgoing trajectory
(bound or unbound), then the line of sight distance
difference would have to be greater than their cur-
rent recession velocity times the age of the universe
(∆lLOS > vLOStH ∼ 34 Mpc) which is greater than
the Hubble flow value: log (DB/DA) ≥ 0.0645; thus,
models where groups A and B are on outgoing tra-
jectories are ruled out with higher confidence than
pure Hubble flow. Therefore, our relative distances
indicate that the subclusters are bound and infalling
with 99% confidence (see Fig. 4).
5.2. Radial Infall Model
Assuming bound and infalling subclusters, we model
these kinematics with a radial infall model (Beers,
Geller & Huchra 1982); this two body model is appro-
priate for separations large compared to the scale of
the virialized region of the cluster and in cases where
there are no other massive clusters in the neighbor-
hood. Given the projected separation, line of sight
velocity difference, and line of sight separation, the
radial infall model provides an estimate of the total
cluster binding mass which depends weakly on the
adopted age of the universe. From Fig. 5, the total
binding mass is at least 1.65 times higher than the
virial sum, a difference significant at ∼ 3σ. The total
binding mass could be much larger than this mini-
mum.
The differences in these two mass estimates could
indicate (1) deviations from radial infall, (2) an error
in the virial estimate, or (3) significant mass beyond
the virialized region. The X–ray image of A2626 pro-
vides no indication of interactions between the two
subclusters, and there is no third subcluster, so the
radial infall model should be appropriate in A2626.
If the virial estimators are biased low, then the radial
infall model indicates that M/LR ∼ 1000h (see Ta-
ble 1), which would be a surprisingly high value (e.g.
Ramella et al. 1989). These arguments suggest that
the minimum merger mass is larger than the virial
mass because it is sensitive to mass in the cluster in-
fall region.
Taking the merger mass as a lower limit on the
mass within the cluster infall region, we can address
whether there is evidence for variation of the mass to
light ratio between the cluster core and infall region.
The galaxy light is calculated within a region which
only extends to a projected radius of 1.5h−1 Mpc;
presumably both the galaxy and mass distributions
extend beyond this region. We have no CCD pho-
tometry over the larger region to directly measure
the total cluster light. Assuming that the projected
galaxy distribution falls off as ∼ R−1.5 (e.g. Mohr
et al. 1996) with a similar luminosity function to
the galaxies in the central region, the total cluster
light grows as ∼ √R for R larger than the core ra-
dius. Under these assumptions, within a diameter of
∼ 4h−1 Mpc there is enough light to yield a mass–to–
light ratio of M/LR ∼ 600h throughout the cluster
even with the higher mass estimate from the radial
infall model. Of course, the binding mass could be
many times larger than the minimum value, so the
mass–to–light ratio could increase significantly out-
side the virialized cluster core. A study of the cluster
light extending to larger radius and a larger sample of
FP distances would yield more concrete information
regarding any variation in the cluster mass–to–light
ratio in A2626.
FP studies of other double clusters, coupled with
(1) a more detailed accounting of the radial distribu-
tion of cluster light and (2) larger redshift samples to
enable a more accurate estimate of the virial mass,
should provide interesting constraints on the distri-
bution of cluster mass and possible differences in the
efficiency of galaxy formation beyond the virialized
region.
We thank Roberto Saglia and Glenn Baggley for
their comments on an earlier version of this manuscript
and for their generosity in allowing us to compare
our measurements to their unpublished results. We
thank Guy Worthey for stimulating discussions, Paul
Schechter for obtaining improved 2.4 m images of
several galaxies in our sample, and Emilio Falco for
attempting additional imaging. This research was
funded in part by NAGW–2367 and NAG5–3401.
REFERENCES
Abell, G. 1958, ApJS, 3, 211
Baggley, G. 1996, D.Phil. Dissertation, Oxford Uni-
versity
Beers, T.C., Geller, M.J. & Huchra, J.P. 1982, ApJ,
257, 23
Beers, T.C., Forman, W., Huchra, J.P., Jones, C. &
Gebhardt, K. 1991, AJ, 102, 1581
11
Beers, T.C., Gebhardt, K., Huchra, J.P., Forman, W.,
Jones, C. & Bothun, G.D. 1992, ApJ, 400, 410
Bender, R., Burstein, D., & Faber, S.M. 1992, ApJ,
399, 462
Bracewell, R.N., 1986, The Fourier Transform and Its
Applications, (McGraw–Hill: New York)
Buote, D. & Tsai, J. 1995, ApJ, 452, 522
Burstein, D., Faber, S., Gaskell, C.M., Krumm, N.
1984, ApJ, 287, 586
Colless, M., Burstein, D., Wegner, G., Saglia, R.P.,
McMahan, R., Davies, R.L., Bertschinger, E. &
Baggley, G. 1993, MNRAS, 262, 475
Colless, M. & Dunn, A. 1995, ApJ, 458, 435
David, L.P., Jones, C. & Forman, W. 1995, ApJ, 445,
578
Djorgovski, S. & Davis, M. 1987, ApJ, 313, 59
Dressler, A., Lynden–Bell, D., Burstein, D., Davies,
R.L., Faber, S.M, Terlevich, R.J. & Wegner, G.
1987, ApJ, 313, 42
Dressler, A. & Shectman, S. 1988, AJ, 95, 985
Faber, S.M., Dressler, A., Davies, R.L., Burstein, D.,
Lynden–Bell, D., Terlevich, R.J. & Wegner, G.
1987, in Nearly Normal Galaxies, ed. S.M. Faber,
(New York: Springer–Verlag), 175
Fabricant, D.G. & Hertz, E. 1990, SPIE Proc. 1235,
747
Frei, Z. & Gunn, J.E. 1994, AJ, 108, 1476
Geller, M.J. & Beers, T.C. 1982, PASP, 94, 421
Guzman, R., Lucey, J.R., Carter, D. & Terlevich, R.J.
1992, MNRAS, 257, 187
Guzman, R., Lucey, J.R., & Bower, R.G. 1993, MN-
RAS, 265, 731
Heisler, J., Tremaine, S. & Bahcall, J.N. 1985, ApJ,
298, 8
Jackson, R.E. 1982, PhD Dissertation, University of
California, Santa Cruz
Jones, C. & Forman, W. 1992, in Clusters and Su-
perclusters of Galaxies (NATO ASI Vol. 366), ed.
A.C. Fabian, London: Kluwer, p. 49
Jørgensen, I. Franx, M. & Kjaergaard, P. 1993, ApJ,
411, 34 (JFK93)
Jørgensen, I. Franx, M. & Kjaergaard, P. 1995, MN-
RAS, 276, 1341
Jørgensen, I. Franx, M. & Kjaergaard, P. 1996, MN-
RAS, 280, 167 (JFK96)
Jørgensen, I. 1997, astro–ph 9702076
Lacey, C. & Cole, S. 1993, MNRAS, 262, 627
Lacey, C. & Cole, S. 1994, MNRAS, 271, 676
Landolt, A. 1992, AJ, 104, 340
Lucey, J.R. & Carter, D. 1988, MNRAS, 235, 1177
12
Lynden–Bell, D., Faber, S.M., Burstein, D., Davies,
R.L., Dressler, A., Terlevich, R.J. & Wegner, G.
1988, ApJ, 326, 19
Mohr, J.J., Fabricant, D.G. & Geller, M.J. 1993, ApJ,
413, 492
Mohr, J.J., Evrard, A.E., Fabricant, D.G. & Geller,
M.J. 1995, ApJ, 447, 8
Mohr, J.J., Geller, M.J., Fabricant, D.G., Wegner,
G., Thorstensen, J. & Richstone, D.O. 1996, ApJ,
470, 724
Mohr, J.J., Geller, M.J. & Wegner, G. 1996, AJ, 112,
1816; (MGW96)
Pearce, F.R., Thomas, P.A. & Couchman, H.M.P
1994, MNRAS, 268, 953
Peebles, P.J.E. 1993, Principles of Physical Cosmol-
ogy, (Pinceton University Press: Princeton), p 535
Press, W.H., Flannery, B.P., Teukolsky, S.A. & Vet-
terling, W.T. 1988, Numerical Recipes in C, (Cam-
bridge University Press: Cambridge)
Ramella, M., Geller, M.J., & Huchra, J.P. 1989, ApJ,
344,57
Rose, J.A., Bower, R.G., Caldwell, N., Ellis, R.S.,
Sharples, R.M. & Teague, P. 1994, AJ, 108, 2054
Saglia, R.P., Bender, R. & Dressler, A. 1993a, A&A,
279, 75
Saglia, R.P., Bertschinger, E., Baggley, G., Burstein,
D., Colless, M., Davies, R.L., McMahan, R.K. &
Wegner, G. 1993b, MNRAS, 264, 961
Saglia, R.P., Bertschinger, E., Baggley, G., Burstein,
D., Colless, M., Davies, R.L., McMahan, R.K. &
Wegner, G. 1997, MNRAS, submitted
Sandage, A. & Bedke, J. 1994, The Carnegie Atlas of
Galaxies: Volumes I & II, (Carnegie Institution of
Washington: Washington, D.C.)
Savage, B.D. & Mathis, J.S. 1979, Ann Rev Astron
Astro, 17, 73
Scodeggio, M., Solanes, J.M., Giovanelli, R. &
Haynes, M.P. 1995, ApJ, 444, 41
Wegner, G., Colless, M., Baggley, G., Davies, R.L.,
Bertschinger, E., Burstein, D., McMahan, R.K. &
Saglia, R.P. 1996, ApJS, 106, 1
White, S.D.M., Navarro, J.F., Evrard, A.E. & Frenk,
C.S. 1993, Nature, 366, 429
Worthey, G., Faber, S.M., Gonzalez, J.J. & Burstein,
D. 1994, ApJS, 94, 687
Worthey, G., Trager, S.C. & Faber, S.M. 1996, in
Fresh Views of Elliptical Galaxies, eds. Buzzoni,
A., Renzini, A. & Serrano, A. ASP Conference Se-
ries, Vol 86, 203
Worthey, G., & Ottaviani, D. L. 1997, ApJS, in press
This 2-column preprint was prepared with the AAS LATEX
macros v4.0.
A. Effects of Culling the Galaxy Sample
Four of our 24 galaxies show peculiar morphologies; here we detail the peculiarities and demonstrate that
eliminating them from the analysis does not significantly alter our conclusions. The following are morphological
types which we derived by comparing deep 2.4 m images and the Carnegie Atlas of Galaxies (Sandage & Bedke
1994):
• AH: bright nucleus with possible bar; probable SB0
• AL: possible spiral structure; SB0 or Sa
• AM: bright nucleus with possible bar; probable SB0
• AP: appears to have a dust ring; probable RSB0
Without these four galaxies, Group A (B) contains 12 (8) galaxies; we now apply the analysis detailed in §3 to
this smaller sample. The best fit form of the FP is
log θe = (1.39 ± 0.65) log σ + (0.32 ± 0.06)hµei + γ,
(A1)
and the scatter around this plane in log Re is 0.080 (20% distance uncertainty). These coefficients are consistent
with the values from the full sample, but the scatter is somewhat smaller and the α uncertainty is larger. The FP
coefficient uncertainties are determined (as before) through bootstrap resampling and refitting the sample.
The RMS scatter around the median distance estimator in group A (B) is 0.066 (0.115); the zeropoint difference
corrected for homogeneous Malmquist bias is
∆γ = −0.026 ± 0.045.
(A2)
The zeropoint difference and uncertainty (from the scatter) are somewhat smaller than for the whole sample (see
Eqn. 3-4). This measurement is inconsistent with the Hubble flow hypothesis at ∼ 2.0σ or 97.7% confidence. The
bootstrap refitting produces a distribution of ∆γ which includes the variation in the FP coefficients; 5000 refits of
this smaller sample rule out the Hubble flow hypothesis with 98.2% confidence. In summary, removing 4 galaxies
which have morphological peculiarities slightly reduces the significance with which the Hubble flow model can be
rejected in Abell 2626.
Table 1
Subcluster Properties
ID N
¯v
σ
Mvir[1014M⊙] Mvir/LR
A
B
67
30
16,533±141
19,164±138
658+111
−81
415+117
−72
6.6+2.4
−1.5
2.3+1.4
−0.7
630
570
Intervals are statistical 90% confidence limits
13
Table 2
Galaxy Properties
Tag
RA(1950)
Decl
cz
σ
ǫσ MR
23 34 29.30
AA
23 35 18.21
AB
23 34 22.71
AC
23 33 40.11
AD
23 33 23.09
AE
23 35 33.91
AF
23 32 40.07
AG
23 33 40.68
AH
23 34 32.46
AI
23 34 35.93
AJ
23 33 58.71
AK
AL
23 34 08.75
AM 23 33 49.52
23 33 57.86
AO
23 33 47.81
AP
23 35 19.60
AZ
BA
23 34 07.96
23 33 19.80
BB
23 33 57.29
BC
23 33 08.31
BD
23 33 17.67
BE
BF
23 33 33.61
23 33 17.50
BG
BH
23 33 12.20
20 22 22.3
20 31 56.7
20 56 55.6
20 45 31.4
20 42 02.1
21 07 54.7
21 00 01.6
20 42 35.6
20 18 14.5
20 59 41.0
21 01 12.4
20 49 29.9
20 49 12.1
20 51 11.8
20 49 21.8
20 30 15.3
20 31 29.6
20 33 12.0
20 35 11.2
20 48 02.8
20 22 54.2
20 49 36.5
20 52 03.1
20 35 20.1
17,538
17,159
16,819
17,911
15,941
16,728
15,979
16,644
17,804
16,665
16,856
17,645
16,129
16,113
16,859
16,063
19,729
19,369
19,240
19,150
19,215
19,825
18,818
19,136
287
223
218
190
205
191
163
161
185
178
202
152
252
190
172
118
148
219
187
214
133
119
133
151
12
22
22
22
28
12
19
6
7
35
20
3
21
27
8
11
3
21
17
10
3
3
6
5
14.06
13.93
14.37
14.75
15.03
14.97
14.71
15.02
14.93
15.35
15.35
14.61
14.40
15.91
15.07
16.02
14.84
14.31
14.36
14.21
15.18
15.89
16.02
17.09
hµei
19.79
19.47
19.56
19.93
19.22
19.23
19.33
20.15
19.24
18.94
19.23
19.67
21.03
17.66
19.70
19.74
19.19
19.61
20.06
20.48
19.17
19.77
19.71
18.34
θa
e
M gb
6.232
5.477
4.846
4.817
3.063
3.161
3.727
4.732
3.231
2.319
2.564
4.554
9.403
0.997
3.743
2.136
3.351
4.900
6.243
8.148
2.836
2.705
2.431
0.761
4.83
4.50
4.41
4.64
4.42
4.17
3.83
4.54
4.14
4.67
4.86
3.67
4.87
4.68
4.86
3.89
3.80
4.27
4.42
4.45
4.27
3.79
4.61
4.98
F b
B
0.53
0.34
0.79
0.63
0.55
0.74
0.85
0.50
0.84
0.90
0.70
0.39
0.64
1.00
0.30
0.70
0.69
0.74
0.69
1.00
0.62
0.38
0.68
0.33
ain arc seconds
bbulge luminosity fraction
14
Fig. 1.- The galaxy velocity distribution around Abell 2626. The histogram contains 108 galaxy velocities from the
redshift survey of MGW96. Group A is centered at v = 16, 500 km/s, and group B is centered at v = 19, 150 km/s.
The high fraction of galaxies with emission lines in the group of 11 centered at v ∼ 21, 500 km/s (73% compared
to 39%/33% for groups A/B) indicates they are part of a low density background structure.
15
Fig. 2.- Two projections of the FP. Both plots contain group A (solid) and B (hollow) galaxies. The lines define
the best fit group A (solid) and B (dashed) FP; offsets in the fits reflect subcluster distance differences. The best fit
FP (with weighting to account for incompleteness– see text) is of the form log θe = 1.30 log σ + 0.31 hµei + γcl where
γcl is the distance dependent cluster zeropoint. The RMS scatter around the best fit in log θe is 0.09, corresponding
to a 23% distance uncertainty per galaxy.
16
Fig. 3.- Plot of log (M gb) vs log (σ) for the galaxies in groups A (solid) and B (hollow). The lines mark the best fit
relations for groups A (solid) and B (dashed) constrained to have the slope found using a much larger galaxy sample
(Jørgensen 1997). The apparent zeropoint offset between the two samples is statistically insignificant (0.012±0.019;
see text), providing no evidence for significant stellar population differences between the subclusters.
17
Fig. 4.- Plot of fundamental plane relative distances versus redshift for galaxies in groups A (solid) and B
(hollow). Galaxies in groups A and B are clearly offset in velocity (see Fig. 1), and also appear to be offset in
distance. Vertical lines mark the (homogeneous Malmquist bias corrected) median distances to groups A (solid) and
B (dashed). Horizontal lines mark the mean velocities of groups A (solid) and B (dashed). The large star marked
with DHubble lies at the expected position of group B if its trajectory relative to group A were pure Hubble flow.
The median distances to groups A and B indicate that log (DB/DA) = −0.037± 0.046; the 1σ distance uncertainty
is indicated by the error bar on the vertical, dashed line. The sample of 24 galaxies rules out the Hubble flow
hypothesis, DB = DHubble and log (DB/DA) = 0.065, with 99% confidence (see §3.3).
18
Fig. 5.- The ratio of the cluster merger mass MG to the sum of the group A and B virial masses MA + MB =
9.1×1014h−1M⊙ as a function of the line of sight separation ∆lLOS of the two groups. The two solutions correspond
to different assumptions regarding the age of the universe tH : tH = 13 Gyr (solid) and tH = 18 Gyr (dotted). The
statistical uncertainty in the merger mass MG is 12%. If groups A and B are merging, as suggested by the FP
distances and the galaxy magnitude distributions, the merger mass is at least 1.65 times greater than the sum of
the virial masses. The FP distances indicate that ∆lLOS = 14 ± 18 h−1 Mpc.
19
|
astro-ph/0210358 | 3 | 0210 | 2004-04-14T17:36:14 | Testable anthropic predictions for dark energy | [
"astro-ph",
"gr-qc",
"hep-ph"
] | In the context of models where the dark energy density $\rD$ is a random variable, anthropic selection effects may explain both the "old" cosmological constant problem and the "time coincidence". We argue that this type of solution to both cosmological constant problems entails a number of definite predictions, which can be checked against upcoming observations. In particular, in models where the dark energy density is a discrete variable, or where it is a continuous variable due to the potential energy of a single scalar field, the anthropic approach predicts that the dark energy equation of state is $p_D=-\rho_D$ with a very high accuracy. It is also predicted that the dark energy density is greater than the currently favored value $\Omega_D\approx 0.7$. Another prediction, which may be testable with an improved understanding of galactic properties, is that the conditions for civilizations to emerge arise mostly in galaxies completing their formation at low redshift, $z\approx 1$. Finally, there is a prediction which may not be easy to test observationally: our part of the universe is going to recollapse eventually. However, the simplest models predict that it will take more than a trillion years of accelerated expansion before this happens. | astro-ph | astro-ph |
Testable anthropic predictions for dark energy
J. Garriga1,2 and A. Vilenkin 2
1 Departament de F´ısica Fonamental, Universitat de Barcelona,
Mart´ı i Franqu`es 1, 08193 Barcelona, Spain and
2 Institute of Cosmology, Department of Physics and Astronomy,
Tufts University, Medford, MA 02155, USA
(Dated: September 27, 2018)
In the context of models where the dark energy density ρD is a random variable,
anthropic selection effects may explain both the "old" cosmological constant problem
and the "time coincidence". We argue that this type of solution to both cosmologi-
cal constant problems entails a number of definite predictions, which can be checked
against upcoming observations. In particular, in models where the dark energy den-
sity is a discrete variable, or where it is a continuous variable due to the potential
energy of a single scalar field, the anthropic approach predicts that the dark energy
equation of state is pD = −ρD with a very high accuracy. It is also predicted that
the dark energy density is greater than the currently favored value ΩD ≈ 0.7. An-
other prediction, which may be testable with an improved understanding of galactic
properties, is that the conditions for civilizations to emerge arise mostly in galaxies
completing their formation at low redshift, z ≈ 1. Finally, there is a prediction
which may not be easy to test observationally: our part of the universe is going to
recollapse eventually. However, the simplest models predict that it will take more
than a trillion years of accelerated expansion before this happens.
I.
INTRODUCTION
The "old" cosmological constant problem - why don't we see the large vacuum energy
density ρΛ which is expected from particle physics? - and the "time coincidence" problem
- why do we live at the epoch when the dark energy component ρD starts dominating? -
may find a natural explanation in models where ρD is a random variable. The idea is to
introduce a dynamical dark energy component X whose contribution ρX varies from place
to place, due to processes which occurred in the early universe. Then
ρD = ρΛ + ρX
will also vary from place to place, and the old cosmological constant problem takes a different
form. The question is not why ρΛ is much smaller than η4, where η is some high energy
physics mass scale, such as the supersymmetry breaking scale η ∼ T eV , but why do we
happen to live in a place where ρΛ is almost exactly cancelled by ρX. This line of enquiry is
rather quantitative, since we can ask what is the probability for us to observe certain values
of ρD ∼ 10−11(eV )4, or what is the probability for the time coincidence.
Explicit particle physics models for a variable ρX have been reviewed in [1]. Two examples
which have been thoroughly discussed in the literature are a four-form field strength, which
can vary through nucleation of membranes [2, 3], and a scalar field with a very low mass [3, 4].
Assuming one such mechanism, and using a theory of initial conditions such as inflation,
2
one can calculate the "a priori" probability distribution P∗(ρD)dρD. This is defined as the
fraction of co-moving volume which at some fiducial initial time (which we conventionally
take to be the time of recombination) had the value of the dark energy density in the interval
dρD. Inflation is also responsible for smoothing out the value of ρD over comoving distances
much larger than the size of our presently observable universe.
By itself, P∗ is not sufficient to calculate probabilities for our observations. Selection
effects which bias the measurement of ρD must be included, and the most important one
1 While ρD may be very large in most places, there is
in this case is anthropic [5 -- 9].
nobody there to observe such extreme values. If ρD > 0, galaxy formation stops once the
dark energy becomes dominant over the matter density. Some galaxies are seen at redshifts
of order z ∼ 5, but not much higher, indicating [9] that galaxies will not form in regions
where ρD & (1 + zEG)3ρ0. Here, ρ0 is the matter density at the present time t0, and zEG ≈ 5
is the redshift at the time tEG ∼ (1 + zEG)−3/2t0 when the earliest galaxies formed. Also,
for a negative ρD the universe recollapses on a time scale tD ∼ GρD−1/2, where G is
Newton's constant. This time should be larger than the earliest time tEI which is required
for intelligence to develop [7, 13]. Thus, observers will only exist within a tiny "anthropic
range"
(1)
−(Gt2
EI)−1 . ρD . (Gt2
EG)−1.
It should be noted that, aside from the above minimal requirements, anthropic selection
includes all other ways in which ρD disfavours the existence of observers. For instance, in
regions where ρD < 0, the matter density is larger than ρD throughout the cosmic evolution.
If ρD is too large, all galaxies formed in that region will be very dense, and as a result,
very inhospitable. This occurs also for a large ρD > 0, since galaxies must form before ρD
starts dominating. We shall come back to this issue in Sections III and V.
The selection effect can be implemented quantitatively by assuming the mediocrity prin-
ciple, according to which our civilization is typical in the ensemble of all civilizations in the
universe. The probability to find ourselves in a region with given values of ρD is thus given
by [14]
dP(ρD) ∝ P∗(ρD)nciv(ρD)dρD.
(2)
Here, nciv(ρD) refers to the number of civilizations which will ever form per unit co-moving
volume in regions where the dark energy density was equal to ρD at the time of recombina-
tion. 2
Needless to say, the determination of both factors in the r.h.s. of Eq. (2) leaves room for
some uncertainties. However, we shall argue that there are reasons to be optimistic. If the
distribution (2) is to explain both cosmological constant problems, then a number of rather
generic predictions can be made, rendering these ideas very testable.
In the next section, we review the calculation of the prior probability distribution P∗(ρD).
The anthropic factor nciv(ρD) is discussed in section III. In the same section, we argue that
the anthropic approach can succeed only if the conditions for civilizations to evolve arise
mostly in galaxies formed at low redshifts, z ∼ 1. Anthropic predictions for the dark energy
1 Anthropic selection effects associated with the possible variation of the amplitude of density fluctuations
[10, 11] and of the baryon to photon ratio [10, 12] have also been discussed in the literature.
2 As we shall argue, in models where both cosmological constant problems can be solved anthropically, ρD
has not varied appreciably since the time of recombination, and therefore it can be treated as constant in
time.
3
equation of state, for the energy density ρD, and for the Hubble parameter h are discussed
in sections IV, V. The prediction for the future of the universe in unveiled in section VI.
Finally, our conclusions are briefly summarized and discussed in section VII.
II. THE PRIOR DISTRIBUTION
The first task in determining (2) is to estimate P∗. The vacuum energy density is of
order ρΛ ∼ η4 & (T eV )4, and therefore ρD must have a natural range of variation of order
η4 or larger. Weinberg noted [9] that a function P∗(ρD) that varies smoothly on scales
ρD ∼ η4, should behave as a constant in the utterly narrower interval (1) - unless of course,
the function would happen to have a zero or a pole in that interval (which would be an utter
coincidence). This led him to conjecture that for values of ρD in the anthropic range the
prior probability would be constant,
P∗(ρD) ≈ const.
(3)
Outside of this range the form of P∗ is irrelevant, because the factor nciv vanishes. Weinberg's
conjecture is subject to verification. As mentioned in the Introduction, P∗ is calculable,
provided that the dynamics of ρX is known, and assuming an inflationary model which
would determine its spatial distribution at the time of recombination. Analysis of explicit
models shows that (3) is not automatically guaranteed [4], but it does seem to be satisfied
in generic models.
There are basically two reasons [1, 4] why a non-flat P∗ may result from the process of
randomization of ρD which occurs during inflation (this randomization is due to quantum
diffusion in the case where X is a scalar field, or to nucleation of membranes in the case when
X is a four-form). The first reason is the differential expansion induced by the dark energy
component. During inflation, the expansion rate is determined by H 2 = (8πG/3)(Vinf +ρD).
Although ρD is very small compared with the inflationary potential Vinf , its effect may
build up over time, in such a way that more thermalized volume is generated with high
values of ρD. In this way, P∗(ρD) could be biased towards large values. Let us denote by
τ (X, H) the characteristic time needed for the dynamics of X to sample (at a fixed point in
space) all values of ρD within the anthropic range (∆ρD)anth. The differential expansion is
characterized by the parameter
q = (∆H)τ = (4πG/3)H −1(∆ρD)anthτ (X, H).
(4)
If q ≫ 1, then P∗ is exponentially steep in the range of interest. This case is ruled out
by observations, because it predicts a very large ρD, even after selection effects have been
factored in. If q ∼ 1, the distribution P∗ may have a moderate dependence on ρD within
the anthropic range. This dependence affects the position of the peak of the distribution for
the observed values of ρD, Eq. (2), and hence it affects our predictions. While models of
this sort are not ruled out, they require a very unnatural adjustment of parameters, since
q is determined by a combination of rather different pieces of dynamics. Hence, we shall
disregard this marginal possibility as non-generic. Finally, there is a wide class of models
where q ≪ 1 is satisfied without any fine tuning [1, 4], and hence we shall take this to
be the generic case. Numerical simulations confirm that in this case the bias effect due to
differential expansion is insignificant [18].
4
The second reason why P∗ may be non-flat is the following. Even if the differential
expansion is negligible, and the prior distribution for X is flat, this does not automatically
guarantee that the prior for ρD will be flat, unless the relation between X and ρD is linear
in the range of interest. Through this effect, it is possible to have a moderate variation of
P∗(ρD) within the anthropic range. But again, this would require a contrived adjustment
of parameters and we shall dismiss this case as non-generic (see also [19] for a discussion of
this issue).
As an example, let us consider the case where ρX = V (φ) is the potential energy density
of a scalar field φ,
ρD = ρΛ + V (φ).
(5)
The field must change very slowly on a cosmological time-scale, so that its potential energy
behaves as an effective cosmological constant. This requires the slow-roll conditions [4]
V ′ ≪ 10ρD/mp,
V ′′ ≪ 102ρD/m2
p
(6)
to be satisfied up to the present time (when ρD ∼ ρ0, with ρ0 the present matter density).
The constraint q ≪ 1 on the differential expansion yields [4]
V ′2H 4/GV 3 ≫ 1.
(7)
During inflation, the scalar field is randomized by quantum fluctuations, and at recombina-
tion it is distributed according to the "length" in field space,
P∗(φ)dφ ∝ dφ.
.
(8)
(9)
Therefore 3,
P∗(ρD)dρD ∝
dρD
V ′(φ)
Thus, the flatness of the prior depends on how much V ′ changes in the anthropic range.
As we shall see, variations in this range may occur, but they do not bias the probability
distribution for ρD in any significant way, unless we adjust some parameters specifically for
this purpose.
Consider a potential of the form
V (φ) =
1
2
µ2φ2,
(10)
where µ2ρΛ < 0, so that it is possible to have ρD very small even if ρΛ is large. Eqs. (6)
lead to the condition [4]
(11)
µ ≪ 10−120m3
pρΛ−1/2.
Such a small mass parameter may seem unrealistic, but it can naturally arise, for instance,
in a low energy effective theory with a suitable discrete symmetry [3] (for other proposals,
see [1, 19, 20] and references therein). Note that (11) does not correspond to a fine tuning,
but just to a strong suppression. The condition (7) translates into
µ ≫ H 3
0 H −2 ∼ 10−169mp,
(12)
3 Note that near the points where V ′(φ) = 0, we have ρD ≈ A + Bφ2 and V ′(φ) ∼ φ ∼ (ρD − A)1/2, which
is integrable. Hence, the zeroes of V ′(φ) are not a concern.
P∗(ρD)dρD =
κ + (M 4/η) cos(φ/η)−1 .
dρD
(17)
where H0 is the present Hubble rate, and in the last step we have used H ∼ 10−7mp,
corresponding to a GUT scale of inflation. The conditions (11) and (12) leave very many
orders of magnitude available for the parameter µ, and so fine tuning is not necessary. From
(5),
5
(13)
0 = −2ρΛ/µ2 and κ = µ2φ0. We are interested in the vicinity of ρD = 0, where it is
ρD = κ(φ − φ0) +
where φ2
easy to show from (9) that [4]
µ2
2
(φ − φ0)2,
P∗(ρD)dρD ∝ [1 + O(ρD/ρΛ)]dρD ≈ dρD.
(14)
Since ρD ≪ ρΛ in the anthropic range, the distribution is indeed flat to a very good accuracy.
For contrast, we may consider the "washboard" potential
where κ was given above and M and η are different mass scales. Let us assume that 4
ρD = ρΛ + κφ + M 4 sin(φ/η),
M 4 ≪ H 2
0 mpη ∼ (η/mp)ρ0.
(15)
(16)
Then the field will typically be found away from the local minima, with a probability dis-
tribution
In the case κ ≫ M 4/η, the distribution (17) is still flat, as in (3).
Both κ and M 4/η should be much smaller than H 2
0 mp in order to satisfy the slow roll
condition.
In the
opposite case, where M 4/η ≫ κ, the a priori distribution can have a sizable variation within
If η ≪ mp, this range is very wide in the field space,
the anthropically allowed range.
δφ & ρ0/κ ≫ mp. This means that the oscillations in P∗ will average out on scales much
smaller than the anthropic range, and effectively we recover (3). Clearly, the only way to
avoid this averaging effect is if η & mp, and
M 4 ∼ (∆ρD)anth.
(18)
The last equation is to ensure that a significant range of values of φ/η is sampled in the
anthropic range (∆ρD)anth . 103ρ0, so that changes in the slope of the potential are appre-
ciable. Otherwise the distribution for P∗ will be almost flat. Thus, aside from the fact that
the washboard potential is already a somewhat contrived example [19], Eq. (18) implies an
otherwise unnecessary adjustment of the parameter M.
In what follows, we shall only consider models where there is no such ad-hoc adjustment.
In this sense, our predictions may not be completely inescapable, but they can be considered
generic. The situation can be compared with the predictions of inflation that the density
parameter is Ω = 1 and the spectrum of density perturbations is nearly flat. It is certainly
possible, in the context of inflation, to have an open universe with Ω < 1, or to have
a markedly non-flat spectrum of density perturbations. But to achieve this, additional
parameters must be introduced and adjusted to the desired outcome.
4 If M 4 ≫ H 2
0 mpη, the slow roll condition is not satisfied today and the field φ will be in any one of the local
minima of the washboard. With some generic requirements on the inflationary parameters, the minima
will have equal a priori probability within the anthropic range [1].
6
III. THE ANTHROPIC FACTOR
We now consider the effect of the anthropic factor nciv in Eq. (2). The physical situation
is rather different for positive and negative ρD, so we consider these two cases separately.
For positive ρD, the main change introduced by nciv is that the time of earliest galaxy
formation tEG in the anthropic range (1) is effectively replaced by the time at which the
bulk of galaxy formation occurs. This is because a few early birds will not make a difference
once we apply the principle of mediocrity. More precisely, we should take into consideration
that the morphology of some galaxies could make them less suitable for the development of
civilizations, and therefore
nciv(ρD) =Z dα n(α, ρD) Nciv(α).
(19)
Here, α denotes the set of parameters characterizing the type of galaxy (e.g. its size, density,
etc.), n(α, ρD) is the number density of such galaxies that form per co-moving volume in
regions characterized by ρD, and Nciv(α) is the number of civilizations per galaxy of type α.
Suppose that the above integral receives a dominant contribution from galaxies of type αG.
Then
nciv(ρD) ∝ n(αG, ρD),
(20)
and the relevant time for anthropic considerations is the time at which this type of galaxies
form, which we shall denote by tG. With the assumption of a flat prior P∗, it was shown in
[11, 39] that the most probable value for a positive ρD is the one characterized by
This fact was used in order to explain the observed time coincidence
tD ∼ tG.
tD ∼ t0.
(21)
(22)
The last relation follows from (21), assuming that stars and civilizations develop on a
timescale not much greater than tG, and therefore tG is comparable to t0, defined as the
time when most civilizations make their first determination of ρD.
Connected with the above discussion, there is a prediction of the anthropic approach,
which can be checked by a combination of observations and theoretical analysis. In a not so
distant future, our understanding of galactic evolution and morphology may improve to the
point where we can tell with some confidence which galaxies are suitable for sustaining plan-
etary systems similar to our own, where civilizations can develop. The anthropic approach
to the cosmological constant problems (CCPs) predicts that the conditions for civilizations
to emerge will be found mostly in galaxies that formed (or completed their formation) at a
low redshift, z ∼ 1.
In the standard cold dark matter cosmology, galaxy formation is a hierarchical process,
with smaller objects merging to form more and more massive ones. We know from observa-
tions that some galaxies existed already at z = 5, and the theory predicts that some dwarf
galaxies and dense central parts of giant galaxies could form as early as z = 10 or even 20.
The fraction of matter bound in giant galaxies (M ∼ 1012M⊙) at z = 1 (∼ 20%) is some-
what less than that in objects of mass ∼ 109M⊙ at z = 3, or in objects of mass ∼ 107M⊙
at z = 5 [30]. If civilizations were as likely to form in early galaxies as in late ones, then
7
Eq. (21) would indicate that, for a typical observer, the cosmological constant should start
dominating at a redshift zG & 5. The corresponding dark energy density,
ρD ∼ (1 + zG)3ρ0,
(23)
would be far greater than observed. Clearly, the agreement becomes much better if we
assume that the conditions for civilizations to emerge arise mainly in the types of galaxies
which form at lower redshifts, zG ∼ 1.
We now point to some directions along which the choice of zG ∼ 1 may be justified. One
problem with dwarf galaxies is that if the mass of a galaxy is too small, then it cannot retain
the heavy elements dispersed in supernova explosions. Numerical simulations suggest that
the fraction of heavy elements retained is ∼ 30% for a 109M⊙ galaxy and is negligible for
much smaller galaxies [37]. The heavy elements are necessary for the formation of planets
and of observers, and thus one has to require that the structure formation hierarchy should
evolve up to mass scales ∼ 109M⊙ or higher prior to the dark energy domination. This gives
the condition zG . 3, but falls short of explaining zG ∼ 1.
Another point to note is that smaller galaxies, formed at earlier times, have a higher
density of matter. This may increase the danger of nearby supernova explosions and the rate
of near encounters with stars, large molecular clouds, or dark matter clumps. Gravitational
perturbations of planetary systems in such encounters could send a rain of comets from the
Oort-type cloud towards the inner planets, causing mass extinctions5.
Our own Galaxy has definitely passed the test for the evolution of intelligence, and the
principle of mediocrity suggests that most observers may live in galaxies of this type. Our
Milky Way is a giant spiral galaxy. The dense central parts of such galaxies were formed
at a high redshift z & 5, but their discs were assembled at z ∼ 1 or later [32]. Our Sun is
located in the disc, at a distance ∼ 8.5 kpc from the Galactic center6. If this situation is
typical, then the relevant epoch to use in Eq. (23) is the epoch zG ∼ 1 associated with the
formation of discs of giant galaxies.
The above remarks may or may not be on the right track, but we emphasize once again
that if CCPs have an anthropic resolution, then, for one reason or another, the evolution
of intelligent life should require conditions which are found mainly in giant galaxies, which
completed their formation at zG ∼ 1.
In order to estimate n(αG, ρD) in Eq. (20), we shall need a simple quantitative criterion
to specify the relevant type of galaxies. The most important parameter characterizing a
5 The cross-section for disruption of planetary orbits is much smaller, and it would take a rather substantial
increase of the density for this process to become statistically important. A.V. is grateful to David Spergel
for a discussion of this issue.
6 It has been noted [31] that this distance is close to the corotation radius, where the orbital velocity of the
stars coincides with the rotational velocity of the spiral pattern. In other words, the motion of the Sun
relative to the spiral arms is rather slow, and as a result, the periods between spiral arm crossings are
rather long (∼ 108 yrs). Spiral arms are the primary sites of supernova explosions. They are also rich in
giant molecular clouds, and are therefore very hazardous to life. It has been argued in [31] that spiral arm
crossings are responsible for the major mass extinctions observed in the fossil record. Then one expects
that habitable planetary systems are to be found mainly in the vicinity of the corotation radius, since
mass extinctions at a rate much greater than once in 108 yrs may be too frequent for intelligent life to
evolve. (Note that it took us 6.5 × 107 yrs to evolve since the last great extinction.)
8
galaxy is its mass M. For the Milky Way it is MM W ∼ 1012 M⊙ [33], and the above
discussion suggests that we identify the relevant galaxies with gravitationally bound halos
of this mass. (Note that this is also the typical mass of L∗ galaxies, which contain most of
the luminous stars in the Universe.) It should be recognized, however, that the choice of
this characteristic mass scale is somewhat uncertain, so we shall illustrate how our results
are affected by choosing a larger or a smaller mass.
Our Galaxy is a member of the Local Group cluster, whose mass has been estimated as
[34] MLG ∼ 4 × 1012 M⊙. It is conceivable that the gas captured in this cluster is later
accreted onto the member galaxies and thus affects the properties of their discs. There
seems to be no justification to consider larger mass objects, and we shall regard MLG as an
upper bound on the potentially relevant mass scales. On the lower mass end, we shall use
M ∼ 1011 M⊙, which is roughly the mass of the bright part of our Galaxy, up to ∼ 10 kpc
from the center. (We note that MM W is probably a more reasonable choice, because the
properties of the disc depend on the total mass of the halo [35].)
We now consider negative ρD. The scale factor of a universe filled with nonrelativistic
matter and dark energy with ρD < 0 is given by
a(t) = sin2/3(cid:18) t
tD(cid:19) ,
(24)
where tD ≡ (1/6πGρD)1/2. The matter density ρM initially decreases while the universe
expands, but at t = πtD/2, when it reaches the value ρM = −ρD, the universe stops its
expansion and starts recontraction. The matter density grows in the contracting phase, and
thus ρM ≥ ρD throughout the evolution. The structure formation in a universe with a
negative ρD proceeds as usual until t ∼ tD, but then the growth of density perturbations
accelerates during the contraction, so that all overdensities collapse to form bound objects
prior to the big crunch. For tD & t0, giant galaxies will form at about the same time as
they did in our part of the universe and will have similar properties (with a possible caveat
indicated below). However, for tD ≪ t0 halos of the galactic size will be forced to collapse
at a much earlier time t ∼ tD, and their density will therefore be much higher than that of
our Galaxy. This would probably make such halos unsuitable for life.
These considerations suggest that the anthropic factor effectively constrains tD to be in
the range
tD & t0
(25)
for both positive and negative ρD. There is, however, an additional factor that could make
negative ρD less probable. For ρD > 0, structure formation effectively stops at t > tD, and
the existing structures evolve more or less in isolation. This may account for the fact that
discs of giant galaxies take their grand-design spiral form only relatively late, at z ∼ 0.3.
The discs are already in place at z ∼ 1, but they have a very unsettled, irregular appearance
[32]. On the other hand, for ρD < 0 the clustering hierarchy only speeds up at t > tD, and
quiescent discs which may be necessary for the evolution of fragile creatures like ourselves
may never be formed.
Another factor to consider is the characteristic time tI needed for intelligence to develop.
For positive ρD, this factor is unimportant, since the time after the dark energy domination
is practically unlimited, but for negative ρD the available time is bounded by t < πtD, and
the effect of tI requires a closer examination.
We first note that tI ≪ t0 is unlikely, since then it is not clear why it took so long for
intelligence to develop on Earth. (The total time of biological evolution, from the origin of
9
life on Earth till present, is estimated at ∼ 3.5×109 yrs.) For tI ≫ t0, we note that the main
sequence lifetime of stars believed to be suitable to harbor life is t⋆ ∼ (5 − 20) × 109yrs ∼ t0
(see [11] for a discussion of this point). If tI ≫ t0 ∼ t⋆, most of these stars will explode as red
giants before intelligence has a chance to develop. Carter [6] has argued that this is the most
likely scenario7. In this case, the number Nciv is suppressed by a factor ∼ min{t⋆, tD}/tI ∼
t0/tI, where we have used (25) in the last step. For positive ρD, the suppression is by a
factor ∼ t⋆/tI, which is of the same order of magnitude.
We conclude that the precise value of tI has little effect on the relative probability of
positive and negative ρD. If the accelerated clustering hierarchy is detrimental for life, then
the probability for negative ρD is suppressed; otherwise the two signs of ρD are equally likely.
In either case, we should not be surprised that ρD is positive in our part of the universe. In
the following sections we shall focus on the positive values of ρD.
IV. PREDICTION FOR THE EQUATION OF STATE
A rather generic prediction of models where both CCP's are solved anthropically is that
the equation of state of dark energy is given by pD = wρD, with
w = −1 ± 10−5.
(26)
The error bars correspond to the precision to which the observable universe can be aproxi-
mated by a homogeneous and isotropic model. In models where ρX is the energy density of
a four-form field, this equation of state is guaranteed by the fact that the four-form energy
density is a constant and can only change by the nucleation of branes (other than that,
it behaves exactly like an additional cosmological constant). If ρX is a generic scalar field
potential, the slow roll conditions (6) are likely to be satisfied by excess, by many orders
of magnitude, rather than marginally. For instance, for the quadratic potential (10), these
conditions imply the constraint (11). It would be contrived to arrange for the condition to
be satisfied marginally, since the whole point of the present approach is to have ρΛ cancelled
regardless of its precise value (which is not known to us even by order of magnitude). If
the slow roll conditions are satisfied by excess by just more than three orders of magnitude,
then the kinetic energy of the scalar field will be less than its potential energy by more than
six orders of magnitude, and Eq. (26) follows.
There are certainly models for dark energy, some of them with anthropic input, were
(26) is not satisfied. For instance, Kallosh and Linde [13] recently considered a supergravity
model where the time coincidence problem is solved anthropically, and where (26) does
not hold. However, their model does not solve the old CCP, since it is assumed that the
cosmological constant vanishes in the observable matter sector due to some unspecified
mechanism. Likewise, (26) does not hold in the usual quintessence models [21], which have
no anthropic input at all, but which do not address the CCP's [1, 19], or in models of
k-essence [22], where only the time coincidence is partially addressed.
A possibility worth discussing is the case of models where the slow roll parameters are
themselves random fields. Consider, for instance, the following model:
ρD = ρΛ + µ2(ψ)φ2.
(27)
7 The coincidence tI ∼ t⋆ is unlikely, since the evolution of life and evolution of stars are governed by
completely different processes.
10
If the probability distribution for the new scalar field ψ were such that all values of µ2 are
equiprobable, then one might imagine that the order of magnitude of µ2 would be such that
the slow roll conditions would be marginally satisfied. 8
V. PREDICTIONS FOR ΩD AND h.
Currently favoured values for the dark energy density and for the Hubble parameter are
ΩD ≈ .7 and h ≈ .7 [24, 25], both with error bars of the order of 10%. While observations
are not very accurate, we would like to challenge the status quo and boldly use the anthropic
approach to the CCP's to make predictions for these two parameters. As we shall see, this
approach predicts that ΩD is likely to be somewhat higher, and that h is likely to be smaller
than those currently favoured values.
The basic reason why we expect ΩD to be larger is the following [14, 15]. The growth
of density fluctuations in a universe with a positive cosmological constant effectively stops
at the redshift zD when the cosmological constant starts dominating. This is given by
(1 + zD) ∼ (ΩD/ΩM )1/3, where ΩM = 1 − ΩD is the matter density parameter. According
to (21), we expect zD ∼ zG, where zG is the epoch when the relevant galaxies were formed.
(For
With zG ∼ 1, this corresponds to (ΩD/ΩM ) ∼ 8, which in turn implies ΩD ∼ .9.
zG > 1, we would obtain an even higher value for ρD.) This prediction can be made more
8 The new field ψ must also be a light field and hence its prior distribution is in principle calculable. We
can actually consider a more general form of the potential for ψ and φ,
ρD = ρΛ + V (ψ, φ).
(28)
Around any point (ψ0, φ0) on the curve γ defined by V (ψ, φ) = −ρΛ, the potential can be approximated
by a linear function of the fields. Moreover, we can always rotate coordinates in field space so that ψ is
directed along γ, and φ is orthogonal to it,
ρD ≈ Vφ(ψ0, φ0)(φ − φ0).
Here Vφ is the gradient of the potential at that point. During inflation, both fields ψ and φ are randomized
by quantum fluctuations. Hence, the prior probability distribution is given by the area in field space
P∗dψdφ ∝ dψdφ, which leads to
(29)
P∗(ρD)dρD =(cid:20)Zγ
dψ
Vφ(cid:21) dρD.
Along the curve γ, the values of Vφ that will carry more weight per unit distance along the curve are
those for which the slope is smaller. In the published version of this paper, it was (incorrectly) concluded
from this observation that "given a model where the slope of the potential is variable, smaller values of
the slope are preferred a priori, and there is no reason to expect that the slow roll conditions should be
satisfied only marginally." However, this is not necessarily correct in general, since the slope could be large
for a very large range of ψ, and large slopes may end up dominating the integral in Eq. (29). A detailed
analysis of this point has been given in [42], where it is shown that there is a wide class of two-field models
for which a small slope is favoured a priori (and for which the conclusions of Section IV do indeed apply).
However, there is an equally broad class of models for which a large slope is favoured a priori. In that
case, the posterior probability distribution for the slope is such that the slow roll condition is satisfied
only marginally, and some departure from the vacuum equation of state may be expected (see [43, 44]).
quantitative [17] by using the distribution (2). As we shall see, the precise predictions
depend not only on ΩD but also on h.
11
Throughout this section, we shall assume that ρD > 0 as part of our prior. In a universe
filled with pressureless matter and with a dark energy component ρD > 0, the scale factor
behaves as
a(t) = sinh2/3(cid:18) t
tD(cid:19) ,
where tD ≡ (1/6πGρD)1/2. A primordial overdensity will eventually collapse, provided that
its value at the time of recombination is larger than a certain value δrec
. In the spherical
collapse model, this is estimated as δrec
rec , where xrec = x(trec) [27]. Here, we
have introduced the variable
c (ρD) = 1.13x1/3
c
x(t) ≡
ΩD(t)
ΩM (t)
= sinh2(cid:18) t
tD(cid:19) .
(30)
The number of galaxies n(M, ρD) of mass M that will form per unit comoving volume in
a region characterized by the value ρD of the dark energy density, is proportional to the
fraction of matter that eventually clusters into this type of galaxies. In the Press-Schechter
approximation [26, 27], this is given by
nciv(ρD) ∝ n(M, ρD) ∝ erfc(cid:18) δrec
√2σrec(M)(cid:19) .
c (ρD)
(31)
Here, erfc is the complementary error function, and σrec(M) is the dispersion in the density
contrast at the time of recombination trec. As argued in the preceeding section, we shall
assume that most civilizations are formed in galaxies characterized by a mass M ∼ MM W ≈
1012M⊙ (although we shall also consider slightly larger and smaller masses).
The factor nciv depends on the parameter σrec, which in turn depends on the amplitude
of density perturbations generated during inflation. The value of σrec can be inferred from
the normalization of CMB anisotropies, but for this task, both the present value of ΩD and
the value of the Hubble parameter h would be needed. Since these are the parameters we
wish to make predictions about, it would be somewhat contrived to use them at this point
to make an inference about σrec.
Another factor to consider is that σrec may be different in distant regions of the universe
(where, as a consequence, galaxies would form earlier or later). In models where the inflaton
field has only one component, the value of σrec is the same in all regions of the universe.
However, if the inflaton field has more than one component, the amplitude of density per-
turbations depends on the path followed by the inflaton on its way to the minimum of the
potential. In such models, it is possible for σrec to vary over distances much larger than the
presently observable universe.
To make our discussion sufficiently general, we shall consider that σrec is itself a random
variable with unspecified prior. This prior may be determined by processes occurring during
inflation, or it may just reflect our ignorance of the actual value of the fixed parameter σrec.
Then, Eq. (2) is generalized to
dP(ρD, σrec) ≈ ncivP∗(ρD, σrec)dρDdσrec.
(32)
In this context, the generic expectation that the prior does not depend on ρD in the anthropic
range [see Eq. (3)], translates into
P∗(ρD, σrec) ≈ P∗(σrec).
dP
dln y
0.08
0.06
0.04
0.02
0.01
0.1
1
10
100
y
FIG. 1: The distribution (34).
Substituting (31) into (32), we have
dP(ρD, σrec) ∝ erfc .80x1/3
σrec !P∗(σrec) dxrecdσrec,
rec
12
(33)
where we have used that ΩM (trec) ≈ 1 in all regions of interest, so that dρD ∝ dxrec.
Introducing y = xrecσ−3
rec, the change of variables (xrec, σrec) → (y, σrec) produces a Jacobian
proportional to σ3
recP∗(σrec)dydσrec, where f (y) does
not depend on σrec. Integrating over σrec leads to the normalized distribution
rec, and we have 9 dP(y, σrec) ≈ f (y)σ3
dP(y) = (.80)3π−1/2erfc(.80y1/3) y d ln y,
(34)
which is uncorrelated with σrec.
The variable y can be expressed in terms of observable quantities, as we shall see below,
and from (34) we should expect y ∼ 1 by order of magnitude (See Fig. 1). More precisely,
9 The appearance of the factor σ3
rec in the posterior distribution was noted in Ref. [11]. This factor implies
that the prior P∗(σrec) should decay faster than σ−3
rec at large values of the density contrast. Otherwise,
the posterior distribution is not normalizable (and we should expect both a large value of the effective
vacuum energy and of the linear density contrast, in contradiction with observations). In the published
version of the present paper this factor was omitted, due to an incorrect normalization of Eq. (33). We
thank Takahiro Tanaka for drawing our attention to this point.
we expect y > .79 with probability
and y > .07 with probability
P (y > .79) = .68
(1σ c.l.),
P (y > .07) = .95
(2σ c.l.).
13
(35)
(36)
We shall denote these two equations as the 1σ and 2σ confidence level predictions for y. Let
us now show how these translate into confidence level curves for the expected values of the
parameters ΩD and h. Here, and in what follows, ΩD will denote the present value of the
dark energy density parameter in our observable universe.
Let us first express the "observed" value of y, which we shall denote as y0, in terms of
ΩD and h. The density contrast at present is given by σ0 = G(x0, xrec)σrec, where, assuming
zrec ≫ 1, the growth factor is given by [17] G(x0, xrec) = x−1/3
dw
rec F (ΩD), with
F (ΩD) =
5
6
Ω−1/2
D
w1/6(1 + w)3/2 .
(37)
Therefore,
0
Z ΩD/(1−ΩD)
(cid:21)3
y0 =(cid:20) F (ΩD)
σ0
.
(38)
The linearized density contrast at present σ0 can be inferred from measurements of CMB
temperature anisotropies, as described e.g.
in [17, 28]. Since the spectrum is expressed as
a function of wavelength, the mass scale has to be converted into a length-scale. A halo
of mass M corresponds to a co-moving radius R(M) = (3M/4πρ0)1/3. The mean matter
density of the universe is given by ρ0 = 1.88 × 10−29ΩM h2 g/cm3, which leads to
R(M) = .95h−2/3Ω−1/3
M (cid:18) M
1012M⊙(cid:19)1/3
Mpc.
Assuming an adiabatic primordial spectrum of scalar density perturbations, characterized
by a spectral index n, we have
σ0(R) = (c100Γ)(n+3)/2δHK 1/2(R).
(39)
Here, c100 = 2.9979 is the speed of light in units of 100km s−1 and
Γ = ΩM h exp[−Ωb(1 + √2hΩ−1
M )]
is the so-called shape parameter, with Ωb the density parameter in baryons. For numerical
estimates, we shall take Ωbh2 ≈ .02. The dimensionless amplitude of cosmological pertur-
bations inferred from the COBE DMR experiment is given by [28, 29]
δH = 1.91 × 10−5
Ω−.80−.05 ln ΩM
M
[1 − .18(1 − n)ΩD − .03rΩD].
(40)
The parameter r denotes the ratio of tensor to scalar amplitudes. Note that the effect of
tensors is to make δH a bit smaller (although not very significantly). Finally,
exp[1.01(1 − n)]
p1 + r(.75 − .13Ω2
D)
K(R) =Z ∞
0
q(n+2)T 2(q)W 2(q Γ h R Mpc−1)dq,
14
h
0.85
0.8
0.75
0.7
0.65
0.6
0.55
0.6 0.65 0.7 0.75 0.8 0.85
ΩD
2
σ
1PSfrag replacements
σ
FIG. 2: Contours of the function y0(ΩD, h) given in (38), corresponding to the 1σ (lower curves)
and 2σ (upper curves) predictions represented by Eqs. (35-36). The excluded region lies to the left
of the curves. The thick solid lines assume that the dominant contribution to nciv is in galaxies of
mass M = MM W = 1012M⊙. For comparison, we show the predictions for different choices of the
mass. The short dashed curves correspond to the mass of the local group MLG = 4× 1012M⊙, and
the long dashed curves correspond to the mass of the bright inner part of our galaxy M = 1011M⊙.
A scale invariant spectrum of density perturbations is assumed.
where the transfer function is given by T (q) = (2.34q)−1 ln(1 + 2.34q)[1 + 3.89q + (16.1q)2 +
(5.46q)3 + (6.71q)4]−1/4 and the window function is given by W (u) = 3u−3(sin u − u cos u).
Substituting (39) in (38), and using ΩD + ΩM = 1, we obtain the function y0 = y0(ΩD, h).
Contour lines of this function, corresponding to the 1σ and 2σ predictions represented by
Eqs. (35-36), are plotted in Fig. 2, assuming that the dominant contribution to nciv is in
galaxies of mass M = MM W = 1012M⊙ (thick solid lines). We also consider the predictions
for different choices of the mass, as discussed in Section III. The short dashed curves cor-
respond to the mass of the local group MLG = 4 × 1012M⊙, and the long dashed curves
correspond to the mass of the bright inner part of our galaxy M = 1011M⊙. The effect of
a tilt in the spectral index is plotted in Fig. 3. Both of these figures ignore the effect of
tensor modes in the normalization (40). Tensor modes tend to lower the value of δH, and
hence they tend to make the bounds somewhat less stringent. The effect, however, is not
dramatic. Even for r as large as .5, the effect on the curves is comparable to the effect of
lowering the spectral index by .05.
Expressions similar to Eqs. (34)-(38) were already contained in the exhaustive analysis of
the problem given by Martel, Shapiro and Weinberg (MSW) in [17], where σrec was treated
as a fixed parameter. However, our use of these expressions is somewhat different. MSW
15
0.6 0.65 0.7 0.75 0.8 0.85
ΩD
2
σ
1
σ
PSfrag replacements
h
0.85
0.8
0.75
0.7
0.65
0.6
0.55
FIG. 3: Effect of a tilt in the spectral index of density perturbations. As in Fig. 2, the thick solid
lines correspond to a scale invariant spectrum n = 1, and a mass M = MM W = 1012M⊙. The
long dashed line and the short dashed lines correspond to tilted spectra, with n = .95 and n = .9
respectively.
noted that the existing observations indicate a value of ΩD ∼ 0.6 − 0.7 and used Eq. (33),
with h = 0.7 to show that this range corresponds to probabilities from 2% to 12%, depending
on the values chosen for the galactic scale M and the spectral index of perturbations n. They
concluded that "anthropic considerations do fairly well as an explanation of a cosmological
constant with [ΩD] in the range 0.6 − 0.7". However, one cannot help but feel disappointed
by the somewhat low values of the probabilities.
Our approach here is that anthropic models should be used as any other models -- to
make testable predictions. Thus, the goal is not so much to explain the value of ΩD after it
is determined by observations, but to predict that value at a specified confidence level. The
contour lines in Figs. 2,3 indicate the 1σ and 2σ predictions of the model. If M ∼ MM W
proves to be the relevant mass giving the dominant contribution to nciv, then the currently
favoured model with ΩD ≈ .7 and h ≈ .7 is virtually excluded by the anthropic approach at
the 2σ level. Instead, this approach favours lower values of h and higher values of ΩD.
These predictions can be turned around. If the values ΩD ≤ 0.7, h ≥ 0.7 are confirmed
by future measurements, then our model will be ruled out at a 95% confidence level, again
assuming M ∼ MM W and a scale invariant spectrum. For a tilted spectrum, slightly lower
values of ΩD are allowed at the same confidence level. The observational situation at the
time of this writing is far from being clear. CMB and supernovae measurements yield [23, 24]
ΩD ≈ 0.7 , while the observations of galaxy clustering give [38] ΩM = 0.18 ± 0.8, and thus
ΩD ≈ 0.8.
16
VI. THE FUTURE OF THE UNIVERSE
We finally discuss the anthropic prediction which is not likely to be tested any time soon.
In all anthropic models, ρD can take both positive and negative values, so the observed
positive dark energy will eventually start decreasing and will turn negative, and our part of
the universe will recollapse to a big crunch.
To be specific, we shall consider a scalar field model with a very flat potential. In the
anthropic range (1), the potential can be approximated as a linear function,
where V ′
the evolution is described by the usual slow roll equations
0 is a constant and we have set φ = 0 at V = 0. Once the dark energy dominates,
V (φ) ≈ −V ′
0φ,
(41)
3H φ = V ′
0,
H 2 =
8π
3m2
p
V ′
0φ,
where H = a/a and a(t) is the scale factor. The solution of (42), (43) is
φ(t) = −φ∗[1 − (t/t∗)]2/3,
where −φ∗ is the present value of φ and
a(t) = exp[4πm−2
p (φ2
∗ − φ2(t))],
t∗ = 8πtD(φ∗/mp)2
(42)
(43)
(44)
(45)
(46)
is the time from the present to the beginning of recollapse.
The slow roll condition (6) implies that φ∗ & mp. As we discussed in Sec. IV, we do not
expect this condition to be only marginally satisfied, and thus φ∗ ≫ mp. Then it follows from
Eqs. (46) and (45) that t∗ ≫ 8πtD and therefore we should expect our region of the universe
to undergo accelerated expansion for at least another trillion years before recollapse.10
The slow roll approximation breaks down at φ ∼ −mp, so the above equations cannot be
used to describe the evolution at φ > 0, where the potential becomes negative. A general
analysis of models with negative potentials has been given in [36], where it is shown that at
φ2 ≫ V (φ).
φ ≫ mp the dynamics becomes dominated by the kinetic energy of the field,
The corresponding evolution is described by
φ(t) =
ln(tc − t) + const,
mp√6π
a(t) ∝ (tc − t)1/3,
(47)
(48)
where tc is the time of the big crunch. The linear approximation (41) for the potential breaks
down at sufficiently large φ, but in this regime the form of the potential is unimportant and
Eqs. (47), (48) still apply.
10 This is in contrast with the model of Kallosh and Linde [13] discussed in Section IV, where the universe
is expected to recollapse within 10-20 billion years.
17
During the dark energy dominated expansion, the ordinary nonrelativistic matter is di-
luted by the exponential factor (45). When the contraction starts, the density of matter
begins to grow as ρM ∝ (tc − t)−1. However, the kinetic energy of the field φ grows much
φ2 ∝ (tc − t)−2, and thus ordinary matter forever remains a subdominant component
faster,
of the universe.
VII. CONCLUSIONS
We now summarize the predictions that follow from the anthropic approach to the CCP's.
(1) In the simplest models where the dark energy density takes discrete values, or where
the dark energy density is due to the potential energy of a single scalar field, the dark energy
equation of state is predicted to be that of the vacuum,
pD = wρD,
(49)
where w = −1 with a very high accuracy. This distinguishes the anthropic models we
discussed here from other approaches, such as quintessence [21] or k-essence [22].11
(2) The anthropic predictions for the dark energy density ΩD and for the Hubble parame-
ter h are given in Figs. 2 and 3 of Section V12. We show the areas in the Ωd−h plane that are
excluded at 1σ and 2σ confidence levels. The excluded areas depend on the assumed galactic
mass M and on the spectral index n of the density fluctuations. For M = MM W = 1012M⊙
the currently popular values ΩD = 0.7, h = 0.7 are marginally excluded at 2σ confidence
level for a scale invariant spectrum n = 1. Lowering the spectral index relaxes the bounds
somewhat. For h > 0.65 and n > .95, the 1σ prediction is ΩD > 0.79. These anthropic
constraints get weaker when the relevant mass scale M is increased. For example, with
M = 4 × 1012M⊙ a value as low as ΩD = 0.63 is still allowed at the 2σ level for a scale
invariant spectrum. The 1σ prediction in this case is ΩD > 0.78 (for h = 0.65).
(3) Conditions for intelligent life to evolve are expected to arise mainly in giant galaxies
that form (or complete their formation) at low redshifts, zG . 1.
(4) The accelerated expansion will eventually stop and our part of the universe will
recollapse, but, at least in the framework of the simplest models, it will take more than a
trillion years for this to happen. Of course, this prediction is not likely to be tested anytime
soon. 13.
The above predictions apply to models where both CCP's are solved anthropically. For
comparison, we may consider other models. For instance, it is conceivable that a small value
of the cosmological constant will eventually be explained within the fundamental theory.
(We note the interesting recent proposal by Dvali, Gabadadze and Shifman [40] in this
regard.) Even then, the coincidence problem will still have to be addressed. One possibility
11 For a discussion of multifield models, see [42]. There is a class of two field models where the above
prediction does not hold. These are the models where the prior distribution favours a large slope of the
potential. In general, the equation of state pD = w(z)ρD depends on the redshift z, and increases with
time as the dark energy fields pick up kinetic energy, on their way to negative values of the potential. The
prediction for w(z) in the general case has been discussed in [43, 44].
12 These predictions are implicit in the earlier analysis by Martel, Shapiro and Weinberg [17]
13 Again, this prediction is somewhat different in two field models where the prior distribution favours a
large slope of the potential [42]
V (φ) = Λ(2 − cosh√2φ)
(50)
is that ρD is truly a constant, while the amplitude of the density fluctuations σrec is a
stochastic variable. With some mild assumptions about the prior probability distribution
P∗(σrec), it can be shown [1] that most galaxies are then formed at about the time of vacuum
domination. In this class of models, predictions (1) and (3) still hold, while the other two
predictions no longer apply.
Another possibility has been recently discussed by Kallosh and Linde [13]. They assumed
18
an M-theory inspired potential
with a stochastic variable Λ. An interesting property of this potential is that its curvature
is correlated with its height (at φ = 0). As a result, the universe tends to recollapse
within a few Hubble times after the dark energy comes to dominate. Assuming that other
contributions to the vacuum energy are somehow cancelled (that is, that the old CCP
is solved by some unspecified mechanism), Kallosh and Linde argue that the coincidence
tD ∼ tI is to be expected, where tI is the time it takes intelligent life to evolve (they assume
it to be ∼ 1010 yrs). Predictions (1)-(3) are not applicable to this model. The model does
predict recollapse of the universe, but the corresponding timescale (∼ 1010 yrs) is much
shorter than the anthropic prediction (4).
Anthropic arguments are sometimes perceived as handwaving, unpredictive and unfal-
sifiable lore, of questionable scientific validity.
In our view, the results presented in this
paper should dispel this notion. Here, we have used the anthropic approach to make sev-
eral quantitative predictions, some of which may soon be checked against observations. It
should also be emphasized that, for the particular case of dark energy, there are at present
no alternative theories explaining both CCP's, or making generic predictions of comparable
accuracy.
The present bound on the equation of state parameter w from the CMB and supernovae
measurements is [41] w < −0.7, which is consistent with the anthropic prediction of w = −1.
The value of w = −1 is usually associated with a plain cosmological constant. However,
if in addition to this equation of state, observations confirm some of the other predictions
presented above, this may be taken as an indication that the dark energy is dynamical. Thus,
a better understanding of structure formation and galactic evolution may in fact reveal a
crucial property of dark energy, with important implications for particle physics.
Acknowledgements
We are grateful to Steven Barr, David Spergel and Rosanne Di Stefano for useful discus-
sions. This work was supported by the Templeton Foundation under grant COS 253. J.G. is
partially supported by MCYT and FEDER, under grants FPA2001-3598, FPA2002-00748.
A.V. is partially supported by the National Science Foundation.
[1] J. Garriga and A. Vilenkin, Phys. Rev. D64 023517 (2001).
[2] J.D. Brown and C. Teitelboim, Nucl. Phys. 279, 787 (1988). For recent discussions, see also
J. L. Feng, J. March-Russell, S. Sethi and F. Wilczek, Nucl. Phys. B602, 307 (2001); R.
Bousso and J. Polchinski, JHEP 0006:006 (2000); T. Banks, M. Dine and L. Motl, JHEP
0101:031 (2001).
19
[3] G. Dvali and, A. Vilenkin, Phys. Rev. D64, 063509 (2001).
[4] J. Garriga and A. Vilenkin, Phys. Rev. D61, 083502 (2000).
[5] B. Carter, In I.A.U. Symposium 63: confrontation of cosmological theories with observational
data (ed. M.S. Longair), p. 291. Dordecht:Reidel (1974).
[6] B. Carter, Phil. Trans. R. Soc. Lond. A 310, 347 (1983).
[7] J.D. Barrow and F.J. Tipler, the Anthropic Cosmological Principle (Oxford University Press,
New York, 1986).
[8] For an early attempt to apply anthropic arguments to the cosmological constant, see also
P.C.W. Davies and S. Unwin, Proc. Roy. Soc. 377, 147 (1981).
[9] S. Weinberg, Phys. Rev. Lett. 59, 2607 (1987).
[10] M. Tegmark and M. Rees, Astrophys. J. 499, 526 (1998).
[11] J. Garriga, M. Livio and A. Vilenkin, Phys. Rev. D61, 023503 (2000).
[12] A. Aguirre, Phys. Rev. D64, 083508 (2001).
[13] R. Kallosh and A. Linde, "M-theory, Cosmological Constant and the anthropic principle",
hep-th/0208157 (2002).
[14] A.Vilenkin, Phys. Rev. Lett. 74, 846 (1995).
[15] G. Efstathiou, M.N.R.A.S. 274, L73 (1995).
[16] S. Weinberg, in Critical Dialogues in Cosmology, ed. by N. G. Turok (World Scientific, Singa-
pore, 1997).
[17] H. Martel, P. R. Shapiro and S. Weinberg, Ap.J. 492, 29 (1998).
[18] V. Vanchurin, A. Vilenkin and S. Winitzki, Phys. Rev. D61, 083507 (2000).
[19] S. Weinberg, Phys. Rev. D61 103505 (2000); "The cosmological constant problems," Talk
given at 4th International Symposium on Sources and Detection of Dark Matter in the Universe
(DM 2000), Marina del Rey, California.
[20] J. Donoghue, JHEP 0008:022 (2000).
[21] P.J.E. Peebles and B. Ratra, Ap. J. Lett. 325, L17 (1988); R.R. Caldwell, R. Dave and P.J.
Steinhardt, Phys. Rev. Lett. 80, 1582 (1998).
[22] C. Armendariz-Picon, V. Mukhanov and P. J. Steinhardt, Phys. Rev. D 63, 103510 (2001).
[23] J. L. Sievers et. al., astro-ph/0205387; W.J. Percival et. al., astro-ph/0206256.
[24] S. Perlmutter et al., Ap.J. 483, 565 (1997); S. Perlmutter et al., astro-ph/9812473 (1998);
B. Schmidt et al., Ap.J. 507, 46 (1998); A. J. Riess et al., A.J. 116, 1009 (1998).
[25] W. L. Freedman et al., arXiv:astro-ph/0012376.
[26] W.H. Press and P. Schechter, Astrophys. J. 187,425 (1974).
[27] H. Martel and P.R. Shapiro, astro-ph/9903425.
[28] A. R. Liddle and D. H. Lyth, Cambridge, UK: Univ. Pr. (2000) 400 p.
[29] E. F. Bunn and M. J. White, Astrophys. J. 480, 6 (1997).
[30] H.J. Mo and S.D.M. White, astro-ph/0202393.
[31] E.M. Leitch and G. Vasisht, astro-ph/9802174.
[32] R.G. Abraham and S. van der Bergh, Science 293, 1273 (2001).
[33] M.I. Wilkinson and N.W. Evans, Mon. Not. Roy. Astron. Soc. 310, 645 (1999).
[34] P.J.E. Peebles, Principles of Physical Cosmology, Princeton University Press, Princeton, 1993.
[35] H.J. Mo, S. Mao and S.d.m. White, MNRAS 304, 175 (1999).
[36] G. Felder, A. Frolov, L. Kofman and A. Linde, hep-th/0202017.
[37] M. Mac Low and A. Ferrara, astro-ph/9801237.
[38] N. Bahcall et. al. astro-ph/0205490.
[39] S. Bludman, Nucl. Phys. A663-664,865 (2000).
[40] G. Dvali, G. Gabadadze and M. Shifman, hep-th/0202174.
[41] J.R. Bond et. al., astro-ph/0210007.
[42] J. Garriga, A. Linde and A. Vilenkin, arXiv:hep-th/0310034.
[43] R. Kallosh, J. Kratochvil, A. Linde, E. V. Linder and M. Shmakova, JCAP 0310 (2003) 015
[arXiv:astro-ph/0307185].
[44] S. Dimopoulos and S. Thomas, Phys. Lett. B 573 (2003) 13 [arXiv:hep-th/0307004].
20
|
astro-ph/9709283 | 1 | 9709 | 1997-09-28T18:24:12 | Image Separation vs. Redshift of Lensed QSOs: Implications for Galaxy Mass Profiles | [
"astro-ph"
] | Recently, Park and Gott reported an interesting observation: image separation of lensed QSOs declines with QSO redshift more precipitously than expected in any realistic world model, if the lenses are taken to be either singular isothermal spheres or point masses. In this Letter I propose that the observed trend arises naturally if the lensing galaxies have logarithmic surface mass density profiles that gradually change with radius. If the observed lack of central (odd) images is also taken into account, the data favor a universal dark matter density profile over an isothermal sphere with a core. Since the trend of image separation vs. source redshift is mostly a reflection of galaxy properties, it cannot be straightforwardly used as a test of cosmological models. Furthermore, the current upper limits on the cosmological constant may have to be revised. | astro-ph | astro-ph |
Mon. Not. R. Astron. Soc. 000, 000 -- 000 (1997)
Printed 28 December 2018
(MN LATEX style file v1.4)
Image Separation vs. Redshift of Lensed QSOs:
Implications for Galaxy Mass Profiles
Liliya L. R. Williams⋆
Institute of Astronomy, Madingley Road, Cambridge, CB3 0HA
Accepted for publication in MNRAS
ABSTRACT
Recently, Park and Gott reported an interesting observation: image separation of
lensed QSOs declines with QSO redshift more precipitously than expected in any
realistic world model, if the lenses are taken to be either singular isothermal spheres
or point masses. In this Letter I propose that the observed trend arises naturally if the
lensing galaxies have logarithmic surface mass density profiles that gradually change
with radius. If the observed lack of central (odd) images is also taken into account,
the data favor a universal dark matter density profile over an isothermal sphere with
a core. Since the trend of image separation vs. source redshift is mostly a reflection of
galaxy properties, it cannot be straightforwardly used as a test of cosmological models.
Furthermore, the current upper limits on the cosmological constant may have to be
revised.
Key words: gravitational lensing -- cosmology -- galaxies: structure
1
INTRODUCTION
Since gravitational lensing takes place over cosmological dis-
tances, it can in principle be used to measure cosmologi-
cal parameters, such as mass density, Ω, and cosmological
constant, Λ. For example, the frequency of multiply imaged
QSOs is a sensitive function of Λ, and therefore the observed
abundance of lensed QSOs has been used to place upper lim-
its on Λ (Turner 1990; Fukugita et al. 1992; Kochanek 1996).
Another example is the relation between image separation,
θim and source redshift, zs of multiply split QSOs, which
mostly depends on the sum of Ω and Λ, and thus should
serve as a good indicator of the curvature of the Universe.
(Gott, Park & Lee 1989; Fukugita et al. 1992, Park & Gott
1997, hereafter PG97). If the Universe is flat and galaxies
are singular isothermal spheres, the θim -- zs relation should
be flat, i.e. independent of source redshift. If the Universe
is open, there should be a mild decline of image separations
with redshift; a ∼ 20% decline is expected between zs ∼ 1.5
and 5, in an extreme Ω = 0 case.
The mildness of the expected θim -- zs trend can be ex-
plained as follows. For any given lens, and fixed lens and
source redshifts the image separation of images with compa-
rable brightness is quite independent of source impact pa-
rameter. Since optimal lens redshift for any high redshift
source is roughly the same, zl ∼ 0.5, and angular diame-
ter distances vary little past z ∼ 0.5, the observed image
separation should change little with zs. If a population of
lenses is considered, having a range of properties, like scale
lengths and velocity dispersions, there will be a correspond-
ing spread in image separations, but still little or no trend
with zs.
Yet the observed θim of lensed QSOs declines strongly
with zs (See Figure 1 of PG97, or Figure 2 of this paper.)
PG97 present a list of source redshifts and image separations
of 20 multiply imaged QSOs. The source redshifts span a
range between 1.4 and 4.5, and image separation between
0.5′′ and 7′′. There is a drop of at least a factor of 4 in image
separation from the low-zs to high-zs cases. According to
PG97, in a flat Universe such distribution can be ruled out
at 99% confidence level, while in the most extreme empty
Universe, it is ruled out at 97%. To make the confidence
level decrease below 95%, one would need to 'remove' 2 -- 3
largest separation or highest redshift cases from the sample.
(See PG97 for a discussion of statistical significance, possible
errors, etc.)
In this Letter, I suggest that the trend, which is due
to the lack of high redshift wide separation lensed QSOs is
naturally reproduced if a set of three conditions is met, (i)
galaxies' central mass profiles have logarithmic slopes that
change with radius, (ii) there is a dispersion in galaxy prop-
erties, like central surface mass densities or velocity disper-
sion, and (iii) the characteristic length scale of galaxy dark
matter halos varies with galaxy luminosity as r ∝ La, where
0 < a <∼ 0.5.
⋆ Email: [email protected]
c(cid:13) 1997 RAS
2
L. L. R. Williams
2 MODEL
In this section I describe the lensing model, and introduce
some simplifying assumptions about the lenses, galaxies, and
sources, QSOs. I do not aim to produce a detailed model be-
cause the small number of lensed QSOs in the current sample
does not necessitate it. I need not assume any particular cos-
mological model, because the effect implied by the present
model is considerably stronger than that of cosmology.
2.1 Model Assumptions
2.1.1 Lenses
I consider two types of galaxy mass profile, an isothermal
sphere with a core, ISC, and a universal dark matter pro-
file, NFW, derived from numerical simulations of Navarro,
Frenk & White (1995, 1996). These profiles were chosen be-
cause their logarithmic surface mass density gradually flat-
tens with decreasing radius. The reason for using these,
as opposed to constant power law profiles, like a singular
isothermal sphere, is discussed in Section 3. Both ISC and
NFW provide a plausible description of the dark matter ha-
los of galaxies. ISC is commonly used because of the ob-
served shapes of the rotation curves of spiral galaxies. Addi-
tionally, based on the results of the HST Snapshot Survey,
Maoz & Rix (1993, hereafter MR93) conclude that ellipticals
must possess ISC-type dark matter halos in order to repro-
duce the observed image separation distribution of multiply
imaged QSOs. The NFW profile has been recently claimed
as a universal dark matter halo profile describing halos of
gravitationally bound structures over 4 orders of magnitude
in mass. It is also supported on theoretical grounds (Evans &
Collett 1997). I assume that all galaxy-lenses are circularly
symmetric.
A description of the lensing properties of these profiles
can be found in Schneider et al. (1992), Bartelmann (1996),
and Williams & Lewis (1997). It suffices to say here that
ISC is roughly flat within core radius rc, and has an isother-
mal density profile outside. The vertical normalization of
the profile is fixed by its central surface mass density, κ0,
in terms of critical density for lensing, Σcrit. NFW is sin-
gular at the centre, its logarithmic slope gradually steepens
from ρ ∝ r−1 at the centre, to ρ ∝ r−3 at large radii, and
is isothermal around the characteristic scale length, rs. The
vertical normalization of its projected surface mass density
is proportional to κs, which is given in terms of Σcrit.
As was pointed out by PG97, the scatter in the observed
θim -- zs points is large. One of the main sources of scatter is
probably the spread in galaxy-lens properties. For example,
a singular isothermal sphere lens of a given σv produces a
constant bending angle at the lens, and hence image sepa-
ration varies as Dls/Dos, where the D's are the lens -- source
and observer -- source angular diameter distances. If all lens-
ing galaxies had the same σv, one would expect the observed
angular image separations to be proportional to the corre-
sponding Dls/Dos values, whereas a plot of Dls/Dos vs. θim,
in 8 systems where both zs and zl are known reveals no cor-
relation at all, and a scatter in θim of a factor of a few larger
than the scatter in Dls/Dos. This strongly suggests that the
spread in galaxy-lens properties is important. Therefore I as-
sume a family of galaxy-lenses, each having the same scale
length, rc and rs for ISC and NFW respectively, but a range
of κ0 and κs, which is equivalent to a range in velocity dis-
persions.
I assume that rc and rs and galaxy luminosity function
do not evolve with redshift, at least in the interval most
relevant here, i.e. optimal lens redshift for high-z sources,
zl ∼ 0.3 − 1. This assumption is supported by the recent
spectroscopic observations of Lilly et al. (1995), which indi-
cate that the luminosity function of elliptical galaxies, i.e.
the galaxy population that is believed to provide the bulk
of the lensing optical depth (see MR93), evolves very little
from z ∼1 to the present.
When the source impact parameter is sufficiently small,
all the NFW lenses, and ISC lenses with κ0 > 1 produce
3 images, whereas the observed multiply imaged systems
mostly have an even number of images, either 2 or 4. This
is thought to be because the central (odd) image is demag-
nified below visibility. I will return to the importance of
the central odd image later, in Section 2.2. The circularly
symmetric lenses considered here cannot produce 5 image
systems; observed 4(+1) cases are a property of elliptical
lenses. However the present treatment will not suffer if only
symmetric lenses are considered.
For sources at large redshifts, the optimal lens redshift
increases very slowly with zs in all cosmologies except for
the most extreme Λ dominated cases. So I will assume that
zl is constant, independent of zs. The critical surface mass
density for lensing, Σcrit changes little with zs, if zl is fixed,
and sources are at high zs. For example, in an Ω=0.3 open
Universe, moving the source from zs=1 to 5 reduces Σcrit by
17%, if zl = 0.2. Therefore, I will assume that galaxy-lens
projected surface mass density is given by κ0 and κs, with
no dependence on redshifts.
2.1.2 Sources
Analytic fits to QSO luminosity function (QLF) usually take
the form of a double power law, with steep bright-end slope
and shallow faint-end slope; the transition occurs at some
characteristic luminosity, L0, which evolves with redshift,
and is commonly parameterized as L0(zs) ∝ (1+zs)α. Boyle
et al. (1990) derive a pure luminosity evolution for z <∼ 2
QSOs, with α=3.2, while Hewett et al. (1993) conclude that
QLF changes shape with redshift, and that evolution must
slow down beyond z ∼ 1.5 compared to the predictions of
Boyle et al. For 2 <∼ z <∼ 3 they derive α ∼1.5. QSO evolu-
tion at higher redshifts, up to 5 is less constrained, though
it is sometimes assumed that the shape of the QLF does
not evolve beyond z ∼ 2, while L0 evolves such that it
'slides' brightward along the high luminosity part of the
QLF (Wallington & Narayan 1993). I assume that at any
given redshift there are no QSOs brighter than L0, and that
QLF at L < L0 is NQSO(L) ∝ L−s, with s = 1.2 (MR93).
I adopt the α parameterization of QSO evolution for high
redshifts, and derive results for α=0, and 2.
2.2 Results
image separation and total
We are interested in two lensing properties of the galaxy-
lenses:
image magnification.
Given a galaxy-lens, a source can have a range of impact
parameters, each resulting in a different image magnifica-
tion, and image separation, even though the latter tends to
c(cid:13) 1997 RAS, MNRAS 000, 000 -- 000
stay rather constant, roughly equal to twice the Einstein
ring radius of the lens. To account for a range of source im-
pact parameters, I calculate the impact parameter weighted
averages of image magnification, hµi and image separation,
hθi as a function of κ0 or κs, for ISC and NFW respectively.
Figure 1 shows the results. Vertical normalization of im-
age separation is irrelevant for now. The important feature
is that for both profiles the image separation increases with
κ0 or κs, while total image magnification decreases. In other
words, given a population of galaxy-lenses with a range of κ0
or κs, image magnification goes inversely with image sepa-
ration. Now, image magnification couples to source redshift,
as follows. Flux of an L0 QSO is given by,
f ∝ µ(zs)L0(zs)/DL(zs)2
∝ µ(zs)(1 + zs)α/DA(zs)2(1 + zs)4
∝ µ(zs)(1 + zs)α−4,
(1)
where µ(zs) is total image magnification, and DL and DA
are the luminosity and angular diameter distances respec-
tively. In the second and third lines I used the fact that
DL = (1 + z)2DA ∝ (1 + z)2, since angular diameter dis-
tance stays roughly constant for high-zs sources. So to make
it into a magnitude limited sample, with a certain flim, an
L0 QSO at zs has to be magnified by at least
µ(zs) ∝ flim(1 + zs)4−α,
(2)
with fainter QSOs magnified more. Because larger redshift
QSOs undergo larger magnifications, a fixed flim of any
given survey translates into a lower limit on hµi as a func-
tion of zs, which in turn implies an upper limit on hθi, as
read off from Figure 1. The resulting relation is plotted as
solid (ISC) and dashed (NFW) curves in Figure 2, for two
values of QSO luminosity function evolution parameters, α;
0, and 2. Both sets of curves reproduce the upper envelope
of the observed points quite well. The paucity of observed
points does not allow to differentiate between α of 0 and 2,
or between ISC and NFW profiles.
The curves are scaled vertically and horizontally to
roughly match the observed distribution. Adjusting the scal-
ing in the vertical direction yields the angular size of rc
and rs respectively. For ISC angular core radius is 0.37′′,
while for NFW angular scale radius is 1.25′′. For typical lens
redshifts of 0.5 these translate into rc = 1.4h−1 kpc, and
rs = 4.6h−1 kpc. These values are roughly what one would
derive from modelling individual multiple-image systems us-
ing ISC/NFW profiles. Scaling the curves in the horizontal
direction gives the average minimum magnification of lensed
QSOs as a function of zs. For α = 0 models, hµi/flim =
0.039(1 + zs)4, and for α = 2, hµi/flim = 0.4(1 + zs)2; i.e. if
QSO characteristic luminosity does not evolve, a lower limit
on magnification of QSOs at z = 3 is hµi/flim=10, whereas
it is 6.4 if L0(zs) ∝ (1+zs)2. Surveys with brighter flim have
higher hµi, but the dependence on (1+zs) is weak, and most
existing surveys have flim of 1 -- 2 magnitudes within each
other.
Let us summarize the conclusions thus far. Roughly,
Figure 1 implies that average image magnification is in-
versely proportional to the galaxy central surface mass den-
sity, hµi ∝ κ−1
0,s; and that average image separation is di-
rectly proportional to the galaxy central surface mass den-
sity, hθi ∝ κ0,s. In general, the observed image separation
is also proportional to the characteristic scale length of the
c(cid:13) 1997 RAS, MNRAS 000, 000 -- 000
Image Separation of Lensed QSOs
3
Figure 1. Magnification and image separation vs. the central
surface density of two different types of lenses: ISC and NFW.
Both magnification and image separation are impact parame-
ter weighted averages of these quantities. The dotted line is the
impact parameter weighted average magnification of a singular
isothermal sphere. Note that for ISC and NFW profiles the mag-
nification and image separation go in the opposite sense with
increasing κ0 and κs, while for SIS the average magnification is
always a constant (see Section 3).
galaxy-lens, so hθi ∝ κ0,src,s, but since I have so far assumed
that all galaxies have the same characteristic scale length, I
have not used the latter dependence. The scale length can
be a function of central concentration, rc ∝ κβ
0,s. Combin-
ing these relations gives, hµi ∝ hθi− 1
1+β , and together with
eq.(2) it implies,
hθi ∝ (1 + zs)−(1+β)(4−α)
(3).
With β = 0 this equation basically reproduces the curves
in Figure 2. Note that β does not have to be 0, any value
between roughly -0.5 and 2 will do, i.e. the total exponent in
eq. (3) should be around -- few. I will return to the discussion
of β and its implications for the dark matter halos of galaxies
in Section 4.
Thus both ISC and NFW galaxy-lens models can re-
produce the observed θim -- zstrend, and neither is preferred
based on these observations alone. However, another widely
documented observation can break the tie; all multiply im-
aged QSOs lack the central odd-numbered image, i.e. ob-
served cases are either doubles or quadruples. This has long
been interpreted as evidence in favor of centrally condensed
galaxy profiles, which would demagnify the central image
below visibility. The other possible explanation, dust obscu-
ration in the centres of galaxies, is largely ruled out because
radio observations of multiply split QSOs also do not reveal
central images (Myers et al. 1995). I apply this argument
here: ISC model predicts that κ0 ∼ 1 − 2 models should
have central images of brightnesses comparable to the pri-
mary image, while with the NFW models the central images
of low κs lenses should be 5-10 times demagnified compared
4
L. L. R. Williams
Figure 2. Image separation vs. source redshift. The points are the
observed multiply imaged QSOs (taken from Table 1 of PG97);
stars represent the lenses from the HST Snapshot Survey (MR93),
circles are from other, ground based surveys. Predictions of the
isothermal sphere with a core, ISC and universal dark matter
halo, NFW models (labeled) were normalized vertically and hori-
zontally, as described in Section 2.2. Both ISC and NFW profiles
can account for the observed trend.
to the primary image. This rules strongly in favor of the
NFW model.
3
IMPLICATIONS FOR GALAXY MASS
PROFILES
The key feature of the galaxy lens population that allows
to reproduce the envelope is that image separation goes in-
versely with image magnification. This allows magnification
bias, which acts through image magnification, to couple im-
age separation to source redshift. This property of a lens
population is a direct consequence of the individual galax-
ies having a changing logarithmic slope, as in the ISC and
NFW models. Lenses that do not have this property, for ex-
ample singular isothermal spheres and point masses, cannot
reproduce the observed envelope.
For a singular isothermal sphere (SIS) the average im-
pact parameter weighted magnification can be calculated to
be 4, independent of the velocity dispersion, σv, while image
separation is proportional to σ2
v. So hµi, being a 'universal'
constant, does not couple to hθi, and thus SIS cannot repro-
duce the observed trend. The property that hµi is constant
is not unique to SIS, but can be generalized to any family of
mass profiles with no in-built scale length, like single power-
laws and point masses. In such models the only parameter is
the absolute mass normalization of the profile, which deter-
mines the size of Einstein ring radius, and hence the image
separation. However, hµi in such self-similar profiles is a con-
stant, independent of absolute mass normalization.
Therefore galaxies must have changing logarithmic sur-
face mass density profiles in order to be able to reproduce
the θim -- zs relation.
4 DISCUSSION
There are two interesting aspects in the observed distribu-
tion of image separations vs. source redshift. In this Letter I
set out to explain one of these, the lack of high redshift wide
separation lenses, i.e. the θim -- zs anticorrelation. It turns out
that the model presented in Section 2 also naturally explains
the other interesting feature of the θim -- zs plot, namely the
existence of wide separation cases; there are 8 lensed QSOs
with θim > 3′′. Most models currently found in the literature
have trouble predicting a large population of wide separation
lenses; for example MR93 predict the peak in θim distribu-
tion at <∼ 1′′, with vanishingly small number of cases above
3′′. So it has been argued that large θim cases are not due to
isolated galaxies, but are the result of cluster-aided galaxy
lensing. QSO 0957+561, with 6.1′′ between its two images,
is adduced as supporting evidence. However, if cluster-aided
galaxy lensing is the correct explanation, then wide sepa-
ration lenses should be found at all redshifts, and not just
at low zs, as is currently the case (Figure 2). Furthermore,
PG97 show that with the help of cluster lensing, image sep-
aration should increase with zs. Since this is clearly not ob-
served, cluster aided lensing is probably not important in
general (0957 must be a special case), and the large number
of wide separation lenses needs an explanation.
Let us derive the distribution of image separations as
predicted by the model of Section 2. (I will use the ISC
model because all the derivations required in this section
can be carried out analytically with this model. However
the general arguments presented here will also apply to the
NFW profile.) So far only the dark matter halo properties
have been discussed. These need to be related to the observ-
able galaxy properties, like luminosity. I will adopt relations
similar to those used in MR93 (their Section 2.2.1);
rc ∝ La, M/L ∝ Lb, hence κ0 ∝ L1−2a+b
(4).
Optical effective radii of ellipticals are observed to scale as
L1.2 (Lauer 1985), and since it is commonly assumed that
they are linearly proportional to the galaxies' dark mat-
ter scale length, a is usually taken to be 1.2. From the
observations of the fundamental plane of ellipticals (Kor-
mendy & Djorgovski 1989), b ≈ 0.25. Similar assumptions
were made by Kochanek (1996). Parameter β of eq. (3) is
β = a/(1 − 2a + b)†.
The frequency of multiply imaged QSOs as a function
of image separation is given by,
dF
dhθi
=
dN (L)
dL
[rcyr(L)]2
dL
flim (cid:17)s
dhθ(L)i (cid:16) hµ(L)i
,
(5)
where dN (L)/dL ∝ L−1.2e−L is the Schechter luminosity
function (Schechter 1976); L is in units of L∗, and I assume
a constant slope of -1.2. The lensing cross section for a galaxy
of luminosity L is given by the radial caustic in the source
† Note that MR93 and Kochanek (1996) assumed a=1.2, and
b = 0.25, and hence effectively their β = −1.04, which would not
reproduce the θim -- zs anticorrelation, see eq. (3).
c(cid:13) 1997 RAS, MNRAS 000, 000 -- 000
plane, whose radius is yr(L) = (κ2/3
0 − 1)3/2. The image
separation is approximately equal to the diameter of the
tangential critical curve in the lens plane, hθi ≈ 2rcpκ2
0 − 1.
Using eq. (4),
dhθ(L)i
dL
=
2rc,∗La−1(κ2
(κ2
0[1 − a + b] − a)
0 − 1)
.
(6)
Here, rc,∗ and κ0,∗ refer to the core radius and κ0 of an L∗
galaxy. I take rc,∗ to be 1.4h−1kpc, equal to the constant
core radius derived in Section 2.2. The value of κ0,∗ is then
calculated from rc,∗ and an assumed asymptotic line of sight
velocity dispersion, σ∗ = 300km s−1; κ0,∗ = 5. The last term
in eq. (5) is the magnification bias.
The function dF/dhθi of eq. (5) is plotted in Figure 3,
for a range of a corresponding to the allowable range of β, in
eq. (3), a=0.0, 0.2, 0.4, and 0.6; b = 0.25 was used through-
out. The angular splitting by an L∗ galaxy is denoted by an
arrow. These predicted distributions apply to the source red-
shift range where the upper envelope, described by eq. (3)
and plotted in Figure 2 does not cut in, i.e. for zs <∼ 2.5. The
solid histogram is the observed distribution for the same
source redshift cutoff. Even though the small number of cur-
rently observed lenses does not allow to make any precise
conclusions, it is apparent that a ∼ 0.4 reproduces the his-
togram quite well. The corresponding value of β is 0.9, which
is perfectly consistent with the allowed range of β, -0.5 to
2 (see eq. [3]). Thus, a ∼ 0.4 (β ∼ 0.9) reproduces both the
θim -- zs anticorrelation, and the observed frequency of wide
separation lenses. In this scenario, the θim ∼ 7′′ lenses are
due to ∼ 2L∗ galaxies, with a mass within 10h−1kpc of
3.7 1011h−1M⊙.
The major deviation of the present model from those
found in the literature is in the value of a, the power law
index relating the dark matter characteristic scale length of
a galaxy to its luminosity. As mentioned earlier, a is usually
assumed to be 1.2, while consistency with lensing observa-
tions in the framework of the present model implies a ∼ 0.4.
5 CONCLUSIONS
If the observed distribution of image separation vs. source
redshift of lensed QSOs is real and not a result of false
lenses, like physical QSO pairs, then it has the following im-
plications for galaxy dark matter halos: (i) galaxies must
have changing logarithmic surface mass density profiles,
with a 'universal dark matter' model being preferred over an
isothermal sphere with a core, (ii) there must be a spread in
galaxy properties, like the central surface mass density, (iii)
the characteristic length scale of dark matter halos should
scale with luminosity as La, where a ∼ 0.4.
The corollary of the present model is that higher zs mul-
tiply imaged QSOs are statistically more magnified, and are
preferentially lensed by galaxies with lower central surface
mass densities and intrinsic luminosities. These predictions
can be tested if each lens case is modelled individually, which
will become possible when the lensing galaxies of most mul-
tiply split QSOs are detected.
The strong dependence of the observed θim -- zs relation
on galaxy properties means that it cannot be easily used as
a test for the curvature of the Universe unless we know a lot
more about the mass distribution of galaxies. Furthermore,
c(cid:13) 1997 RAS, MNRAS 000, 000 -- 000
Image Separation of Lensed QSOs
5
Figure 3. Predicted image separation distribution for lensed
QSOs. The curves are labeled by parameter a, defined by rc ∝ La.
Vertical normalization is arbitrary. The predictions are valid for
zs where the magnification bias effect, discussed in Section 2.2
and shown in Figure 2, does not impose an upper limit on θim of
high-zs lensed QSOs, i.e. for zs <∼ 2.5. Accordingly, the observed
distribution to which the curves apply, is restricted to zs < 2.5
and is shown as the solid histogram. Values of a ∼ 0.4 seem to
describe the observed distribution adequately. In these models
image separation is a monotonically increasing function of galaxy
luminosity; image separation produced by an L∗ galaxy is indi-
cated by an arrow.
because of the implications the present model has on the
lensing properties of galaxy population, like the dependence
of the predicted frequency of lensed QSOs on parameter a
(see Figure 3), the current upper limits on Λ may have to
be revised.
ACKNOWLEDGMENTS
I am grateful to Joachim Wambsganss for useful suggestions.
I would like to acknowledge the support of PPARC fellow-
ship at the Institute of Astronomy, Cambridge, UK.
REFERENCES
Bartelmann, M. 1996, A&A, 313, 697
Boyle, B. J., Shanks, T., & Peterson, B. A. 1990, MNRAS, 243, 1
Evans, N. W., & Collett, J. L. 1997, Preprint, also available as
astro-ph/9702085
Fukugita, M., Futamase, T., Kasai, M., & Turner, E. L. 1992,
ApJ, 393, 3
Gott, R. J. III, Park, M.-G., & Lee, H. M. 1989, ApJ, 338, 1
Hewett, P. C., Foltz, C. B., & Chaffee, F. H. 1993, ApJ 406, L43
Kochanek, C. S., 1996. ApJ, 466, 638
Kormendy, J. & Djorgovski, G. 1989, ARA&A, 27, 235
Lauer, T. R. 1985, ApJ, 292, 104
Lilly, S. J., Tresse, L., Hammer, F., Crampton, D., & Le Fevre,
O. 1995, ApJ, 455, 108
6
L. L. R. Williams
Park, M.-G., & Gott, R. J. III. 1997, Preprint, also available as
astro-ph/9702173 (PG97)
Maoz, D., & Rix, H.-W. 1993, ApJ, 416, 425 (MR93)
Myers, S. T., et al. 1995, ApJ, 447, L5
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
Schechter, P. L., 1976. ApJ, 203, 297
Schneider, P., Ehlers J. and Falco E. E. , 1992, Gravitational
Lenses, Springer-Verlag Press
Turner, E. L., 1990. ApJ, 365, L43
Wallington, S., & Narayan, R. 1993, ApJ, 403, 517
Williams, L. L. R., & Lewis, G. F. 1997, MNRAS, submitted
c(cid:13) 1997 RAS, MNRAS 000, 000 -- 000
|
astro-ph/9910336 | 1 | 9910 | 1999-10-19T05:32:43 | Conference highlights of "Toward a New Millennium in Galaxy Morphology" | [
"astro-ph"
] | A summary of the highlights of the conference, "Toward a New Millennium in Galaxy Morphology" is presented. In this review, I cover the major topics addressed at the conference, including both the observational and theoretical frameworks suggested for understanding the morphology of galaxies. | astro-ph | astro-ph |
Conference Highlights
Toward a New Millennium in Galaxy Morphology1
The most basic process in any observational science is taxonomy - the classification of objects by their
natural relationships. For almost a century, the Hubble classification has been the dominate morphological
paradigm, but it appears new ideas and systems are needed. Many galaxies at high redshift as well as a
fraction of galaxies in the nearby universe are unclassifiable in the Hubble sequence. Some nearby examples
are low surface brightness galaxies, including dwarf spheroidals, and galaxies in clusters, especially large cD
galaxies. With so much complexity in the galaxy population, new systems and ideas are needed.
We can ask at the turn of the century if optical images contain enough information to adequately
classify galaxies. Perhaps other wavelengths, such as the infrared, UV, radio and sub-mm are also needed
to fully cover the basic features of extragalactic systems. Spectroscopy of a galaxy, including kinematic
information, or HI line profiles might also be better morphological indicators than a galaxy's image.
To address these issues, over 90 astronomers clustered in Johannesburg, South Africa between 13-18
September to discuss the future of galaxy morphology, and how classification relates to galaxy evolution
and formation. Over 60 talks and a number of posters were presented during the meeting.
Several new methods of classifying galaxy images were proposed at the conference. These new methods
are in many ways different from the qualitative Hubble system. Not only are they quantitative in nature,
with parameters such as asymmetry, concentration of light, Fourier components, measures of bar strengths,
and automatic neural networks, but most do not reproduce the Hubble sequence, but attempt to modify or
replace it (Abraham, Bershady, Conselice, Frei, Takamiya, Windhorst). Physical morphological systems
based on comparing the form of a galaxy with its spectral-type look promising; these are similar to the
ideas proposed by W.W. Morgan in the 1950s. There is still however, no all encompassing system that can
relate all galaxies in different environments.
High redshift galaxies, such as those seen in the Hubble Deep Field (HDF) have brought about this
new interest in morphology. Although, most of the HDF, and medium deep survey (MDS) images of distant
galaxies are sampling rest frame ultraviolet morphologies, the most disturbed galaxies remain disturbed
when viewed with NICMOS on HST. Therefore, these galaxies are intrinsically different from nearby ones;
their morphologies are not the result of morphological k-corrections. We are however, only sampling the
young to middle aged stars when viewing galaxies in optical/near infrared wavelengths. Limiting ourselves
to these wavelengths, we are missing a large fraction of the light output from galaxies. The effects of dust
on the morphology of galaxies, and much of the gas is being missed.
Our view of the dusty universe is provided by ISO, IRAS, SCUBA and other instruments that probe
long-ward of 5µm (Cesarsky, Laurent, Mirabel, Sanders, Sauvage, Scoville). Not only does a significant
fraction of the light emitted from galaxies occur at these wavelengths, but because of dust effects several
features can only be seen by examining galaxies at these wavelengths. Hidden star formation in mid
ISOCAM images of nearby galaxies, prove that dust is a significant factor affecting the morphology of
galaxies. Luminous and ultra-luminous infrared galaxies are probably major galaxy mergers, with their
1Conference was held in Johannesburg, South Africa from 13-18 September 1999. Proceedings will be edited by D.L Block,
I. Puerari, A. Stockton and D. Ferreira and published in a special edition of Ap&SS
-- 2 --
optical morphology dominated by dust absorption. SCUBA sources seen at redshifts z < 6, are probably
analogs to nearby ULIGs and possibly are progenitors of ellipticals with L > 1012 L⊙. Many of these
questions will be answered with the next generation of sub-mm instrumentation (Combes).
Using near infrared images, particularly in the K band is useful for determining the underlying
structure of the old stellar populations in galaxies. Bars and other morphological features are more easily
detected in the NIR. The spiral structure of galaxies, specifically the pitch angle, changes between optical
and NIR wavelengths (Block). Using Fourier analysis, three principle NIR spiral galaxy morphologies are
found. The pitch angle of the spiral arms in these families, viewed in the NIR correlates with the shape of
velocity rotation profiles (Block), an effect not seen in optical wavelengths.
Using dust corrected rest frame optical fluxes, infrared, sub-mm fluxes from high redshift galaxies and
the sub-mm background, it is possible to obtain an estimate of the star formation history of the universe
(Lagache, Thompson). Determining the star formation rate in a galaxy is tricky, and estimates of star
formation rates are based on dust correction assumptions. Although, we do not know if the dust in high
redshift galaxies behaves the same as dust in nearby galaxies. To paraphrase Mayo Greenberg, "Galaxies
evolve, stars evolve, so why not dust?". Additionally, we do not know the full effects of stellar winds, SNe,
magnetic fields, and possibly other feedback mechanisms (Chu) to say nothing of the effects of black holes
and AGN phenomenon on the morphology of galaxies.
The masses of black holes correlate with the masses of their host galaxy's spheroidal components
(Macchetteo) suggesting a possible feedback mechanism and related formation. AGNs might also play an
important role in star formation in galaxies. The space density of AGNs is related to the star formation
history of the universe (Miley). Jet outflows are an important component for the formation of stars in
galaxies, and hence could have a significant effect on their morphologies.
These physical mechanisms can change how a galaxy looks in the general, but perhaps more detailed
features deserve a closer look. Rings, bars, and globular clusters all hold clues to galaxy formation. Rings
(Buta) can be located at resonances, are sites of star formation and are potentially triggered by bars, or
interactions with other galaxies. Bars also produce spiral arms, star formation, and drive secular evolution
in spiral galaxies. The bar frequency among spiral galaxies in the local population seen in the near infrared
is at about 60% (Eskridge, Knapen). The bar fraction decreases at higher redshifts (Abraham), suggesting
that bars might form later in the evolution of galaxies. The specific frequency and properties of globular
clusters, especially metallicity distributions, can also be used to determine the evolutionary history of
galaxies (B. Elmegreen).
An overlooked feature for morphological studies of galaxies is their gas content. While some forms of
ionized gas can be seen at optical wavelengths, most gas is detectable in the radio. HI 21cm lines images of
galaxies are generally more extended than in the optical, and more commonly show signs of interaction and
mergers than galaxies in optical light (Sancisi). Interactions between galaxies is a critical aspect for their
evolution, and any complete morphological system must include dynamical indices. The effects of mergers
and interactions on galaxies, including the intensity of the induced star formation, can vary depending on
the initial orbital conditions (Horellou).
HI gas dynamics can be used to determine the underlying physical nature of galaxy disks, including
dark matter halos. Recent observations suggest that dark matter halos are triaxial and possibly rotating in
some galaxies (Freeman). These rotating dark matter halos, predicted by CDM could be responsible for
the structure of HI disks. High resolution rotation curves also show that the luminous mass of galaxies
does not trace dark matter potentials in halos (Sofue). Vertical structures of edge-on disk galaxies are also
-- 3 --
inconsistent with slowly rising rotation curves (van der Kruit).
Knowing the structure of galaxies naturally leads towards learning how galaxies formed and evolved.
Observationally we know that the galaxy population looked different in the past (Abraham, Conselice,
Illingworth, Windhorst) but how did this evolution occur? Evidence for dissipative collapse, where galaxies
are created from the inside out, includes metallicity gradients that correlate with bulge luminosity, and
old bulges (Goudfrooij). On the other hand, properties of spirals from Sm to S0s suggest galaxies formed
through a secular process (Pfenniger). Additionally dynamical evolution of galaxies based on semi-analytical
and N-body simulates using likely initial conditions of the universe suggest galaxies formed hierarchically.
Although, dark matter halos in these models are too concentrated and fail to produce correct sizes for disk
galaxies (Steinmetz). There is however, observational evidence for hierarchical formation of galaxies at
high redshifts (Illingworth). Morphologies of certain spirals can also change in special environments such as
galaxy clusters through high speed galaxy interactions (Lake).
The observations and theory described above is beginning to reveal how galaxies were formed, evolved
and how fundamental properties relate to morphology. A good taxonomy system for extragalactic objects
should classify galaxies according to characteristics that best describe galaxy evolution; perhaps interactions,
star formation history, gas content, and dark matter/rotation curve profiles, are among these. Although
such a morphological system does not presently exist, the accumulation of observational data and the
further development of theory will suggest what parameters are the best for understanding the structure
and evolution of galaxies.
Christopher J. Conselice
University of Wisconsin, Madison
|
0804.2074 | 1 | 0804 | 2008-04-13T16:44:06 | Bohdan's Impact on Our Understanding of Gamma-ray Bursts | [
"astro-ph"
] | Bohdan Paczy'nski was one of the pioneers of the cosmological GRB model. His ideas on how GRBs operate and what are their progenitors have dominated the field of GRBs in the hectic nineties during which the distances and the origin of GRBs were revealed. I discuss here Bohdan's contributions in some historical perspective. | astro-ph | astro-ph |
**FULL TITLE**
ASP Conference Series, Vol. **VOLUME**, **YEAR OF PUBLICATION**
**NAMES OF EDITORS**
Bohdan's Impact on Our Understanding of Gamma-ray
Bursts
Tsvi Piran
Racah Institute for Physics, The Hebrew University, Jerusalem, Israel
91904
Abstract.
Bohdan Paczy´nski was one of the pioneers of the cosmological
GRB model. His ideas on how GRBs operate and what are their progenitors have
dominated the field of GRBs in the hectic nineties during which the distances
and the origin of GRBs were revealed. I discuss here Bohdan's contributions in
some historical perspective.
1. Prologue
I first met Bohdan in the summer of 1977 at the IOA. I don't remember the
exact date, but I remember the time. It was between 3am and 4am. Several
of us were trying to open the last remaining bottle of wine at the traditional
IOA summer (Argentinean) barbecue. Something did not work out. Bohdan
approached us and succinctly pointed out that the process will be much easier
if we remove first the plastic warp that covered the cork. We did. It worked.
This was a simple demonstration of Bohdan's ability to see the obvious simple
fact that everyone else around stared at - but no one sees.
Our next meeting took place several years later at Princeton. Bohdan and
his family moved to the Institute housing, where I lived as well. I have spent
many pleasant evenings visiting them. We enjoyed many discussions. When we
did not discuss life in America versus life in Europe those conversations turned
naturally to Gamma-Ray Bursts (GRBs) - a subject that fascinated both of us.
2.
Introduction - GRBs as we understand them today
I begin with a brief description of the present picture of GRBs and their mod-
els. This section is not intended to be a comprehensive review (see Piran 1999;
M´esz´aros 2002; Piran 2005; M´esz´aros 2006; Woosley & Bloom 2006; Zhang,
2007 for recent reviews and Fishman & Meegan, 1995; Piran 1995 for a histor-
ical perspective). My objective is to outline our current understanding so that
Bohdan's contributions and impact can be put into the right perspective and be
appreciated.
GRBs are short and intense (∼ 10−7 −10−5ergs/cm2 ) bursts of soft (10 keV
- 2 MeV) gamma rays coming from random directions in the sky. GRBs were
discovered accidentally in the late sixties by the Vela satellites - defense satellites
that were sent to monitor the outer space treaty. The discovery was reported
only in 1973 (Klebesadel et al.). In the early nineties the BATSE detector on-
board of the COMPTON-GRO satellite indicated that GRBs are cosmological
1
2
Tsvi Piran
(Meegan 1992, Mao & Paczy´nski 1992, Piran 1992). Towards the late nineties
the BeppoSAX satellite discovered X-ray afterglows (Costa et al., 1997). Op-
tical (van Paradijs et al., 1997) and Radio (Frail and et al., 1997) afterglows
followed shortly. Direct redshift measurement of the optical afterglow of GRB
970508 (Matzger et al., 1997) established the cosmological origin of GRBs. Ra-
dio scintillation (Goodman, 1997) and later VLBI observations (Taylor et al.,
2004) confirmed the relativistic bulk motion, predicted earlier by the fireball
model.
In the following years afterglow observations revealed a wealth of in-
formation on host galaxies (Djorgovski et al. 2003) and the positions of GRBs
within host galaxies, which was always within star forming regions (Paczy´nski
1998, Fruchter, 2006).
GRB belong to two groups (Kouveliotou et al. 1993) according to their
durations T90: Long GRBs (with T90 > 2sec) and short (with T90 < 2). Short
bursts are also typically harder and hence they are called at times short hard
bursts (SHBs). The association of several long GRBs with supernovae (Galama
et al. 1998; Bloom et al. 1999; Stanek et al. 2003; Hjorth, et al. 2003) confirmed
the Collapsar model (Woosley 1993; Paczy´nski 1998; McFadyen & Woosley,
1998) according to which (long duration) GRBs arise during the collapse of
massive stars. The origin of short GRBs is less certain. Unlike long GRBs, short
ones do not arise in star forming regions. They arise in both elliptical and spiral
galaxies (see e.g. Nakar 2007 for a review). The redshift distributions shows that
the observed short bursts are much nearer than the observed long ones (Guetta
& Piran, 2006; Nakar et al., 2006). These observations are consistent with
earlier suggestions that short GRBs arise from neutron star mergers (Paczy´nski,
1986; Eichler et al., 1999). Within this model the delay corresponds to the
gravitational radiation spiral in time. The delay is crucial also in positioning
the progenitors away from their birth places, possible even outside the original
galaxy.
The internal-external shocks model (also called the fireball model) (Good-
man 1986; Paczy´nski, 1986; Shemi & Piran, 1990; M´esz´aros & Rees 1992, Rees &
M´esz´aros, 1994, Narayan, Paczy´nski & Piran, 1992; Sari & Piran 1997, Wijers,
Rees & M´esz´aros, 1997) is depicted in Fig. 1. It suggests that GRBs involve
ultrarelativisitic (baryonic or Poynting flux) jets that emerge from compact ob-
jects, most likely black holes. Internal collisionless shocks or instabilities within
the relativistic outflow, that take place at 1014 − 1016cm from the central engine,
accelerate particles to relativistic velocities and generate strong magnetic field.
Synchrotron or SSC emission of these particles produce the prompt emission.
The outflow interacts later with the surrounding medium producing a relativis-
tic blast wave - a relativistic analogue of a Supernova remnant. This long lasting
blast wave accelerates particles and generates magnetic fields that produce the
multiwavelength afterglow. The predictions (Sari, Narayan & Piran, 1998; Sari,
Halpern & Piran, 1999) of this "standard model" agreed nicely the observed
light curves and spectra of GRB afterglows (Wijers & Galama, 1999; Kulkarni
et al., 1999; Panaitescu & Kumar 2002; Yost S.A., et al., 2003). Recent Swift
observations indicate that the picture is somewhat more complicated than what
is described here. Still for the purpose of this review this description is sufficient.
Bohdan's GRBs
3
Figure 1.
The standard internal-external shocks model of GRBs (see e.g.
Piran 1999; M´esz´aros 2002; Zhang and M´esz´aros 2004; Piran 2005; M´esz´aros
2006 for reviews). 1. Central engine and acceleration of the flow. 2. Coasting
phase; terminal Lorentz factor Γ > 100. 3. Internal shocks; prompt gamma-
ray emission . 4. The interaction of the outflow with the surrounding matter
produces the external shocks system with a blast wave propagating outwards
producing the afterglow.
3. Prehistory - GRBs before BATSE
The history of GRBs should be divided to two period. Pre and post BATSE. At
the pre-BATSE era (up to the late eighties and very early nineties) a consensus
formed that GRBs originate in magnetic instabilities on galactic neutron stars.
What was considered as a strong evidence for this model was the detection of
20 and 40 keV lines which were interpreted as cyclotron lines of a ∼ 1012 Gauss
field - a magnetic field expected in neutron stars.
Bohdan was skeptical. He never believed the observed 20 and 40 keV lines
e.g.(see Narayan, Paczy´nski & Piran, 1992) - insisting that there was never the
case that the same line was observed simultaneously from the same burst by
two different detectors. With no lines there was no compelling case for galactic
neutron stars and the road to cosmological sources was open.
If the bursts
are cosmological they release much more energy ∼ 1051ergs or more. A basic
problem with any cosmological model was that GRBs have non-thermal spectra
indicating that they are optically thin. However, the temporal variability (less
4
Tsvi Piran
than a fraction of a second) indicated that the emission region is small (less than
109cm). If 1051ergs of soft gamma-rays (with numerous photons above 500 keV)
are released from such a source there would be copious pair production that will
lead to an optically thick source. This is the so called "Compactness problem".
The "Compactness problem". Left:
Figure 2.
light curves for different
bursts. Variability on very short time scales is clearly seen. Right: The spec-
trum of GRB 910601. The nonthermal spectrum is clearly seen. If 1051ergs
of radiation with many photons above mec2 (as seen in this spectrum) are
released within a small region (as indicated by the temporal variability) nu-
merous pairs will form and the radiation-pairs plasma will be optically thin.
The breakthrough concerning the resolution of the "Compactness problem"
came in 1986 when two back to back papers where publisehd in the Astrophysical
Journal Letters by Jeremy Goodman (1986) and Bohdan Paczy´nski (1986). Both
papers describe an idealized problem in which a hot radiation, a fireball, with
T0 ≫ mec2 is released within a small region. They argues that the radiation will
form a radiation-pair plasma fluid that will expand rapidly. As the fireball
accelerates it cools (in its local frame). When the local temperature drops
below mec2 the pair begin to annihilate. There are enough pairs to keep the
fireball optically thick until the local temperature drops to ∼ 20keV when the
fireball become optically thin. By this time the fireball has reached a relativistic
motion with Γ ∼ T0/(20keV). The photons escape freely now. Because of the
relativistic motion the escaping photons will be seen by an observer at rest with
Bohdan's GRBs
5
approximately the original temperature T0. Jeremy solved the problem in the
impulse approximation, considering the case of an "explosion". Bohdan, on
the other hand, considered a steady state approximation in which the evolution
was approximated as a steady wind. The results were more or less similar.
The papers were the first to suggest that GRBs involve relativistic motion. A
concept that was verified more than ten years later first using radio scintillations
(Goodman, 1997) and later with direct VLBI measurements (Taylor et al., 2004).
While the two papers are closely related, they are not similar. Goodman focuses
on the physics and is very concerned with the fact that at the end the resulting
photons have a thermal like spectrum. Bohdan, on the other hand, is not so
worried about this issue and he uses this model to put forwards his idea that
GRBs are in fact cosmological. He even goes on to suggest, in the discussion
section of his work, that some reoccurring sources arise because of gravitational
lensing of cosmological events.
At that time we spent many hours discussing what could drive such events.
Our discussions focused on neutron star mergers. Objects that at the time
became popular among relativists as it was realized that they are prime candi-
dates for detection as sources of gravitational radiation. Our discussions evolved
around the final stages in which tidal interaction between the two stars will tear
them apart, and Bohdan borrowed here a lot from his work on binaries. Still
Bohdan was reluctant to publish a paper with a model describing mergers as
GRB sources. My guess was that what he really cared most at the time was
to "win the cosmological war" on the distances to the bursts. He was proba-
bly worried that linking the cosmological idea to the very speculative merger
model (Neutron star mergers were considered as somewhat esoteric events in
those days), might weaken his cosmological case. Bohdan speculated on this
merger possibility very briefly in two or three sentences in his 1986 paper and
returned to it (actually to a variant - black hole neutron star merger (Paczy´nski,
1991)) only in the early nineties. I went on and published these ideas, with a
detailed model on how a merger can produce a GRB, in 1989 (Eichler, Livio,
Piran & Schramm 1989). We wrote our joint paper on this subject only in 1992
(Narayan, Paczy´nski & Piran, 1992).
I will return to this paper later in this
talk.
4. BATSE, the Cosmological-Galactic War and the Great Debate
The Burst and Transient Source Experiment, BATSE, on board of the Compton-
GRO satellite, was expected to demonstrate that GRBs are galactic. BATSE's
spectrograph was supposed to detect the lines and its ability to localize the
bursts within a few degrees was sufficient to demonstrate that the bursts are
within the galactic disk. It is difficult to imagine the shock when BATSE's first
results were announced in the fall of 1991: No lines, Isotropic distribution
with no dipole moment (corresponding to our position relative to the galactic
center), and last but not least a paucity of weak bursts (Meegan et al., 1992).
The results were accepted with disbelief and at first were sort of ignored. Most
of the talks, at least most of the theoretical talks, given at the first Huntsville
symposium, that took place in October 1991, dealt with various aspects of the
galactic neutron star model. A whole session was devoted to magnetospheric
6
Tsvi Piran
instabilities on neutron stars. One speaker, J. P. Lasota, withdrew his talk on
a galactic model and stated that it became irrelevant in view of the new data.
Bohdan (Paczy´nski, 1992) and me (Piran, Narayan & Shemi 1992) were the only
speakers that dealt with cosmological models in that meeting.
Both Bohdan (Mau & Paczy´nski, 1992) and me (Piran 1992) rushed to
publish letters on the essential cosmological interpretation of BATSE's result.
We stressed that the combination of an isotropic distribution, that at that stage
was demonstrated by upper limits on the dipole moment of the distribution (see
e.g. Fishman et al.1994; Hakkila et al.1994) and the paucity of weak bursts,
which at that stage was not demonstrated by a full log N − log S distribution
(see Fig. 3) but by the very low hV /Vmaxi = 0.32 value (see e.g Fishman et
al.1994; Hakkila et al.1994) forced a cosmological distribution. The argument
was surprisingly simple.
If the sources are galactic their angular distribution
must show a significant dipole moment, unless (i) Their radial distribution is
extremely local (and we don't see the galactic structure) or (ii) The radial dis-
tances are very large so that the typical distance compared to our distance from
the galactic center is large. Possibility (i) is ruled out by the hV /Vmaxi value.
A small industry begum by proponents of the galactic models that threw GRB
sources to larger and larger distances in the galactic halo (e.g. Podsiadlowski
et al.1995). As the limits on the dipole moment become tighter and tighter
the average distance needed for the sources continuously increased. I vividly re-
member Bohdan joking that with a typical distance of 200kpc one might call the
model cosmological... Additionally, as the distances increased so did the energy
budget until even the galactic models suffered from the Compactness problem
(Shemi & Piran, 1993).
Bohdan was very excited during this period. He enjoyed the lively (at
times too lively) debate that went on. He told me that in order to prevent
misunderstanding or misrepresentations of his ideas he was signing, at that time,
all GRB referee reports that we wrote. He was most amused when Nature
published a galactic origin paper (Lingenfelter & Higdon 1992) and refused to
publish his rebuttal because "our readers will be confused if we published two
contradicting opinions". Both of us had a great time making fun of some of
the models. Surprisingly the fight went on for quite some time. At the second
Huntsville meeting that took place at the fall of 1993 more than half of the
audience still believed in galactic origin and this situation persisted even in the
"Great Debate" that took place in 1995.
The debate culminated in the "Great Debate" that too place in April 1995
at the Smithsonian in Washington, D.C., between Bohdan (Paczy´nski, 1995) and
Don Lamb (Lamb 1995). Bohdan's abstract presents his argument best: "The
positions of over 1000 gamma-ray bursts detected with the BATSE experiment
on board of the Compton Gamma Ray Observatory are uniformly and randomly
distributed in the sky, with no significant concentration to the galactic plane or
to the galactic center. The strong gamma-ray bursts have an intensity distri-
bution consistent with a number density independent of distance in Euclidean
space. Weak gamma-ray bursts are relatively rare, indicating that either their
number density is reduced at large distances or that the space in which they are
distributed is non-Euclidean. In other words, we appear to be at the center of
a spherical and bounded distribution of bursters. This is consistent with the
Bohdan's GRBs
7
Figure 3.
Right: The distribution of BATSE bursts on the sky, in galactic
coordinates. The isotropy is self evident now. Left: the log(N ) − log(S) plot.
The number of bursts with N with a peak flux larger than S. In a homogenous
distribution we should see a straight line with a slope of −3/2.
distribution of all objects that are known to be at cosmological distances (like
galaxies and quasars), but inconsistent with the distribution of any objects which
are known to be in our galaxy (like stars and globular clusters). If the bursters
are at cosmological distances then the weakest bursts should be redshifted, i.e.,
on average their durations should be longer and their spectra should be softer
than the corresponding quantities for the strong bursts. There is some evidence
for both effects in the BATSE data. At this time, the cosmological distance scale
is strongly favored over the galactic one, but is not proven. A definite proof (or
dis-proof ) could be provided with the results of a search for very weak bursts in
the Andromeda galaxy (M31) with an instrument 10 times more sensitive than
BATSE. If the bursters are indeed at cosmological distances then they are the
most luminous sources of electromagnetic radiation known in the universe. At
this time we have no clue as to their nature, even though well over a hundred
suggestions were published in the scientific journals. An experiment providing
1 arc second positions would greatly improve the likelihood that counterparts of
gamma-ray bursters are finally found. A new interplanetary network would offer
the best opportunity."
8
Tsvi Piran
It was the 75th anniversary of the Curtis and Shapley 'Great Debate' (where
the size of the Universe was contested just three years before Edwin Hubble made
his seminal discovery about its expansion and thus the size of the Universe). The
parallel of the Curtis/Shapley and GRB debates could not be more poignant:
just as few changed their minds about the size of the Universe after the 1920
debate as have changed their mind about the distances of GRBs after this one.
The audience was split roughly 50% 50% at the end of this discussion. I must
say that when I heard the results of the vote at this debate I was shocked. To
me at that time the question what are the distances of GRBs was not an open
question (Piran 1995).
Bohdan was very happy about this work. In a report of his achievements in
a NASA grant that he posts (in a very unusual manner) on astro-ph (Paczy´nski,
1996) he writes: "The research project: 'Models and Scenarios for Gamma-Ray
Bursts' resulted in a total of 20 published research papers. The central issue was
the distance scale to gamma-ray bursters, the issue brought up by the remarkable
discovery of the distribution properties of gamma-ray bursts by the BATSE on
Compton GRO. The last paper on the reference list is the PI's contribution to
the debate on the distance scale to gamma-ray bursts held in Washington DC on
22 Apr. 22 1995. When this project got started, right after the announcement
of the BATSE results at the conference in Annapolis in the fall of 1991, only a
small fraction of astrophysicists seriously considered the possibility that gamma-
ray bursts are at cosmological distances. By now the cosmological distance scale
has become a majority view, to a large extent because of the publications listed
in this final report. PI considers this to be the most important and lasting result
from the research supported by this grant."
The debate ended with the measurement of the redshift of the optical af-
terglow of GRB 970508 (Metzger et al., 1997). It is amusing to recall that even
after this observation "conspiracy theories" suggesting coincidence between this
GRB and the observed afterglow was put forwards for a while.
5. Building a theoretical model - how GRBs work
In 1986 Bohdan contributed to the early ideas on a thermal fireball. However,
it was clear that this idea is rather preliminary. The resulting emission from
this simple fireball has a quasi-thermal spectrum, quite unlike what is observed.
Additionally,it was clear that this picture of a pure radiation (and pairs) fireball
was too idealized.
5.1. From Radiation and Baryons to Ultra relativistic Baryonic out-
flow
The next question was what would be the effects of baryons on such a pure
radiation fireball. Here, once more we have worked in parallel on the same prob-
lem and the results were published more or less simultaneously (Abramowicz,
Novikov & Paczy´nski, 1990; Shemi & Piran, 1990). Both works have shown that
the addition of even a small baryonic load would change drastically the outcome.
The baryons will be dragged along and accelerated by the radiation field. If the
total baryonic load is not too small eventually all the initial thermal energy will
be transferred to the kinetic energy of the baryons. If the baryonic load is not
Bohdan's GRBs
9
too large, that is m < E0/c2 the final outcome will be a relativistic baryonic
outflow with Γ = E0/M (E0 being the total energy and M the baryonic mass).
This was a step in the right direction but it was not good enough. Relativistic
baryons can serve as a source of cosmic rays. But what we have to find a way
to convert their kinetic bulk energy back to radiation.
5.2. Astro-ph/9204001
In the spring of 1992 Bohdan gave a colloquium, on the cosmological origin
of GRBs at Harvard.
I was amused by the fact that talk was on March 5th
- the anniversary of the famous 1979 March 5th GRB. This was was the only
well localized GRB at that time, and it was in the LMC - giving credence to
the Galactic origin. By now we know that it was not a regular GRB but a
soft Gamma repeater (SGR) but this is another story. Bohdan stayed for a
whole week during which we wrote, together with Ramesh Narayan, "Gamma-
Ray Bursts as the death throes of massive binary stars" (Narayan, Paczy´nski &
Piran, 1992- NPP92).
This paper, astro-ph/9204001, is the first astro-ph.
It is a well known,
highly cited(more than 400 citations) paper. Most remember it because of the
serious attempt to build a consistent merger model for GRBs. However, in
addition to discussing neutron star mergers that it contained several new ideas
that helped shape our theoretical understanding of GRBs:
• Internal shocks
• GRBs are powered by accretion onto a newborn black holes.
• The duration of the GRB is determined by the activity of the inner en-
gine (accretion onto the black hole) while the duration of the pulses is
determined by the dynamical time scale of the source
• GRBs should be followed by a long lasting Afterglow
• Binary neutron stars has a long spiraling-in phase before they merge. As
they receive a large kick when the binary is born they can escape during
this phase from their host galaxy.
• 1015Gauss magnetic fields may exist at GRBs' inner engines.
Somewhat surprising, or maybe not, the paper was not easily accepted. The
first referee simply rejected it as (i) too speculative and (ii) irrelevant because
GRBs are galactic. The second one accepted it reluctantly stating: "It three
well respected well known scientists what to make fools of themselves who am I
to stop them. I was very happy to be referred to as a "respectable and well known
scientist" but I suspected that I gained this title because of my co-authors. I
was even happier that the paper was accepted. In retrospect given the "anti-
cosmological" atmosphere at that time we should have been grateful to have had
such a broad minded referee.
10
Tsvi Piran
5.3. From Matter to Light Internal Shocks
The most urgent question was how to convert the relativistic baryonic outflow
to radiation. The solution proposed in NPP92 was internal shocks: "These
ejecta should, through collisions with one another at large radii and low optical
depth, give a non Planckian spectrum by various nonthermal mechanisms. For
instance given the strong magnetic fields, synchrotron processes might naturally
produce the observed power-law spectrum.". This was of course just a short
exposition of the idea and Bohdan continues to elaborate on it later in 1993
together with Ghuohong Xu (Paczy´nski & Xu 1993). The basic idea is that
the source is irregular and it emits a wind with a variable Lorentz factor. An
idealized picture of such a wind is a situation when the outflow is in the form
of shells that moves with different Lorentz factors. Faster shells collide with
slower ones. Paczy´nski & Xu (1993) consider within this context proton proton
collisions that produce pions that decay producing Gamma-Rays. However this
won't be very efficient. Later on Rees & M´esz´aros (1994)suggested the picture
accepted now according to which the collisionless shocks that form accelerate
particles to very high energies and possibly also generate strong magnetic fields.
These shocks are the source of the observed prompt emission. The schematic
picture is depicted as a central part of Fig. 1 and in Fig. 4. It is central now
to our current model. However at the time the simpler external shocks picture
proposed a few years earlier by M´esz´aros and Rees (1992) - namely interaction
of the outflow with the surrounding matter- was considered as the source of
the prompt emission. Only in 1997 we (Sari & Piran, 1997) have shown that
external shocks cannot produced the observed variable light curves and as of
today internal shocks are the only known viable way to convert the kinetic
energy of the outflow to radiation and produce a highly variable light curve!
It should be stressed that the same ideas hold for Poynting flux dominated
outflow, a possibility that is considered seriously now as an alternative to the
baryonic outflow. The kinematic arguments that require internal process within
the outflow are applicable regardless of the exact nature of the outflow. The
main difference of course is that in the Poynting flux case instabilities such as
reconnection replace the collisionless shocks. But the basic feature that external
interaction cannot produce the variable light curves remains.
The idea of internal shocks is linked directly to another important issue.
In internal shocks the temporal structure of the burst is determined by the
inner engine. The duration is determined by the time that the source is active
while the dynamical time scale of the source dictates the short term fluctuations
time scale seen in the individual pulses. The need of a prolonged activity of
the inner has pointed out towards accretion onto a newborn black hole as the
maincandidate for the activity
5.4. From Matter to Light Afterglow
Clearly, not all the kinetic energy can be dissipated in internal shocks. Only
the relative kinetic energy can be dissipated there. The bulk center of mass
kinetic will remain. This leads immediately to the idea of an afterglow: "The
ejecta should much later also produce something similar to a supernova rem-
nant." (NPP92). With James Rhodas, (Paczy´nski & Rhodas, 1993) Bohdan
continues to elaborate on this idea. They use the analogy with supernova to es-
Bohdan's GRBs
11
Figure 4.
Internal shocks are produced from a variably flow (schematically
depicted as shells). The faster shells catch up with the slower ones and produce
the shocks. The system is, most likely, powered by an accretion disk and 1015
Gauss field could arise in this disk.
timate that GRBs will be followed by a radio transients. More detailed models
focusing on higher energies were suggested later by numerous authors, (M´esz´aros
& Rees, 1997; Wijers, Rees & M´esz´aros, 1997; Waxman, 1997; Sari, Piran &
Narayan, 1998). Still this was the first suggestion that GRBs will be followed
by a long duration afterglow.
The afterglow prediction was verified in 1997 with the discovery of X-ray
afterglow by BeppoSAX (Costa et al., 1997). This discovery was followed by
detection of optical (van Paradijs et al., 1997) and radio (Frail et al., 1997)
afterglows as well.
6. GRB progenitors - mergers and hypernovae
The question what makes GRBs is of course the central one for the whole prob-
lem. Once it was realized that the bursts are cosmological and the energy budget
was set to be around 1051 − 1052ergs it was clear that the sources involve a com-
pact object. Most likely the formation of the object and the release of its binding
energy, and if not that at least a significant catastrophic event. The neutron
star merger model, which is currently the most likely to work for short GRBs,
12
Tsvi Piran
was proposed already in the eighties (Paczy´nski, 1986; Eichler et al., 1999). For
a while it was thought that this is the model for all GRBs.
However, when BeppoSAX begun detecting X-ray afterglows and host galax-
ies were revealed, it turned out that GRBs (actually long GRBs, as BeppoSAX
detected afterglow only from long GRBs) are located within small irregular star
forming galaxies. Now we know that within these galaxies GRBs are located in
the highest star forming regions and that the rate of GRBs is roughly propor-
tional to the square of the rate of star formation (Fruchter et al., 2006). Bohdan
did not need much data to make a far reaching conclusion. At the fall of 1997
he concluded on the basis of just three well localized GRBs, 970228, 0970508
and 980828 that GRBs arise in the vicinity of star forming regions (Paczy´nski,
1998). He concludes that the merger model must be abandoned and GRBs must
be linked to death of massive stars. "There is tentative evidence that the GRBs
970228, 970508, and 970828 were close to star-forming regions. If this case is
strengthened with future afterglows, then the popular model in which GRBs are
caused by merging neutron stars will have to be abandoned, and a model linking
GRBs to cataclysmic deaths of massive stars will be favored."
In a typical manner Bohdan ignored the theoretical prejudice against this
idea. At the time baryonic contamination was a major concern. Within the
fireball model discussed above it was clear that a significant baryonic load will
result in a non-relativistic outflow which won't be able to drive a GRB. Collaps-
ing stars have large envelopes that could be a strong source of contamination.
However, observations are more important than theory and if the bursts are near
star forming regions they must involve stellar death. Bohdan outlines a model
based on Woosley's (1993) "failed supernova" on accretion and Blandford-Znajek
mechanism (1976) and suggests that GRBs operate like microquasars.
Once again Bohdan's ability to grasp the basic point from a dismal amount
of data leads to a great success. A few month later Galama et al., (1998)
discovered that a very powerful type Ic SN 1998bw is associated with GRB
980425! At the same time MacFadyen & Woosley (1999) demonstrate that a
sufficiently powerful jet can punch a hole in a stellar atmosphere, (provided that
the later is sufficiently small as would be the case if the Hydrogen envelope has
escaped). As more and more SN like bumps were discovered on GRB afterglow
light curves (Bloom et al., 1999) the model gained credibility. It was eventually
proven with GBR 030339/SN 2003dh where an 98bw like SN spectrum arose
just as expected from the GRB optical light curve (Staneck et al., 2003; Hjorth,
J., et al. 2003).
In the process Bohdan coined the name Hypernovae (Paczy´nski, 1998).
Bohdan anticipated that GRB associated collapse events are more powerful than
a regular supernovae and hence they should be given a name indicating that.
Hyper is clearly more powerful than Super. It turned out that the first GRB
associated supernova SN 1998bw was indeed much more powerful than average
and indeed GRB associated supernovae do show higher velocity ejecta and are
more powerful than other type Ic SNe.
Bohdan's GRBs
13
7. Some ideas for the future: Neutrinos, GRB remnants and optical
flashes preceding GRBs
Bohdan has left many ideas to be tested in the future.
7.1. GRB neutrinos
Already in 1994 Bohdan noticed (Paczy´nski & Xu, 1993) that internal shocks
within GRBs can produce high energy neutrinos. In this specific model proton-
proton collisions between protons from different shells produce pions,which in
turn produce the observed gamma-rays as well as 30 GeV neutrinos. In fact,
more energy is released in this case in neutrinos than in gamma-rays.
Proton-proton collisions are not very efficient and they can produce only
a weak GRB signal. To overcome this problem Paczy´nski and Xu proposed
that the bursts are very narrow beamed. As far as we understand GRBs today
both ideas are invalid. GRBs are beamed but the jets are much wider. Proton-
proton collisions are indeed not efficient enough to produce the observed prompt
emission. However, what we do take from this paper today is first the basic
concept of internal shocks within the outflows and second the idea that GRBs
are prime candidates for being sources of high energy neutrinos - an idea that is
generally accepted today (see e.g. Achterberg et al.(2007) for a description of a
recent search using AMANDA).
7.2. GRB remnants
Several times we have worked in parallel on related idea. The last one was in the
early 2000. It was realized at that time that GRBs are beamed (Rhodas, 1999;
Sari et al., 1999; Kulkarni et al., 1999). It was also realized that the afterglow, in
which the Lorentz factor is much lower is essentially less beamed. One expects
therefore, orphan afterglows. Afterglows whose GRB prompt emission does not
points towards us. An intriguing question was how to search for such orphan
afterglows. The late radio phase is naturally the longest one. But this phase is
the weakest. Additionally, if one waits too long the afterglow becomes spherical
and indistinguishable from a regular Supernova remnant.
However, for a period of 5000 years it is possible to see within the GRB
remnant a bipolar structure that is induced by the original nonspherical GRB
jets (Ayal & Piran, 2001). This will enable us to distinguish GRB remnants
from SNRs. Bohdan (Paczy´nski, 2001a) build on these ideas and estimated that
at any time there should be several dozen relatively nearby GRB radio remnants
which can be resolved with VLBA as being bipolar rather than spherical. He
suggested to combine this idea with other methods to detect imprints of GRBs
such as the effect of the original gamma-rays on the interstellar medium (Draine,
2000) and perform a systematic search for GRB remnants. Such a search was
not done yet. It is something that Bohdan has left for the future.
7.3. Optical Flashes
In recent years Bohdan was fascinated with the transient Universe (Paczy´nski,
2001). This was not surprising. After all both microlensing events and GRBs
to whom he devoted his research in recent years are transient phenomenon. He
kept constantly looking for new possible "flashes" that will shine in the night sky.
14
Tsvi Piran
Figure 5. Density profile of a GRB remnant as a function of µ ≡ M c2/E0,
where M is the accumulatd mass and E0 is the initial energy (from Ayal &
Piran, 2001).
Along this line Bohdan noticed an exciting idea of Beloborodov (2002) that the
prompt emission can cause a cascade of pairs at a distance of ∼ 1018cm from the
origin. The idea is simple. Some gamma-rays will interact with the surrounding
matter and will be reflected backwards. Each photon that is reflected backwards
will certainly interact with the outgoing radiation and produce a pair. The
produced pairs will increase the number of back scattered photons and will
accelerate the process causing a possible runaway. Bohdan suggested that with
the right conditions (Kumar & Panaiteschu, 2004) a cloud of pairs will form and
this cloud will be opaque to gamma-rays. Until this cloud is cleared away only
lower energy photons will be seen. Bohdan suggested, therefore, that an optical
flash will appear in such a case and that this flash will precede the gamma-rays
(Paczy´nski, 2001). He was worried, however, that such flashes might be missed,
as they appear before the GRB trigger, and he searched for method to detect
them.
8. Epilogue
Bohdan was dominant in convincing the community that GRBs are cosmolog-
ical. However, he also had far reaching contributions that shaped our current
Bohdan's GRBs
15
understanding of how GRBs operate. The scope of these contribution is best
realized by presenting once again Fig. 1, that describes the basic theory of
GRBs, but now with Bohdan's contributions superimposed on it. His ideas on
hypernova, links with star formation, internal shocks, afterglow that looks like
a SNR, and above all the cosmological origin of GRBs all paved the ways to the
present "standard" model.
Figure 6.
Bohdan's major contributions to the current GRB model.
The fast response satellite Swift that was launched three years ago have
changed to a large extend the simple picture that we had before. It turns out
that, like in other cases in Astronomy, the afterglow picture is more complicated
than what was originally thought. In particular the early X-ray light curve, as
seen by Swift is rather different from the previous expectations, showing unex-
pected rapid decline that is followed by a shallow phase before joining the more
familiar light curve at about 104sec (Nousek et al., 2006). At present it is not
clear what are the processes that control this light curve. One widely discussed
possibility is that energy is added to the blast wave during the shallow phase.
It is interesting to note that already in 1993 Bohdan discussed the possibility
that a significant amount of energy can come out from the inner engine in a
low Lorentz factor material (Paczy´nski & Xu, 1993). "The slower material with
a low Lorentz factor will be gradually added to the blast wave". This kind of
"energy injection" is in fact the leading interpretation today of this phase (see
e.g. Zhang et al., 2006; Granot, Konigl & Piran, 2006).
Another interesting issue is the fact that Swift has observed bursts that are
further out than what is expected if the bursts simply follow the SFR (Natarajan
et al., 2005,, Daigne et al., 2006; Guetta & Piran, 2007). There are more distance
bursts than the simple model suggests. This result is consistent with the fact
16
Tsvi Piran
that in space GRBs are more concentrated in high SFR regions, as if they follow
a higher power of the SFR. It could be that the relevant factor is low metalicity
(Fynbo, 2003) but the possibility of evolution of the luminosity function or
another unexpected factor cannot be ignored.
Other long standing open questions have existed before Swift and are still
with us: What is the exact working of the inner engine, what is the nature of
the relativistic outflow and how do collisionless shock accelerate particles and
generate magnetic fields? It is illuminating to read what were Bohdan's thoughts
about all that: "It is not likely that the concept of a GRB as a microquasar
powered by the Blandford & Znajek (1977) mechanism can be proven or disproven
on purely theoretical grounds. It is useful to realize, that while we have plenty
of sound evidence that Type II supernovae explode as a result of some 'bounce',
or whatever process following the formation of a hot neutron star, there is no
generally accepted physical process which would be efficient enough to make this
happen. The theoretical problem with the SN II explosions persists in spite of 2 or
3 decades of intense effort by a large number researchers. The problem is vastly
worse with the GRBs as they are 104 − 105 times less common than supernovae.
This might imply that a very special set of circumstances is necessary to generate
the suitable energetic explosion".
Bohdan was probably right. It will take time, a lot of detailed observations
and ingenious theoretical insight to figure out all those details. However, regard-
less of the different variants of the current model and the current observational
puzzles I am sure that Bohdan's basic ideas on how GRBs work will be with us
forever to stay.
Acknowledgments. This work was partially supported by an ISF center
for Excellence in High Energy Astrophysics.
References
Bohdan Paczy´nski publications on Gamma-Ray Bursts
Li, L.-X., & Paczy´nski, B. 2006, Improved correlation between the variability and peak
luminosity of gamma-ray bursts, MNRAS, 366, 219
Paczy´nski, B., & Haensel, P. 2005, Gamma-ray bursts from quark stars, MNRAS, 362,
L4
Gehrels, N., et al. 2005, Erratum: "The Swift Gamma-Ray Burst Mission" ApJ, 611,
1005 [2004]), ApJ, 621, 558
Stockdale, C. J., Van Dyk, S. D., Weiler, K. W., Sramek, R. A., Panagia, N., Rupen,
M. P., Paczynski, B., & Weiler, K. W. 2004, VLBA Observations of SN 2001em:
Supernova, Misdirected Gamma-Ray Burster, or Both?, Bulletin of the American
Astronomical Society, 36, 1464
Gehrels, N., et al. 2004, The Swift Gamma-Ray Burst Mission, ApJ, 611, 1005
Paczynski, B. 2001, Gamma-Ray Bursts at Low Redshift, Acta Astronomica, 51, 1
Paczynski, B. 2001, Gamma-ray burst - supernova relation, Supernovae and Gamma-
Ray Bursts: the Greatest Explosions since the Big Bang, 1
Paczy´nski, B. 2001, Monitoring Variability of the Sky, IAU Colloq. 183: Small Telescope
Astronomy on Global Scales, 246, 45
Paczy´nski, B. 2000, Monitoring All Sky for Variability, PASP, 112, 1281
Paczynski, B. 1999, Gamma-Ray Burst - Supernova Relation, ArXiv Astrophysics e-
prints, arXiv:astro-ph/9909048
Li, L.-X., & Paczy´nski, B. 1998, Transient Events from Neutron Star Mergers, ApJ,
507, L59
Bohdan's GRBs
17
Paczynski, B. 1998, Are Gamma-Ray Bursts in Star-Forming Regions?, ApJ, 494, L45
Paczynski, B. 1998, Gamma-Ray Bursts as Hypernovae, American Institute of Physics
Conference Series, 428, 783
Paczynski, B. 1996, Studies of Gamma-Ray Bursts Time Variability, Princeton Univ. Re-
port,
Paczynski, B. 1995, How Far Away Are Gamma-Ray Bursters?, PASP, 107, 1167
Wijers, R. A. M. J., & Paczynski, B. 1994, On the nature of gamma-ray burst time
dilations, ApJ, 437, L107
Paczy´nski, B. 1994, Cosmic Gamma-ray bursts, 26th Meeting of the Polish Astronomical
Society, 59
Paczynski, B., & Xu, G. 1994, Neutrino bursts from gamma-ray bursts, ApJ, 427, 708
Paczy´nski, B. 1994, Cosmic bursts of gamma radiation., Urania, 65, 41
Paczy´nski, B., Rhoads, J., & Xu, G. 1994, Radio and Neutrino Emission from Theoret-
ical Gamma-Ray Bursters, Gamma-Ray Bursts, 307, 542
Rhoads, J. E., & Paczynski, B. 1993, Radio Transients from Gamma-Ray Bursters,
BAAS, 25, 1296
Paczynski, B., & Rhoads, J. E. 1993, Radio Transients from Gamma-Ray Bursters,
ApJ, 418, L5
Stanek, K. Z., Paczynski, B., & Goodman, J. 1993, Features in the spectra of gamma-ray
bursts, ApJ, 413, L7
Paczy´nski, B. 1993, Gamma-Ray Bursts, Texas/PASCOS '92: Relativistic Astrophysics
and Particle Cosmology, 688, 321
Paczy´nski, B. 1993, Gamma-ray bursts., American Institute of Physics Conference Se-
ries, 280, 981
Narayan, R., Paczynski, B., & Piran, T. 1992, Gamma-ray bursts as the death throes
of massive binary stars, ApJ, 395, L83
Paczynski, B. 1992, Gamma-Ray Bursts: Facts and Fantasies, BAAS, 24, 751
Mao, S., & Paczynski, B. 1992, On the Galactic disk and halo models of gamma-ray
bursts, ApJ, 389, L13
Mao, S., & Paczynski, B. 1992, On the cosmological origin of gamma-ray bursts, ApJ,
388, L45
Paczynski, B. 1992, Estimating redshifts for gamma-ray bursts, Nat, 355, 521
Paczynski, B. 1992, On the galactic origin of gamma ray bursts, The Compton Obser-
vatory Science Workshop, 287
Paczy´nski, B. 1992, Gamma-ray bursts from colliding neutron stars., Los Alamos Work-
shop on Gamma-Ray Bursts, p. 67 - 74, 67
Paczynski, B. 1992, Gamma-ray bursts from colliding neutron stars, Gamma-Ray Bursts
- Observations, Analyses and Theories, 67
Paczynski, B. 1992, How Far Away are Gamma-Ray Bursters, Comments on Astro-
physics, 16, 241
Paczynski, B. 1992, Extragalactic scenarios for gamma-ray bursts, American Institute
of Physics Conference Series, 265, 144
Paczynski, B. 1992, GB 790305 as a very strongly magnetized neutron star, Acta As-
tronomica, 42, 145
Paczynski, B. 1992, On the two population model for gamma-ray bursts, Acta Astro-
nomica, 42, 1
Haensel, P., Amsterdamski, P., & Paczy´nski, B. 1991, Gamma-ray bursts from coalesc-
ing strange stars., Nuclear Physics B Proceedings Supplements, 24, 162
Haensel, P., Paczynski, B., & Amsterdamski, P. 1991, Gamma-ray bursts from colliding
strange stars, ApJ, 375, 209
Abramowicz, M. A., Novikov, I. D., & Paczynski, B. 1991, The appearance of highly
relativistic, spherically symmetric stellar winds, ApJ, 369, 175
Zdziarski, A. A., Svensson, R., & Paczynski, B. 1991, Bursts of gamma rays from
Compton scattering at cosmological distances, ApJ, 366, 343
Paczynski, B. 1991, Cosmological gamma-ray bursts, Acta Astronomica, 41, 257
18
Tsvi Piran
Paczynski, B. 1991, On the Galactic origin of gamma-ray bursts, Acta Astronomica, 41,
157
Paczynski, B. 1990, A test of the galactic origin of gamma-ray bursts, ApJ, 348, 485
Paczy´nski, B. 1989, Electromagnetic fireball for gamma-ray bursts., ST ScI-GSFC Work-
shop on Ultra-Hot Plasmas and Electron-Positron Pairs in Astrophysics, p. 47,
47
Paczynski, B. 1989, Comet Showers and Gamma-Ray Bursts, Nat, 337, 689
Paczynski, B. 1988, Gamma-ray bursts from cusps on superconducting cosmic strings
at large redshifts, ApJ, 335, 525
Paczynski, B., & Long, K. 1988, Distribution of intensities of gamma-ray bursts, ApJ,
333, 694
Paczynski, B. 1987, Gravitational microlensing and gamma-ray bursts, ApJ, 317, L51
Babul, A., Paczynski, B., & Spergel, D. 1987, Gamma-ray bursts from superconducting
cosmic strings at large redshifts, ApJ, 316, L49
Paczynski, B. 1986, Gamma-ray bursters at cosmological distances, ApJ, 308, L43
Other references:
Achterberg, A., et al. 2007, ApJ, 664, 397
Ayal, S., & Piran, T. 2001, ApJ, 555, 23
Beloborodov, A., 2002, ApJ, 565, 808
Blandford, R. D., & Znajek, R. L. 1977, MNRAS, 179, 433
Bloom, J. S., et al. 1999, Nat, 401, 453
Costa, E., et al. 1997, Nat, 387, 783
Daigne, F., Rossi, E., & Mochkovitch, R., 2006, MNRAS, 372, 1034
Djorgovski, S. G., et al. 2003, Proc. SPIE, 4834, 238
Draine, B. T. 2000, ApJ, 532, 273
Eichler, D., Livio, M., Piran, T., & Schramm, D. N. 1989, Nat, 340, 126
Fishman, G. J., et al. 1994, ApJS, 92, 229
Fishman, G. J., & Meegan, C. A. 1995, ARA&A, 33, 415
Frail, D. A., Kulkarni, S. R., Nicastro, L., Feroci, M., & Taylor, G. B. 1997, Nat, 389,
261
Fruchter A.S., et al., 2006, Nature, 441, 463
Goodman, J. 1986, ApJ, 308, L47
Fynbo, J. et al., 2003, A&A 406, L63.
Galama, T. J., et al. 1998, Nat, 395, 670
Goodman, J. 1997, New Astronomy, 2, 449
Hjorth, J., et al. 2003, Nature, 423, 847
Granot J., Konigl A. & Piran T., 2006, MNRAS, 370, 1946
Hakkila, J., Meegan, C. A., Pendleton, G. N., Fishman, G. J., Wilson, R. B., Paciesas,
W. S., Brock, M. N., & Horack, J. M. 1994, ApJ, 422, 659
Klebesadel, R. W., Strong, I. B., & Olson, R. A. 1973, ApJ, 182, L85
Kouveliotou, C., Meegan, C. A., Fishman, G. J., Bhat, N. P., Briggs, M. S., Koshut,
T. M., Paciesas, W. S., & Pendleton, G. N. 1993, ApJ, 413, L101
Kulkarni S. R., et al., 1999, Nature, 398, 389
Kumar, P., & Panaitescu, A. 2004, MNRAS, 354, 252
Lamb, D. Q. 1995, PASP, 107, 1152
Lingenfelter, R. E., & Higdon, J. C. 1992, Nat, 356, 132
Meegan, C. A., Fishman, G. J., Wilson, R. B., Horack, J. M., Brock, M. N., Paciesas,
W. S., Pendleton, G. N., & Kouveliotou, C. 1992, Nat, 355, 143
M´esz´aros P., Rees M.J., 1992, MNRAS, 257, 29P
Meszaros, P., & Rees, M. J. 1997, ApJ, 476, 232
M´esz´aros P., 2002, ARA&A, 40, 137
M´esz´aros P., 2006, Rep. Prog. Phys, 69, 2259
Metzger, M. R., Djorgovski, S. G., Kulkarni, S. R., Steidel, C. C., Adelberger, K. L.,
Frail, D. A., Costa, E., & Frontera, F. 1997, Nat, 387, 878
MacFadyen, A. I., & Woosley, S. E. 1999, ApJ, 524, 262
Bohdan's GRBs
19
Nakar E., et al., 2006, ApJ, 650, 281
Nakar E., 2007, PhR, 442, 166
Natarajan, P. et al., 2005, MNRAS 364, L8
Nousek J. A. et al., 2006, ApJ, 642, 389
Panaitescu A., Kumar P., 2002, ApJ, 571, 779
Piran, T. 1992, ApJ, 389, L45
Piran, T., & Shemi, A. 1993, ApJ, 403, L67
Piran T., 1995, in the Proceedings of "Some Unsoved Problems in Astrophysics", Prince-
ton April 1995, Eds. J. Bahcall and J. Osriker, Princeton University Press
Piran T., 1999, Phys Rep., 314, 575
Piran T., 2005, RMP, 76, 1143
Piran, T., Narayan, R., & Shemi, A. 1992, American Institute of Physics Conference
Series, 265, 149
Podsiadlowski, P., Rees, M. J., & Ruderman, M. 1995, MNRAS, 273, 755
Rees, M.J., & M´esz´aros P.,1994, ApJ, 430, 93
Rhoads, J.E., 1999, ApJ, 525, 737
Sari R., & Piran T., 1997, ApJ, 485, 270
Sari R., Piran T., & Narayan R., 1998, ApJ, 497, L17
Sari R., Piran T., Halpern J.P., 1999, ApJ, 519,
Shemi A., & Piran T., 1990, ApJ, 365, L55
Stanek, K. Z., et al. 2003, ApJ, 591, L17
Taylor, G. B., Frail, D. A., Berger, E., & Kulkarni, S. R. 2004, ApJ, 609, L1
van Paradijs, J., et al. 1997, Nat, 386, 686
.
Waxman, E. 1997, ApJ, 485, L5
Wijers R.A.M.J., Rees M.J., & M´esz´aros P., 1997, MNRAS, 288, L51
Wijers R.A.M.J., & Galama T.J., 1999, ApJ, 523, 177
Woosley S. E., 1993, ApJ, 405, 273
Woosley S. E., Bloom J. S., 2006, ARA&A, 44, 507
Yost S.A., et al., 2003, ApJ, 597, 459
Zhang, B., & M´esz´aros, P. 2004, IJMPA, 19, 2385
Zhang, B., Fan, Y. Z., Dyks, J., Kobayashi, S., M´esz´aros, P., Burrows, D. N., Nousek,
J. A., & Gehrels, N. 2006, ApJ, 642, 354
Zhang B., 2007, ChJAA, 7, 1
|
astro-ph/0203464 | 1 | 0203 | 2002-03-26T17:02:27 | The luminosity functions and stellar masses of galactic disks and spheroids | [
"astro-ph"
] | We present a method to obtain quantitative measures of galaxy morphology and apply it to a spectroscopic sample of field galaxies in order to determine the luminosity and stellar mass functions of galactic disks and spheroids. We estimate, for each galaxy, the bulge-to-disk luminosity ratio in the I-band using a two-dimensional image fitting procedure. Monte Carlo simulations indicate that reliable determinations are only possible for galaxies approximately two magnitudes brighter than the photometric completeness limit, leaving a sample of 90 galaxies with well determined bulge-to-total light ratios. We construct the luminosity functions of disks and spheroids and, using a stellar population synthesis model, we estimate the stellar mass functions of each of these components. The disk and spheroid luminosity functions are remarkably similar. We do, however, find evidence in the bi-variate luminosity function that spheroid-dominated galaxies occur only among the brightest spheroids, while disk-dominated galaxies span a much wider range of disk luminosities. Remarkably, the total stellar mass residing in disks and spheroids is approximately the same. For our sample, we find the ratio of stellar masses in disks and spheroids to be 1.3+/-0.2. Ongoing large photometric and redshift surveys will lead to a large increase in the number of galaxies to which our techniques can be applied and thus to an improvement in the current estimates. (abridged) | astro-ph | astro-ph |
The luminosity functions and stellar masses of galactic disks and
spheroids
A. J. Benson1 C. S. Frenk2 & R. M. Sharples2
1. California Institute of Technology, MC 105-24, Pasadena, CA 91125, U.S.A.
(e-mail: [email protected])
2. Physics Department, University of Durham, Durham, DH1 3LE, England
ABSTRACT
We present a method to obtain quantitative measures of galaxy morphol-
ogy and apply it to a spectroscopic sample of field galaxies in order to deter-
mine the luminosity and stellar mass functions of galactic disks and spheroids.
For our sample of approximately 600 galaxies we estimate, for each galaxy, the
bulge-to-disk luminosity ratio in the I-band using a two-dimensional image fit-
ting procedure. Monte Carlo simulations indicate that reliable determinations
are only possible for galaxies approximately two magnitudes brighter than the
photometric completeness limit, leaving a sample of 90 galaxies with well de-
termined bulge-to-total light ratios. Using our measurements of individual disk
and bulge luminosities for these 90 galaxies, we construct the luminosity func-
tions of disks and spheroids and, using a stellar population synthesis model, we
estimate the stellar mass functions of each of these components. The disk and
spheroid luminosity functions are remarkably similar, although our rather small
sample size precludes a detailed analysis. We do, however, find evidence in the bi-
variate luminosity function that spheroid-dominated galaxies occur only among
the brightest spheroids, while disk-dominated galaxies span a much wider range
of disk luminosities. Remarkably, the total stellar mass residing in disks and
spheroids is approximately the same. For our sample (which includes galaxies
brighter than M∗ + 2, where M∗ is the magnitude corresponding to the char-
acteristic luminosity), we find the ratio of stellar masses in disks and spheroids
to be 1.3 ± 0.2. This agrees with the earlier estimates of Schechter & Dressler,
but differs significantly from that of Fukugita, Hogan & Peebles. Ongoing large
photometric and redshift surveys will lead to a large increase in the number of
galaxies to which our techniques can be applied and thus to an improvement in
the current estimates.
-- 2 --
1.
Introduction
It has long been known that galaxies in the nearby Universe display a range of morpho-
logical characteristics that distinguish between disk-dominated (e.g. spiral) and spheroid-
dominated (e.g.
elliptical) galaxies, with many fine subdivisions within each class (e.g.
Hubble 1926; de Vaucouleurs 1959). The origins of these different types of galaxy and the
evolutionary connections, if any, between them are still unclear, although there is a wealth
of observational data and several proposed theories.
Traditionally, morphology has been assigned by human classifiers directly from galaxy
images, a process that is accurate only to within about two T-types (Naim et al. 1995).
More recently, computer-based algorithms have been developed which show a reasonable
correlation with "eyeball" estimates, at least for bright galaxies (e.g. Abraham et al. 1996).
Unfortunately, all these classifications tend to give considerable weight to detailed mor-
phological features, such as spiral arms or asymmetries in the image, and are difficult to
compare with current theoretical predictions which focus on simpler quantities such as the
total stellar mass or luminosity in the disk and spheroidal components. Here, we consider a
morphological quantifier that is more easily related to theoretical models.
It is widely believed that disks form by slow accretion of gas which acquired angular
momentum through tidal torques (Hoyle 1949; Peebles 1969), although whether this pic-
ture works in detail remains an open question (Navarro, Frenk & White 1995; Navarro &
Steinmetz 1997, 2000; van den Bosch, Burkert & Swaters 2001). Spheroids, on the other
hand, are thought to form either by a "monolithic collapse" (Eggen, Lyden-Bell & Sandage
1962; Jimenez et al. 1999), or as a result of mergers of pre-existing galaxies (Toomre 1977;
Barnes & Hernquist 1992 and references therein). Detailed theoretical predictions for the
statistical morphological properties of galaxies and their evolution have been calculated for
the hierarchical merging formation mechanism appropriate to cold dark matter cosmolo-
gies (Kauffmann, White & Guiderdoni 1993; Kauffmann 1995, 1996; Baugh, Cole & Frenk
1996a,b; Somerville, Primack & Faber 2001). Among other things, these models give the
relative luminosities and stellar masses of the spheroids and disks of galaxies.
In this work, we measure the I-band bulge-to-total light ratio, B/T, for a large sample
of galaxies with spectroscopic redshifts by fitting two-dimensional models to the observed
galaxy images. We use this information to estimate the spheroid and disk luminosity func-
tions, as well as the total stellar mass that resides in disks and spheroids. An earlier attempt
to estimate the relative contributions to the luminosity from spheroids and disks was carried
out by Schechter & Dressler (1987), based on "eyeball" estimates of the bulge-to-total ratio.
The remainder of this paper is arranged as follows. In §2 we describe our basic dataset.
-- 3 --
In §3 we introduce the method used to fit model images to the data and thereby extract
B/T (along with other interesting parameters) and describe how we estimate errors. In §4
we examine the accuracy of our technique and determine how well the B/T ratio can be
measured as a function of the apparent magnitude of a galaxy. We then compute luminosity
functions and total stellar masses for spheroids and disks. Finally, in §5 we present our
conclusions.
2. Data
The galaxy sample used in this work is that of Gardner et al. (1997). The reader
is referred to that work for a full description of the data. Here we summarize the most
important features of the dataset.
Imaging of two fields of total area 10 deg2 was carried out in the B, V and I bands
using the T2KA camera on the Kitt Peak National Observatory (KPNO) 0.9-m telescope,
resulting in images with 0.68-arcsec per pixel. Exposures of 300s reached 5σ detection depths
of B=21.1, V=20.9 and I=19.6 in 10-arcsecond circular apertures. Imaging was also carried
out in the K-band using the IRIM camera on the KPNO 1.3-m telescope resulting in 1.96-
arcsecond per pixel images, and a 5σ detection depth of K=15.6 in a 10-arcsecond circular
aperture. The positions of the fields were chosen randomly (the field centers are RA 14h15m,
Dec. +00◦ and RA 18h0m, Dec. +66◦). The I-band images, which we will use in this work,
were bias-subtracted, flattened using twilight flats and with median sky flats. Objects were
identified with the sextractor program (Bertin & Arnouts 1996), using a 3σ threshold.
The seeing in the optical images varied in the range 1.3 < FWHM < 2.0 arcsec. One field
contained a nearby rich galaxy cluster.
Spectroscopic follow-up was obtained for a K-selected sample in sub-regions of total
area 4.4 deg2, using the Autofib-2 fiber positioner and WYFFOS spectrograph on the 4.2m
William Herschel Telescope on La Palma. Spectra were obtained for 567 galaxies with K< 15,
which allowed redshifts to be measured for 510 galaxies (a redshift completeness of 90%).
Although the spectroscopic sample is K-selected, this does not introduce any incompleteness
in the I< 16 sample used extensively in this work (i.e. there are no galaxies with I−K< 1
in the sample). We also briefly consider an I< 18 sample, for which the spectroscopic
completeness falls to around 50% because of the K-band selection (which also introduces a
bias in this fainter sample against objects that are blue in I−K). For the I< 16 and I< 18
samples the median redshift is z = 0.08 and 0.14 respectively.
-- 4 --
3. Method
Wadadekar, Robbason & Kembhavi (1999) have proposed a two-dimensional galaxy
decomposition technique that can efficiently recover B/T ratios (and other parameters) of
model galaxy images with high accuracy (see also Byun & Freeman 1995; de Jong 1996).
They present a detailed study of the effects of uncertainties in the point spread function
(PSF), the presence of nearby stars and the stability of the B/T estimates as a function
of signal-to-noise. Our approach is similar to theirs, but we apply the technique to a large
photometric sample of real galaxies1. We consider the "real-world" problems of automated
masking of nearby galaxies and stars and make a thorough assessment of the errors in
the measured parameters. We also present, in an Appendix, estimators for the luminosity
functions of disks and spheroids.
For this analysis, we have used the sample of 636 I < 18 galaxies of Gardner et al. (1997)
imaged in B, V, I and K, as described in §3. Postage stamp images of 33×33 pixels (0.68′′/pix)
around each galaxy were extracted from the I-band data. This was the best observed band
by Gardner et al. (1997) and is particularly well-suited for our purposes because it minimizes
the effects of young blue stars. This image size is large enough to include the entire region of
the galaxy for which reasonable signal-to-noise is achieved (and in most cases extends well
beyond it.) To determine the bulge-to-total ratio, B/T, we fit the two-dimensional surface
brightness profile of each galaxy using a combination of an exponential disk,
and an r1/4-law spheroid,
Σd(θ) = Σd,0 exp(−θ/θd),
Σs(θ) = Σs,e exp(−7.67[(θ/θe)1/4 − 1]),
(1)
(2)
where θ is the angular distance from the galaxy center. We will also consider a more general
r1/n spheroid profile, as Wadadekar, Robbason & Kembhavi (1999) did. The disk is allowed
to be inclined, and to have arbitrary position angle. The spheroid is allowed an ellipticity
(defined as the ratio of semi-major to semi-minor axes) in the range 1 to 6, and can also
have an arbitrary position angle. To mimic seeing, we construct mock images using these
profiles which we then smooth with a Gaussian filter (integrated over each pixel to account
for the variation of the PSF across the pixel),
p(θ) = exp[−(θ/σ)2/2]/2πσ2.
(3)
1Wadadekar, Robbason & Kembhavi (1999) applied their technique to three galaxies for which previous
estimates of B/T were available and found reasonable agreement.
-- 5 --
The width of the Gaussian is treated as a free parameter, to account for variations in seeing
between the images. (We will examine in §4.1 the effect of using a more realistic PSF.) The
postage stamp images were centered on the galaxy of interest, but we allow the position of
the image center to vary since in many cases the resulting sub-pixel variations lead to lower
values of χ2. We also allow a small contribution from a faint, constant surface brightness
background in order to take into account small inaccuracies in sky subtraction. The best
fitting parameters for each galaxy were then obtained by minimizing χ2 using Powell's algo-
rithm (Brent 1973). There are a total of 12 fit parameters (13 if we include n when fitting
r1/n spheroid profiles) summarized in Table 1. In §4.1 we will consider how accurately this
procedure recovers the B/T ratio of the galaxies.
A significant fraction of the postage stamp images was contaminated by a secondary
galaxy (and occasionally by more than one). We use a simple algorithm to identify such
contaminants and mask them from the image. The aim is to remove objects which are
physically distinct from the galaxy of interest without masking any pixels of the galaxy
itself. We first rank the pixels in the image by surface brightness and then proceed to
find groups of bright pixels. The brightest pixel is assigned to the first group. Successive
pixels are assigned to a pre-existing group if they touch it (i.e.
if they are adjacent either
horizontally, vertically or diagonally), or else are assigned to a new group. In the case where
a pixel touches more than one group, the two groups are merged. This process is continued
until pixels of a fixed signal-to-noise are reached (specifically, we consider only pixels more
than 3σ above the sky background). If more than one group exists at this point, the group
at the center is deemed to be the galaxy of interest and the pixels of all secondary groups are
marked as being contaminated and are not included in the χ2 sum. This simple algorithm
works well in the majority of cases, but it fails in a few (28 out of 636 galaxies), either by
not removing a contaminating galaxy, or by removing a significant fraction of the primary
galaxy. Rather than attempting to use a more complex algorithm in these cases, we resorted
to cleaning the image by hand (i.e. we view the image and manually mark the contaminated
pixels).
-- 6 --
Table 1: The parameters used to construct mock galaxy images in the fitting procedure.
Each parameter is described in the text.
Σd,0 Disk central surface brightness
Σs,e
Spheroid surface brightness at the ef-
fective radius
Disk angular scalelength
Spheroid effective radius
Disk position angle
Spheroid position angle
Disk inclination angle
Spheroid ellipticity
θd
θe
Pd
Ps
i
e
(x, y) Center of image
B
σ
n
Excess background surface brightness
Seeing (c.f. eqn. 3)
Spheroid profile index (fixed at n = 4
unless otherwise stated)
-- 7 --
Fig. 1. -- (cont.) Postage stamp images of three representative galaxies from our sample.
The top row shows a galaxy that is fit very well by our procedure, the bottom row shows
one that is poorly fit and the middle row shows a more typical result. The left hand column
is the original 33 × 33 pixel galaxy image with contours indicating the pixel value in ADUs.
The right hand column is the residual image after subtracting the best fit model galaxy.
Contours show the absolute value of the residual in units of σ, the rms uncertainty on
each pixel value. Hatched regions contained contaminant galaxies and were removed by our
automated cleaning procedure before fitting. (Where an entire row or column is hatched,
the postage stamp image was recentered prior to fitting.) Between the original image and
the residual maps we quote the value of χ2 per degree of freedom, Q (the probability that
a random fluctuation exceeds this value of χ2) and the I-band apparent magnitude. Also
shown is a histogram of dP/d(B/T), the distribution of bulge-to-total ratios found from the
Monte Carlo simulations described in the text, with a vertical dashed line indicating the
best-fit B/T value for the original image.
-- 8 --
The majority of galaxies (363 out of 626) are fit reasonably well by this procedure.
We regard a galaxy as being reasonably well fit when Q, the probability that the measured
value of χ2 is exceeded by random fluctuations, is greater than 5%. Not surprisingly though,
many galaxies are not well fit. These typically show signs of strong morphological disturbance
(perhaps due to a recent or imminent merger) or other inhomogeneities. Examples of well-fit
and poorly-fit galaxies are given in Figure 1. Poorly-fit galaxies are easily identified by their
large χ2 values and so may be excluded from further analysis if desired. It should be noted
that a poor-fit does not indicate a failure of our fitting procedure per se, rather it signals
that the galaxy is not well described by a combination of a spheroid and a disk. We choose
to show results computed using the entire sample, regardless of how well a galaxy was fit,
but we will comment on how our results change if badly fit galaxies are excluded from the
analysis.
Errors on the fitted parameters could, in principle, be determined using a ∆χ2 approach,
but this would require mapping χ2 in the 12 dimensional parameter space of the fit --
an exceedingly time consuming exercise -- and, in any case, the errors are unlikely to be
normally distributed given that the model is highly nonlinear in the parameters. We therefore
adopt a Monte Carlo approach to error estimation. Using the best fitting model for each
galaxy, we generate 30 realizations of that model, add random noise at the same level as in
the real image, and mask out any pixels which were masked out in the original. We then
find the best fitting parameters for each realization and take their distribution as indicative
of the uncertainties in the actual fit. It should be noted that this is only a valid procedure
if the original image is well fit by the model. In cases where this is not the case, there is no
reason to expect the Monte Carlo distributions to give an estimate of the true errors. The
B/T distributions for the three galaxies illustrated in Figure 1 are shown in that figure.
4. Results
4.1. Accuracy Checks
We begin by assessing the reliability of our procedure for recovering the true B/T ratio of
a galaxy (assuming, of course, that real galaxies are well described by our model). Our Monte
Carlo procedure for error estimation allows a determination of the accuracy of our technique.
For each galaxy, the value of B/T input into the Monte Carlo simulations may be compared
to the mean and standard deviation of the distribution of 30 recovered B/T values. Figure 2
gives the results of these accuracy tests. The left-hand panel shows the standard deviation of
the recovered B/T ratio, as a function of the I-band apparent magnitude of the mock image.
For bright galaxies (mI <∼16), σMC is fairly small, typically less than about 0.1. However,
-- 9 --
for fainter galaxies, σMC increases very rapidly, resulting in rather poorly constrained B/T
values. (In reality, σMC depends also upon the other parameters that describe the mock
image, but the correlation with apparent magnitude is the most important.)
In the right-hand panel of Figure 2, we plot the mean value of B/T recovered from the
Monte Carlo simulations against the true value for the mock image. The large solid circles
indicate those images for which σMC ≤ 0.1. Evidently, for these galaxies the value of B/T
is recovered accurately and without any strong systematic bias. The small dots show the
results for all other galaxies. Now the scatter is much larger and, more importantly, there
are systematic biases in the mean recovered B/T, such that very low and very high values
are avoided. This effect is not surprising: the values of B/T for these faint galaxies are
almost entirely unconstrained. (Note that the standard deviation for a completely uniform
distribution of B/T is approximately 0.3.) As a result, the distribution of B/T from the
Monte Carlo simulations becomes close to uniform, with the mean tending towards 0.5 as
σMC increases. For the 90 galaxies in our sample with mI ≤ 16, there is a tight correlation
between (B/T )true and the mean value recovered from the Monte Carlo simulations and it is
this subsample that we will use below to compute luminosity functions. Unfortunately, its
relatively small size limits the statistical accuracy of our estimates quite considerably.
-- 10 --
Fig. 2. -- Left-hand panel: the standard deviation, σMC, of the distribution of recovered
B/T ratios from 30 Monte Carlo realizations of a model galaxy image, as a function of the
apparent I-band magnitude of the model image. Right-hand panel: the mean recovered B/T
ratio from 30 Monte Carlo realizations of a model galaxy image plotted against the true B/T
value. Large, filled circles show those galaxies for which σMC ≤ 0.1, while small open circles
show all other galaxies.
-- 11 --
To find the best fit solution we must choose initial values for the parameters to be fit and
then use Powell's algorithm to search for values producing a better fit. For parameters such
as the position angles, disk inclination and spheroid/disk sizes we make initial guesses based
on the image being fitted. Other parameters are initially assigned "typical" values. We have
checked the effect of altering these initial values. For galaxies with σMC ≤ 0.1, the choice of
initial value makes almost no difference to the recovered values of the parameters, indicating
that our technique is finding the true minimum χ2. As σMC becomes larger, however, the
recovered parameters begin to depend strongly upon the initial values chosen. For these
images, the χ2 surface in the 12 dimensional parameter space does not possess an obvious
minimum (i.e.
it is very noisy). This is just another way of saying that the values of the
fitted parameters for these faint images are highly uncertain.
Andredakis, Peletier & Balcells (1995) have demonstrated that the bulges of spiral
galaxies are more accurately fit by an r1/n, rather than by the more usual r1/4, surface
brightness profile, with values of n ranging from around 1 to 6 (similar variations in n are
seen for elliptical galaxies; Bingelli & Cameron 1991; Caon, Capaccioli & D'Onofrio 1993).
They show that the value of n is strongly correlated with morphological type. Although
our present data provide only rather poor constraints on the value of n (due to limited
angular resolution and signal-to-noise), we have nevertheless repeated the fitting procedure
using r1/n profiles for the spheroids, treating n as a free parameter. For galaxies where the
B/T ratio is well determined, we find that there is a strong correlation between the B/T
values obtained with r1/4 and r1/n profiles, although inevitably some scatter is present. The
disk and spheroid luminosity functions computed using B/T ratios from r1/n fits show no
statistically significant difference from those using r1/4 fits.
Finally, we remind the reader that our analysis makes use of a Gaussian PSF to mimic
the effects of seeing in the data. A Gaussian accurately describes the core of the PSF
measured from bright stars in the images. However, a profile consisting of a Gaussian
core plus power-law wings provides a better match to many of the stellar profiles.
(The
variation of Gaussian and core components from night to night in the imaging data is not
so well characterized, however, and this is why we make use of a simple Gaussian for our
main analysis.) Fitting the images using such a profile (keeping the relative proportions of
Gaussian and power-law wings fixed, but allowing the overall radial scale of the PSF to be a
free parameter) results in small changes in the B/T ratio, typically significantly smaller than
the error in the best fit value. Thus, the luminosity functions presented below are unaffected
by the exact choice of PSF. However, it is clear that a good characterization of the PSF and
its variation will be crucial to obtain accurate disk and spheroid luminosity functions from
larger, higher quality datasets.
-- 12 --
4.2. Luminosity Functions
Using the I< 16 sample of approximately 90 galaxies for which we have good estimates
of the B/T ratio we now proceed to estimate the disk and spheroid luminosity functions.
Our aim here is to develop the techniques required for this measurement and demonstrate
them using a particular dataset. Given the small size of the dataset we must expect that
both statistical (due to the small number of galaxies) and systematic (due, for example, to
the lack of rich clusters in the dataset) errors will be present. These issues are considered
further in §5.
To determine the present-day luminosity functions, we need to apply k+e corrections to
the galaxy luminosities. We use the type-dependent k+e corrections obtained by Gardner
et al. (1997). Briefly, a set of model galaxy colors was computed using an updated version
of the Bruzual & Charlot (1993) stellar population models with a range of star formation
histories. The observed colors of each galaxy were matched to one of the models and that
particular model was then used to extrapolate the observed galaxy luminosity to z = 0. Note
that our type-dependent k+e corrections are based on the total (i.e. disk plus spheroid) color
of each galaxy. In principle, k+e corrections could be applied to each component separately
if spheroid/disk decompositions were carried out in several bands. Given the uncertainties
in our present estimates of B/T, we refrain from this degree of complexity in this analysis.
We use the stepwise maximum likelihood (SWML) estimator proposed by (Efstathiou,
Ellis & Peterson 1988, hereafter EEP) and also the parametric maximum likelihood method
proposed by (Sandage, Tammann & Yahil 1979, hereafter STY) to compute disk and spheroid
luminosity functions. The detectability of a spheroid depends on both its apparent magnitude
and the B/T ratio, and we must account for this in constructing the likelihood function.
This leads us to define a two-dimensional function, Φ(M, B), such that Φ(M, B)dMdB is
the number of galaxies per unit volume with B/T ratio in the range B to B + dB and
spheroid absolute magnitude in the range M to M + dM (with an equivalent definition for
disks). The application of the maximum likelihood estimator to this function is discussed
in detail in Appendix A. The normal luminosity function of spheroids is readily derived
0 Φ(M, B)dB (and similarly for disks). For the STY method we must
assume some parametric form for the luminosity function. We have tried fitting the disk and
spheroid luminosity functions with a "Schechter⊗exponential" form, namely Φ(M, B) =
φ(M) exp(βB), where φ(M) is the normal Schechter function and β is a parameter to be fit,
motivated by the shape of the SWML estimate of these luminosity functions.
using φ(M) = R 1
-- 13 --
Fig. 3. -- I-band luminosity functions. Triangles, circles and stars show the SWML estimates
of the total, spheroid and disk luminosity functions respectively. Only galaxies brighter than
I = 16 have been used, k+e corrections have been applied to all galaxies, and distances
have been calculated assuming (Ω0, Λ0) = (0.3, 0.7). Errorbars are the sum in quadrature
of the standard SWML errors and the variance in estimates of the luminosity function from
30 Monte Carlo realizations of the spheroid/disk decomposition procedure. The very heavy
solid line shows the best fit Schechter function to the total luminosity function, while the
inset shows the values of α and M⋆ for this fit, together with their 1 and 2σ error ellipses.
The heavy and thin solid lines show the best-fit "Schechter⊗exponential" functions to the
spheroid and disk luminosity functions respectively, and confidence regions for α and M⋆ for
these fits are given in the inset (the remaining parameter of the fits was β = 0.0 ± 0.37 and
2.1 ± 0.37 for spheroids and disks respectively).
-- 14 --
Figure 3 shows the resulting I-band luminosity functions with distances computed as-
suming (Ω0, Λ0) = (0.3, 0.7)2 and H0 = 100h km/s/Mpc. Triangles show the total luminosity
function; circles and stars show the spheroid and disk luminosity functions separately. These
SWML luminosity functions are normalized to the I-band number counts in the Sloan Dig-
ital Sky Survey using the procedure described in the Appendix. For the total luminosity
function, we plot the standard SWML errorbars (obtained from the covariance matrix of
the luminosity function as described by EEP), but for the spheroid and disk luminosity
functions, the errorbars are the sum in quadrature of the standard SWML errors and the
variance in the luminosity function estimated from the 30 Monte Carlo realizations of the
spheroid/disk decomposition process. The errors from each source are of comparable mag-
nitude (although the variance from the Monte Carlo realizations is the smaller of the two).
The very heavy solid line shows the best-fitting Schechter function to the total luminosity
function (determined using the STY method); the inset shows the values of α and M⋆ for this
fit, together with their 1 and 2σ error contours. (The small sample size is reflected in rather
large and correlated uncertainties in M⋆ and α.) Heavy and thin solid lines show the best fit
STY "Schechter⊗exponential" luminosity function fits to the spheroid and disk luminosity
functions respectively (with the corresponding confidence ellipses for α and M⋆ shown in the
inset, and the values of β given in the figure caption). A likelihood ratio test (Efstathiou,
Ellis & Peterson 1988) shows that the Schechter⊗exponential luminosity function is not a
particularly good fit to the data. With the present small dataset we have been unable to
find a better functional form. This situation will be rectified with a larger dataset (assuming
that some suitable functional form does actually exist).
The I-band luminosity functions of disks and spheroids are remarkably similar. The
only significant difference is that the spheroid luminosity function is somewhat lower at faint
magnitudes. However, given the small size of the present sample, this difference may not be
robust. The luminosity densities in disks and spheroids obtained by integrating the SWML
luminosity functions over the range of absolute magnitudes shown in Fig. 3 are 5.8 ± 0.8
and 4.7 ± 0.7 × 107hL⊙/Mpc3 respectively (where we have taken M⊙ = 4.14 in the I-band;
Cox 2000). In principle, we can use our Schechter⊗exponential fits to estimate the total
luminosity density, extrapolating to include the contribution from arbitrarily faint spheroids
and disks. Doing so yields results which agree with the SWML estimates within the quoted
errors, suggesting that our determination may have suitably converged. However, it must
be kept in mind that the Schechter⊗exponential form is not a particularly good fit to the
current datasets.
2Assuming (Ω0 , Λ0) = (1, 0) instead changes our results only slightly, shifting the luminosity function
faintwards due to the smaller luminosity distance in this model.
-- 15 --
Fig. 4. -- Slices through the SWML estimate of the bi-variate luminosity function, Φ(M, B),
for different absolute magnitudes, M, as indicated in the figure. Points were computed in bins
of size (∆M, ∆B) = (0.48, 0.33) and errors obtained as described by EEP. The upper panel
shows the spheroid luminosity function, and the lower panel the disk luminosity function.
The points without errorbars above the dotted line in each panel indicate the mean B/T
(and D/T) for spheroids (and disks) in the corresponding absolute magnitude bins. Solid
lines indicate the best-fit "Schechter⊗exponential" parametric luminosity function for the
MI − 5 log h = −19.82 bin.
-- 16 --
In Fig. 4, we show slices through the bi-variate luminosity function, Φ(M, B), at constant
M for several values of M (as indicated in the figure, and in bins of width ∆M = 0.48). It
is evident that Φ(M, B) is not independent of B and that, in fact, it may not be separable
into a simpler form Φ(M, B) = φ(M)g(B). This is particularly noticeable for the spheroid
luminosity function. Figure 4 shows that spheroid-dominated systems (i.e. B/T> 2/3) are
found only in the brightest spheroids, while disk-dominated systems (i.e. D/T> 2/3) have
disks with a much broader range of luminosities. This point is made more clearly in Fig. 5
where we show the spheroid luminosity function of spheroid-dominated systems and the
disk luminosity function of disk-dominated systems. We find spheroid-dominated systems in
abundance only brightwards of MI − 5 log h ≈ −21, but disk-dominated systems across the
whole range of luminosities.
-- 17 --
Fig. 5. -- SWML estimates of the I-band luminosity functions of spheroids in spheroid-
dominated galaxies (i.e. B> 2/3; circles) and disks in disk-dominated galaxies (i.e. D/T>
2/3; stars). The samples include 13 and 36 galaxies respectively. Errors are the sum in
quadrature of the standard SWML errors and the variance found in luminosity functions
estimated from 30 Monte Carlo realizations of the spheroid/disk decomposition procedure.
-- 18 --
4.3. The stellar mass in disks and spheroids
The stellar mass associated with the luminosity of each galaxy is easily obtained by
a similar procedure to that employed to calculate the k-corrections (namely fitting their
BVIK colors to a set of template galaxies). We use these estimates to construct the stellar
mass functions of spheroids and disks in the local Universe and thereby estimate the total
mass content in each component. The method that we adopt is the same that Cole et al.
(2001) used to measure the total stellar mass density in the Universe from a combination
of 2dFGRS and 2MASS data. Cole et al. (2001) found the stellar mass density3 in units
of the critical density to be Ωstars = 0.0016 ± 0.00024h−1 or Ωstars = 0.0029 ± 0.00043h−1,
depending on whether a Kennicutt (1983) or a Salpeter (1955) stellar initial mass function
(IMF) was assumed.
(Both estimates include the effects of dust on galaxy luminosities
as described by Cole et al. 2001.) We normalize our stellar mass functions by requiring
them to produce the same number density of galaxies more massive than M∗ as the Cole
et al. (2001) stellar mass function. The total stellar mass density inferred from our I < 18
sample is then Ωstars = 0.0009 ± 0.00007h−1 or Ωstars = 0.0017 ± 0.00012h−1 for the same
two IMFs respectively and the same prescription for dust-extinction. Errors on the stellar
mass density were found by summing in quadrature the error from each individual bin in the
SWML mass function, together with the error in the overall normalization. Our estimates
include contributions from galaxies with stellar masses greater than 109h−2M⊙ below which
the SWML stellar mass function is not well determined. We can check our result using the
STY stellar mass function. For this, we fit a Schechter function convolved with a Gaussian
of width 0.1 in log10 Mstars to account for the scatter in the relation between stellar mass and
I-band absolute magnitude. We find that our estimates of Ωstars using the SWML and STY
mass functions agree within the errors, suggesting that the result has already converged to
sufficient accuracy. Our estimates, however, are lower than those of Cole et al. (2001), a
reflection of the flatter faint end slope of our mass functions which may well be due to the
small size of our sample.
For our I < 16 sample, for which the bulge-to-total ratio is well measured, we find a
slightly higher total stellar mass density of Ωstars = 0.0012 ± 0.00014h−1 for the Kennicutt
(1983) IMF. Splitting into spheroidal and disk components, we find Ωstars,spheroids = 0.00039±
0.00006h−1 and Ωstars,disks = 0.00051 ± 0.00008h−1 for this same IMF. (Note that with the
SWML method the stellar mass densities of disks and spheroids are not guaranteed to sum
to give the total stellar mass density.) If, instead, we assume a Salpeter IMF, the ratio of
3Specifically, Cole et al. (2001) estimated the mass locked up in stars and stellar remnants which differs
from the time integral of the star formation rate due to recycling of material by massive stars. We adopt
the same definition of stellar mass here.
-- 19 --
disk to spheroid stellar mass densities increases slightly from 1.31 to 1.37, but this change is
negligible given the current errors in these quantities. Although the small size of our sample is
clearly a significant limitation, this initial result suggests that spheroids and disks contribute
about equally to the stellar mass density of the Universe. The techniques developed in this
paper, when applied to a much larger dataset, should allow their contributions to be more
accurately determined.
5. Discussion
We have presented a detailed method to determine the bulge-to-total ratios of galaxies
by fitting to two-dimensional photometry, and have applied this technique to determine the
I-band bulge-to-total luminosity ratios of a sample of approximately 600 galaxies brighter
than I= 18 with spectroscopic redshifts. Our approach is designed to work with realistic
galaxy images, dealing automatically with contamination by nearby objects, a varying PSF
and small changes in the background from image to image. A crucial part of the fitting
procedure is a Monte Carlo determination of the errors on the fitted parameters, an approach
which is favored since it is fast and automatically accounts for the highly non-linear nature
of the model parameters. For the current sample of galaxies, around 60% are well fit by a
combination of an exponential disk and an r1/4-law spheroid. Those that are not well fit
frequently show signs of morphological disturbance. We find that bulge-to-total ratios are
determined accurately (i.e. with errors of around 10%) only for galaxies brighter than I≈ 16.
For the 90 galaxies brighter than I= 16 in this sample we measure the B/T ratio with
reasonable accuracy. We have used the resulting disk/spheroid decomposition of these bright
galaxies to construct separate luminosity functions for disks and spheroids. We find no
significant differences between them when considered purely in terms of luminosity, although
the statistical uncertainties associated with the small sample size make the detection of any
differences difficult. However, when we consider the bi-variate distributions of luminosity and
bulge-to-total or disk-to-total light, we find that spheroid dominated systems (B/T> 2/3)
only occur for the brightest spheroids, while disk-dominated systems (D/T> 2/3) occur for
a much broader range of disk luminosities.
The relative contributions of disks and spheroids to the total stellar mass density in the
Universe is a very important constraint on theories of galaxy formation which attempt to
describe the assembly of galaxies as a function of time. We find, perhaps surprisingly, that
the disks and spheroids in our sample contribute almost equally to the stellar mass density
today (in a ratio of 1.3 ± 0.2). Since the stellar populations in disks are generally younger
than those in spheroids, it is an interesting coincidence that the total stellar mass in the two
-- 20 --
kinds of structural components should be so similar at the present time.
Schechter & Dressler (1987) reached a similar conclusion to ours using a photometric
comparison technique to estimate the B-band bulge-to-disk ratios of galaxies. While this
technique may not be as accurate as our own on a galaxy-by-galaxy basis, it should provide
a good estimate of the total contribution of each component to the stellar mass density. It
is therefore reassuring that our results agree well with those of Schechter & Dressler (1987).
A different result was obtained by Fukugita, Hogan & Peebles (1998), who derived a ratio
of disk to spheroid stellar mass density of 0.33 ± 0.23. Although they found comparable B-
band luminosity density in spheroids and disks, they adopted a spheroid mass-to-light ratio
around four times greater than that for disks, resulting in spheroids making a significantly
greater contribution to the stellar mass density. While we use a more accurate technique for
converting from luminosity to stellar mass (a technique which could be improved further if
B/T ratios were measured for each galaxy in several bands), the small size of our sample
limits the accuracy of our results. In particular, our sample may not contain enough rich
clusters which are known to contain higher fractions of spheroid dominated galaxies than
the field (e.g. Dressler 1980) and this could introduce a small bias in our results.
Clearly the greatest limitation of this work is the small size of the sample of galaxies for
which accurate disk/spheroid decompositions can be performed. Fortunately, this problem
should be remedied in the near future with the advent of high quality, large area photometric
surveys such as that being carried out by the SDSS project.
Acknowledgments
We thank Jon Gardner and Carlton Baugh for supplying data used in this work and for
valuable discussions, and the referee, Alan Dressler, for valuable suggestions. We also thank
Istvan Szapudi for his assistance in the early stages of this work.
REFERENCES
Abraham R. G., van den Bergh S., Glazebrook K., Ellis R. S., Santiago B. X., Surma P.,
Griffiths R. E., 1996, ApJS, 107, 1
Andredakis Y. C., Peletier R. F., Balells M., 1995, MNRAS, 275, 874
Arnaud M., Rothenflug R., 1985, A&AS, 60, 425
Barnes J. E., Hernquist L., 1992, ARA&A, 30, 705
-- 21 --
Baugh C. M., Cole S., Frenk C. S., 1996a, MNRAS, 282, L27
Baugh C. M., Cole S., Frenk C. S., 1996b, MNRAS, 283, 1361
Bertin E., Arnouts S., 1996, A&AS, 117, 393
Bingelli B., Cameron L. M., 1991, A&A, 252, 27
Brent R. P., 1973, Algorithms for minimization without derivatives (Englewood Cliffs, New
Jersey: Prentice Hall), Chapter 7
Bruzual & Charlot, 1993, ApJ, 405, 538
Byun Y. I., Freeman K. C., 1995, ApJ, 448, 563
Caon N., Capaccioli M., D'Onofrio M., 1993, MNRAS, 265, 1013
Cole S. et al. (The 2dFGRS Team), 2001, astro-ph/0012429 (submitted to MNRAS)
Cox A. N., 2000, Allen's astrophysical quantities (4th edition; New York, Springer)
de Jong R. S., 1996, A&AS, 118, 557
de Vaucouleurs G., 1959, in Flugge S., ed., Handbuch der Physik 53, Springer-Verlag, Berlin,
p. 275
Dressler A., ApJ, 236, 351
Efstathiou G., Ellis R. S., Peterson B. A., 1988, MNRAS, 232, 431 (EEP)
Eggen O. J., Lynden-Bell D., Sandage A. R., 1962, ApJ, 136, 748
Fukugita M., Hogan C. J., Peebles P. J. E., 1998, ApJ, 503, 518
Gardner J. P., Sharples R. M., Frenk C. S., Carrasco B. E., 1997, ApJ, L99
Hoyle F., 1949, in Burgers J. M., van de Hulst H. C., eds., Problems of Cosmical Aerody-
namics, Central Air Documents Office, Dayton, p. 195
Hubble E., 1926, ApJ, 64, 321
Jimenez R., Friaca A. C. S., Dunlop J. S., Terlevich R. J., Peacock J. A., Nolan L. A., 1999,
MNRAS, 305, 16
Kauffmann G., White S. D. M., Guiderdoni B., 1993, MNRAS, 264, 201
-- 22 --
Kauffmann G., 1995, MNRAS, 274, 161
Kauffmann G., 1996, MNRAS, 281, 487
Kennicutt R. C., 1983, ApJ, 272, 54
Loveday J., Peterson B. A., Efstathiou G., Maddox S. J., 1992, ApJ, 390, 338
Menanteau F., Ellis R. S., Abraham R. G., Barger A. J., Cowie L. L., 1999, MNRAS, 309,
208
Naim A., Lahav O., Buta R. J., Corwin H. G., de Vaucoulers G., Dressler A., Huchra J. P.,
van den Bergh S., Raychaudhury S., Sodr´e L., Storrie-Lombardi M. C., 1995, MNRAS,
274, 1107
Navarro J. F., Frenk C. S., White S. D. M., 1995, MNRAS, 275, 56
Navarro J. F., Steinmetz M., 1997, ApJ, 478, 13
Navarro J. F., Steinmetz M., 2000, ApJ, 538, 477
Peebles P. J. E., 1969, ApJ, 155, 393
Sandage A., Tammann G. A., Yahil A., 1979, ApJ, 232, 352
Salpeter E. E., 1955, ApJ, 121, 61
Schechter P., 1976, ApJ, 203, 557
Schechter P. L., Dressler A., 1987, AJ, 94, 563
Somerville R. S., Primack J. R., Faber S. M., 2001, MNRAS, 320, 504
Toomre A., 1977, in Tinsley B. M., Larson R. B., eds., The Evolution of Galaxies and Stellar
Populations. Yale Univ. Press, New Haven, p. 401
van den Bosch F. C., Burkert A., Swaters R. A., 2001, astro-ph/0105158
Wadadekar Y., Robbason B., Kembhavi A., 1999, AJ, 117, 1219
This preprint was prepared with the AAS LATEX macros v5.0.
-- 23 --
Appendix A. Estimators for Spheroid and Disk Luminosity Functions
The traditional 1/Vmax estimator is trivially adapted to the case of disk and spheroid
luminosity functions. The estimator is applied just as in the case of the standard luminosity
function, except that the total luminosity of the galaxy (i.e. disk plus spheroid luminosity)
is used to compute Vmax, since it is this total luminosity that determines the volume within
which the galaxy could have been detected.
The maximum likelihood estimator of Efstathiou, Ellis & Peterson (1988, hereafter EEP)
is also easily generalized to the case of spheroid and disk luminosity functions. Consider the
case of the spheroid luminosity function (the same arguments apply to disks). As noted
earlier, the detectability of a spheroid depends upon both its absolute magnitude, M, and
on the bulge-to-total ratio which we denote by B in this Appendix. We begin therefore by
defining a two-dimensional function, Φ(M, B), such that Φ(M, B)dMdB is the number of
galaxies with bulge-to-total ratio in the range B to B + dB and spheroid absolute magnitude
M to M + dM per unit volume. The normal luminosity function of spheroids is easily recov-
0 Φ(M, B)dB. The probability that galaxy i with spheroid magnitude
Mi and bulge-to-total ratio Bi is seen in a magnitude limited survey is
ered using φ(M) = R 1
pi ∝ Φ(Mi, Bi),Z 1
0 Z M ′
−∞
lim(zi,B)
Φ(M, B)dMdB ,
(4)
where M ′
lim(zi, B) = Mlim(zi) − 2.5 log10 B and Mlim(zi) is the limiting absolute magnitude
of the survey at redshift zi. The use of M ′
lim is necessary since arbitrarily faint spheroids
will make it into the survey provided that they have a sufficiently low bulge-to-total ratio
(corresponding to sufficiently bright disks).
From this definition we can construct the usual likelihood function
ln L =
N
Xi=1
ln Φ(Mi, Bi) −
N
Xi=1
lim(zi,B)
ln(Z 1
0 Z M ′
−∞
Φ(M, B)dMdB) + const,
(5)
where N is the total number of galaxies. There are now two ways to proceed. In the first
we assume a simple parametric form for Φ(M, B) and maximize the likelihood with respect
to the parameters. This is analogous to fitting a Schechter function (Schechter 1976) to the
normal luminosity function (e.g. Sandage, Tammann & Yahil 1979). A simple parametric
form which we have tried to fit our data is
Φ(M, B) = φ(M) exp(βB),
(6)
where φ(M) is the usual Schechter function and β is a parameter to be estimated from the
fit. With this method, the likelihood function of eqn. (5) can be evaluated for each value
-- 24 --
of the three parameters α, β and M⋆, and hence the parameter values which maximize the
likelihood are readily obtained.
The second approach involves splitting Φ(M, B) into bins in M and B and treating each
as a parameter. This is equivalent to the SWML method of EEP for estimating the standard
luminosity function.
We represent Φ(M, B) as follows:
Φ(M, B) = Φk,h,(cid:26) Mk − ∆M/2 < M < Mk + ∆M/2, k = 1, . . . , Np
Bh − ∆B/2 < B < Bh + ∆B/2,
h = 1, . . . , Nq.
(7)
The likelihood function may then be written as
N
Xi=1
ln L =
where
W (Mi−Mk, Bi−Bh) ln Φk,h−
Xi=1
N
ln( Nq
Xh=1
Np
Xk=1
Φk,h∆M∆BH[Mk, Bh, Mlim(zi)])+const,
(8)
Wk,h(Mi, Bi) =(cid:26) 1 if Mk − ∆M/2 < Mi < Mk + ∆M/2, and Bh − ∆B/2 < Bi < Bh + ∆B/2,
0 otherwise
(9)
and
H[Mk, Bh, Mlim(zi)] =
1
∆M∆B Z Bh+∆B/2
Bh−∆B/2 Z Mk+∆M/2
Mk−∆M/2
Q(M, B)dMdB,
(10)
where Q(M, B) = 0 if M > Mlim(zi)−2.5 log10 B and Q(M, B) = 1 otherwise. Since only the
shape of the luminosity function is constrained by the above likelihood function, we introduce
an additional constraint, g =PkPh Φk,h(Lk,h/Lf)β∆M∆B − 1 = 0, where Lk,h is the total
luminosity of a galaxy with spheroid magnitude Mk and bulge-to-total ratio Bh and Lf is a
fiducial luminosity (which we will take to be that corresponding to MI − 5 log h = −20.5),
using a Lagrangian multiplier λ as did EEP. Maximizing ln L′ = ln L + λg then yields
Φk,h =
i=1 H[Mk, Bh, Mlim(zi)]/PNp
PN
l=1PNq
PN
i=1 W (Mi − Mk, Bi − Bk)
m=1 Φl,mH[Ml, Bm, Mlim(zi)]
(11)
which are easily solved with an iterative procedure. The covariance matrix for the parameters
is obtained in a manner entirely analogous to that outlined by EEP.
Normalization of the maximum likelihood luminosity function can be achieved using the
actual redshift data as described by Loveday et al. (1992), but using M ′
lim in the selection
function to account for the effects of the bulge-to-total ratio. A better approach is to nor-
malize by performing a least squares fit to the number counts of galaxies from a wide area
-- 25 --
survey. The cumulative number count to apparent magnitude m is given by
0 Z 1
n(m) =Z ∞
0 Z M ′
−∞
lim
Φ(m − D(z) − K(z) − 2.5 log10 B, B)
dV
dz
dMdBdz,
(12)
where D(z) and K(z) are the distance modulus and k+e correction respectively at redshift
z, which we can compute from the SWML estimate of Φ.
|
astro-ph/0412210 | 1 | 0412 | 2004-12-09T13:46:30 | Generation of Gravitational Radiation in Dusty Plasmas and Supernovae | [
"astro-ph",
"gr-qc"
] | We present a novel nonlinear mechanism for exciting a gravitational radiation pulse (or a gravitational wave) by dust magnetohydrodynamic (DMHD) waves in dusty astrophysical plasmas. We derive the relevant equations governing the dynamics of nonlinearly coupled DMHD waves and a gravitational wave (GW). The system of equations is used to investigate the generation of a GW by compressional Alfv\'{e}n waves in a type II supernova. The growth rate of our nonlinear process is estimated, and the results are discussed in the context of the gravitational radiation accompanying supernova explosions. | astro-ph | astro-ph |
Generation of Gravitational Radiation in Dusty Plasmas and Supernovae
Gert Brodin,1, ∗ Mattias Marklund,1, ∗ and Padma K. Shukla1, 2, ∗
1Department of Physics, Umea University, SE -- 901 87 Umea, Sweden
2Institut fur Theoretische Physik IV, Fakultat fur Physik und Astronomie,
Ruhr-Universitat Bochum, D -- 44780 Bochum, Germany
(Dated: February 4, 2019)
We present a novel nonlinear mechanism for exciting a gravitational radiation pulse (or a gravi-
tational wave) by dust magnetohydrodynamic (DMHD) waves in dusty astrophysical plasmas. We
derive the relevant equations governing the dynamics of nonlinearly coupled DMHD waves and a
gravitational wave (GW). The system of equations is used to investigate the generation of a GW
by compressional Alfv´en waves in a type II supernova. The growth rate of our nonlinear process is
estimated, and the results are discussed in the context of the gravitational radiation accompanying
supernova explosions.
PACS numbers: 04.30.Db, 97.60.Bw, 98.62.En
It is well known that there exist numerous mechanisms
for the conversion between gravitational waves (GWs)
and electromagnetic waves [1, 2, 3, 4, 5, 6, 7, 8, 9, 10, 11,
12, 13, 14, 15, 16, 17, 18, 19]. For example, the propaga-
tion of GWs across an external magnetic field gives rise to
a linear coupling to the electromagnetic field [1] , which
may lead to the gravitational wave excitation of ordinary
electromagnetic waves in vacuum, or of magnetohydrody-
namic (MHD) waves in a plasma [2, 3, 4]. Furthermore,
various nonlinear coupling mechanisms give rise to three-
wave couplings between GWs and electromagnetic waves
in matters. We also note that four-wave processes may
cause graviton-photon conversion even in the absence of
external matters or fields [5]. Moreover, GWs can cou-
ple to other types of waves, e.g. sound waves, also in
neutral media [6]. There are numerous motives for con-
sidering wave couplings involving GWs. In some cases,
the emphasis is on the basic theory [5, 6, 7, 8, 9, 10]. In
other works, the focus is on GW detectors [11, 12, 13],
on cosmology [14, 15, 16], or on astrophysical applica-
tions such as binary mergers [17], gamma ray bursts [18]
or pulsars [19]. Many of the previous works have con-
centrated on the conversion from GWs to electromag-
netic waves, which can be analysed within a test matter
approach which neglects the back reaction on the grav-
itational field. We note that such an approach can be
justified if the background energy density is low.
In this Letter, we consider the three-wave coupling be-
tween two DMHD waves and a GW, including the ef-
fects of dust particles [20] in a dense medium such as
the supernova where electrons, protons, and charged dust
macroparticles are abundant. For this purpose, we derive
the dust Hall MHD equations [20], i.e. equations describ-
ing the dust MHD waves, including the effect of a GW.
We emphasize that for a low-beta plasma, the system
of equations has a structure which can describe both a
dust-dominated plasma, as well as an ordinary Hall-MHD
plasma (if we replace the dust mass density by the ion
mass density). Using the normal mode approach [21],
the three-wave coupling equations are derived, includ-
ing the back reaction on a GW from the Einstein equa-
tions. The system is shown to fulfil the Manley-Rowe
relations [21] (which means that the interaction process
can be viewed quantum mechanically) and to be energy
conserving. The three-wave equations are then used to
analyse the generation of a GW by the compressional
Alfv´en waves in the iron core of the type II supernova
[22, 23]. It turns out that the characteristic timescale for
the Alfv´en-GW conversion can be less than a millisecond,
which implies that the mechanism is potentially relevant
for the high-frequency part (> 1MHz) of the supernova
GW spectrum.
The plasma dynamics, due to the response to a gravi-
tational wave
ds2 = −dt2+(1 + h+) dx2+(1 − h+) dy2+2h×dx dy+dz2
(1)
can be formulated according to
∂tns + ∇ · (nsvs) = 0,
(2)
and
msns(∂t + v · ∇)v = −∇ps + qsns(E + v × B) + msnsgs,
(3)
e1 =
(1 + h+/2) ∂y −
where we have introduced the tetrad e0 = ∂t,
(1 − h+/2) ∂x − h×/2∂y, e2
h×/2∂x, e3 = ∂z, and ∇ = (e1, e2, e3). Moreover,
=
gs = − 1
− 1
− 1
2(cid:2)(v2
2 (1 − vsz)(vsx∂th+ + vsy∂th×)e1
2 (1 − vsz)(vsx∂th× − vsy∂th+)e2
sx − v2
sy)∂th+ + 2vsxvsy∂th×(cid:3) e3
(4)
represents the gravitational acceleration of the particle
species s due to the GWs. We have assumed that ∂z ≈
−∂t holds for the GWs.
The electromagnetic field is determined through the
gravity modified Maxwell's equations. Using the same
notation as above, they take the form
Again using Eqs. (10) and (11) to eliminate the electric
2
∂tE = ∇ × B −Xs
∂tB = −∇ × E − jB,
qsnsvs − jE,
(5)
(6)
with the constraints ∇ · E = Ps qsns and ∇ · B = 0.
Here, the effects of the GWs (1) are represented by the
effective currents
jE = − 1
− 1
2 [(Ex − By)∂th+ + (Ey + Bx)∂th×] e1
2 [−(Ey + Bx)∂th+ − (Ex − By)∂th×] e2, (7)
and
jB = − 1
− 1
2 [(Ey + Bx)∂th+ − (Ex − By)∂th×] e1
2 [(Ex − By)∂th+ + (Ey + Bx)∂th×] e2. (8)
With the general setting established above, we will
from now on focus on the case of a three-component
dusty plasma, for which we have the equation of state
ps = kBTsns. Thus, the plasma is composed of electrons
(e), ions (i), and dust particles (d). The mass md of
the dust particles is assumed to be much larger than the
electron and ion masses, viz. me and mi, respectively.
We will assume that the plasma is approximately quasi-
neutral, i.e. qini = ene − qdnd. Moreover, the waves un-
der consideration are supposed to propagate with phase
velocities much smaller than the speed of light c. Thus,
we may neglect the displacement current in Amp´ere's law
(5), i.e.
∇ × B =Xs
qsnsvs + jE.
(9)
Due to the constraint me, mi ≪ md the momentum con-
servation equation (3) for the inertialess electrons and
ions becomes
0 = −kBTe∇ne − ene(E + ve × B) + menege,
(10)
and
0 = −kBTi∇ni + qini(E + vi × B) + minigi,
(11)
respectively. Adding Eqs. (10) and (11), using the
quasineutrality condition, assuming that the number
densities of the electrons and ions are not much larger
than the number density of the dust, and using the heavy
dust approximation, the dust momentum equation takes
the form
ρd (∂t + vd · ∇) vd = −kB(cid:18)Td −
qd
qi
+(∇ × B) × B − jE × B + ρdgd
Ti(cid:19) ∇nd
(12)
field, Faraday's law (6) becomes
md
qd
∂tB = ∇×(vd × B)−
∇×[(∂t + vd · ∇) vd − gd]− jB,
(13)
where we have used the dust momentum equation (12).
Thus, Eqs. (12) and (13) together with the dust conti-
nuity equation
∂tρd + ∇ · (ρdvd) = 0,
(14)
constitute the dust MHD equations in the presence of a
GW. For a low-beta plasma, the pressure term in (12) is
negligible, which means that the structure of Equations
(12)-(14) is the same as in an ordinary Hall-MHD plasma
without dust. Henceforth, we will consider a low-beta
plasma, drop the index d on all quantities and thus let
q/m be either the ion charge to mass ratio or the consid-
erably smaller charge to mass ratio of the dust particles.
As a result, our mathematical analysis below will then
apply either to a dust Hall-MHD plasma, or to an ordi-
nary Hall-MHD plasma without dust.
To simplify the problem, we consider the case when
the dust-acoustic speed cs = kBT /m is much smaller
such
that the pressure term in (12) can be neglected. As a
prerequisite for the nonlinear calculations, we first study
the linear modes of the system (12)-(14) omitting the
than the dust Alfv´en velocity CA = (cid:0)B2
0/µ0ρ(cid:1)1/2
gravitational contributions. Letting B = B0bz + B1, ρ =
ρ0 + ρ1, where the index 1 denotes the perturbation of
the equilibrium part, and linearizing Eqs. (12)-(14) and
Fourier analysing, we readily obtain the dispersion rela-
tion
(cid:0)ω2 − k2
z C 2
A(cid:1) −
A(cid:1)(cid:0)ω2 − k2C 2
ω2k2
z k2C 2
A
ω2
c
= 0,
(15)
where ωc = qB0/m is the gyrofrequency. For frequencies
much smaller than the gyrofrequency, we note that the
z C 2
modes separate into the shear Alfv´en wave, ω2 − k2
A ≈
0, and the compressional Alfv´en wave, ω2 − k2C 2
A ≈ 0.
Below we will consider the more general case described
by (15), however. For later applications it is convenient
to use the linear equations to express all quantities in
terms of a single variable. Thus, we let the wavevector
of the dust MHD waves lie in the x − z-plane, and write
vy = i
ω
ωc
k2
z C 2
A
z C 2
(ω2 − k2
A)
vx, vz = 0, ρ1 = ρ0
kxvx
ω
(16)
Bx = −B0
vx
By = −iB0
kz
(ω2 − k2
z C 2
A)
where ρd = mdnd.
approximation [Te + (e/qi)Ti]ne ≪ [Td − (qd/qi)Ti]nd.
In Eq. (12) we have also used the
Bz = B0
ωkz
k2C 2
A
ω2
ωc
ωkx
k2C 2
A
vx.
vx
(17)
(18)
(19)
Next, we consider a system of three weakly inter-
acting waves. Two dust MHD waves with frequencies
and wave-numbers (ω1, k1) and (ω2, k2) respectively, and
an arbitrarily polarized gravitational wave propagating
along the z-direction with the frequency and wavenum-
ber (ωg, kgbz). Noting that the gravitational dispersion
relation reads ωg = kgc and that CA ≪ c, the frequency
and wavenumber matchings (energy and momentum con-
servation) can be approximated
ωg = ω1 + ω2, kg = k1 + k2 ⇒ 0 ≈ k1 + k2.
(20)
time, i.e. ρ = ρ0 +P2
We will thus use k1 ≈ −k2 below, and we define
k1x = −k2x ≡ kx as well as k1z = −k2z ≡ kz (let-
ting ky = 0 for convenience). All quantities are now
assumed to be superpositions of two dust MHD waves
waves whose amplitudes are weakly varying functions of
j=1 ρj(t) exp[i(kj · r − ωt)] + c.c.,
where c.c stands for the complex conjugate, In princi-
ple, the gravitational wave should also contribute, but we
note that within a fluid model the gravitational wave con-
tribution to all plasma perturbations (velocity, magnetic
field and density) are second order in the gravitational
wave amplitudes, provided that the GW propagates par-
allel to B0, as we have assumed. Thus, the only linear
perturbations due to the gravitational wave are those of
the metric as described by Eq. (1).
Next, in order to simplify the algebra, we introduce
where
3
ω2
1,2
k2
1,2C 2
A
z1,2C 2
A
ρ0
W1,2 =
2"1 +
c(cid:0)ω2
V× = i(cid:20) ω1k2
1,2k4
1,2 − k2
z C 2
ωc (ω2
ω2
ω2
+
z1,2C 2
Ak2 − ω2ω2
1 − k2
z1C 2
and
V+ = i"1 +
ω1ω2
ω2
c
k2
z C 2
(ω2
1 − k2
A(cid:0)k2
z C 2
1,2
1k2
z
ω2
z1,2C 2
k2
A(cid:1)2 1 +
A) k2 + 1 ↔ 2(cid:21) ,
A + ω1ω2(cid:1)
2 − k2
z C 2
A)
z C 2
A) (ω2
+
A!# (25)
(26)
ω1ω2k2
z
k4C 2
A # .
(27)
Next using the Einstein equations, linearized in h+, h×,
keeping only the resonantly varying part of T µν we obtain
for the × and +-polarization, respectively
iωg
dh×
dt
= κ(cid:20)ρ0vx1vy2 +
2(cid:20)ρ0 (vx1vx2 − vy1vy2) +
κ
Bx2By1
µ0
(cid:21) ,
(28)
Bx1Bx2 − By1By2
µ0
(cid:21) ,
(29)
and
iωg
dh×
dt
=
which is reduced to
vyj −
ω2
j
k2
j
kzj
kxjB0
Bxj
and
dh×
dt
= −
ρ0ωg
Wg
V×vx1vx2,
(30)
the normal mode aj defined by
zjC 2
A
zj C 2
ω2
j k2
j − k2
aj =
vxj − i
ωj
kxj
zj C 2
j k2
ω2
A
zjC 2
j − k2
j C 2
j − k2
ωc(cid:0)ω2
A(cid:1)
A(cid:1) ω2
k2
j ωj
Bz
C 2
A
B0
c − k2
+
ω2
c(cid:0)ω2
+i(cid:0)ω2
j C 2
Aω2
j
A(cid:1) kxj
kzj
ωckxjB0
By.
(21)
Returning to Eqs.
(12)-(14) and including the nonlin-
ear terms [26], we can, by keeping the part varying as
exp[i(k1 · r − ω1t)], derive
2xh+ + v∗
∂a1
∂t
ω2
1
2k2
kz
ω1
= −
2kx(cid:0)v∗
kxB0(cid:0)B ∗
kz(cid:2)(cid:0)ω2
1 − k2C 2
−
+i
2xh+ − B ∗
A(cid:1) ω2
k2ω1ωckxB0
2yh×(cid:1) − i
2yh×(cid:1)
1(cid:3)
Aω2
c − k2C 2
ω2
1k2
z C 2
2ωc (ω2
A(cid:0)v∗
2xh× − v∗
1 − k2
z C 2
A) kx
2yh+(cid:1)
(cid:0)B ∗
2xh× − B ∗
(22)
2yh+(cid:1) ,
and we obtain a similar result for ∂a2/∂t by letting 1 ↔ 2.
After some algebra, using Eqs. (16)-(19) and (21), we
find that Eq. (22) reduces to
dvx1
dt
=
ρ0ω1
W1
v∗
x2 (V+h+ + V×h×) ,
(23)
and similarly for mode 2
dvx2
dt
=
ρ0ω2
W2
v∗
x1 (V+h+ + V×h×) ,
(24)
dh+
dt
= −
ρ0ωg
Wg
V+vx1vx2,
(31)
where Wg = ω2
g/2κ. The total wave energy is Wtot =
W1 vx12 + W2 vx22 + Wg(h×2 + h+2) and it is eas-
ily verified from (23)-(24) together with (30)-(31) that
Wtot is conserved. Furthermore, the appearance of the
same coupling coefficients V×, V+ in Eqs.
(23)-(24)
as well as in (30)-(31) assures that the Manley-Rowe
relations are fulfilled, which implies that each mode
changes energy in direct proportion to its frequency, i.e.
(dW1/dt)/(dW2/dt) = ω1/ω2 etc. The system of (23)-
(24) together with (30)-(31) describing the energy conver-
sion between DMHD waves and GWs is one of the main
results of the present letter. A more elaborate calcula-
tion scheme, including effects such as inhomogeneity and
background curvature, is a project for future research.
We now apply our results to the gravitational radia-
tion arising from the iron core of a type II supernova
where the densities can be of the order 1017 kg/m3 [23].
The large neutrino outflow (which can reach powers of
1033 W/cm2 see e.g. Ref. [22]) can generate MHD waves
described by (15). From the flux conservation, we expect
the iron core to be strongly magnetized (comparable to
pulsars), and for magnetic field strengths B0 ∼ 108 T,
the gyrofrequency will be much larger than all other fre-
quencies of the problem. The dispersion relation then
separate into the shear Alfv´en waves ω2 − k2
A ≈ 0 and
the compressional Alfv´en waves ω2 − k2C 2
A ≈ 0. Assum-
ing that the pump MHD wave is a compressional mode
with a frequency 5 MHz, the matching conditions (20)
can be fulfilled for a GW with a typical frequency 3MHz
and a shear Alfv ´en wave with the frequency 2 MHz [27].
For the assumed geometry, the MHD waves couple only
to the h×-polarization (to a good approximation), and
by combining (24) and (30), we obtain
z C 2
d2h×
dt2 = −h×
ω2
ωg
V×2
W2
16πG
c2 ρ0 vx12 .
(32)
Thus, noting that the factor ω2 V×2 /ωgW2 is negative
[27] and has a magnitude of order unity [28] for the given
parameters, the growth rate is
γ ∼p16πGρ0
vx1
c
,
(33)
which for a weakly relativistic pump quiver speed,
vx1 /c ∼ 1/10, implies γ ∼ 10 kHz. Thus, we deduce
that excitation of a GW by MHD waves is a reasonably
fast process in a dense medium such as the supernova iron
core. Furthermore, the present process can give rise to
GWs of higher frequencies than many of the previously
considered excitation mechanisms, see e.g.
[24]. Thus,
our model contributes to the understanding of gravita-
tional radiation emissions accompanying supernova ex-
plosions [25].
∗ Also at: Centre for Fundamental Physics, Ruther-
ford Appleton Laboratory, Chilton Didcot, Oxfordshire,
OX11 0QX, UK
[1] L. P. Grishchuk and A. G. Polnarev, General Relativity
and Gravitation Vol. 2 ed. A. Held (Plenum Press, New
York, 1980) pp. 416-430.
[2] D. Papadopoulos et al., A & A 377, 701 (2001).
[3] J. Moortgat and J. Kuijpers, A & A 402, 905 (2003).
[4] J. Moortgat and J. Kuijpers, Phys. Rev. D, 70, 023001
(2004).
[5] A. Kallberg, G. Brodin, and M. Marklund, submitted
(2004) (gr-qc/0410005).
[6] A. M. Anile, J. K. Hunter and B. Turong, J. Math. Phys.
40, 4474 (1999)
4
[7] M. Servin and G. Brodin, Phys. Rev. D 68, 044017
(2003).
[8] M. Servin et al., Phys. Rev. D 67, 087501 (2003).
[9] J. T. Mendon¸ca, Plasma Phys. Control. Fus., 44, B225,
(2002).
[10] A. B. Balakin et al., J. Math. Phys., 44, 5120 (2003)
[11] F.Y. Li., M. X. Tang., Int. J. Mod. Phys. D, 11, 1049
(2002)
[12] G. Brodin and M. Marklund, Class. Quantum Grav. 20,
45 (2003).
[13] R. Ballantini et al., Class. Quantum Grav. 20, 3505
(2003).
[14] D. Papadopoulos, Class Quantum Grav. 19, 2939 (2002).
[15] M. Marklund et al., Phys. Rev. D 62, 101501 (2000).
[16] P. A. Hogan and E. M. O'Shea, Phys. Rev D 65, 124017
(2002).
[17] G. Brodin et al., Phys. Rev. D 63, 124003 (2001).
[18] L. Vlahos et al., Astrophys. J., 604, 297 (2004).
[19] H. J. M. Cuesta, Phys. Rev. D 65, 64009 (2002).
[20] P. K. Shukla and A. A. Mamun, Introduction to Dusty
Plasma Physics (Institute of Physics, Bristol, 2002).
[21] J. Weiland and H. Wilhelmsson, Coherent Nonlinear In-
teraction of Waves in Plasmas, (Pergamon Press, New
York, 1977).
[22] G.G. Raffelt, Stars as Laboratories for Fundamental
Physics (The University of Chicago Press, Chicago,
1996).
[23] S. E. Woosley et al., Rev. Mod. Phys. 74 1015 (2002).
[24] C. Cutler and K. Thorne, in Proceeding of the 16th Inter-
national Conference on General Relativity and Gravita-
tion, Durban, South Africa, 2001, eds. N.T. Bishop and
S.D. Maharaj (World Scientific, 2002).
[25] M. V. Sazhin et al., JETP Lett. 64, 871 (1996).
[26] When including the nonlinear GW-coupling,
it is in
principle important to separate the coordinate compo-
nents (indices x, y, z) from the tetrad components (in-
dices 1, 2, 3), since the difference is first order in h+, h×.
However, for notational convenience we let indices x, y, z
denote tetrad components 1, 2, 3 in equation (22) and
henceforth.
[27] Note that from the Manley-Rowe relations the pump
wave must have the highest frequency for the growth
rates to become real and positive, Thus using the match-
ing condition (20) we see that the shear Alfv´en mode
formally has a negative frequency. Furthermore, we note
that the dispersion relations together with the match-
ing conditions means that the pump mode cannot prop-
agate purely along the magnetic field. Specifically, the
wavevector matching implies kxCA ≈ 4.6MHZ, where
CA ≈ 300m/s for the given parameters.
[28] This can be seen by taking the limit ωc → ∞ of the
c from
2 − k2
expression, noting that(cid:0)ω2
the dispersion relation (15).
z2C 2
A(cid:1) scales as 1/ω2
|
astro-ph/0202343 | 1 | 0202 | 2002-02-19T00:19:44 | Atomic Processes in Planetary Nebulae | [
"astro-ph"
] | A hot central star illuminating the surrounding ionized H II region usually produces very rich atomic spectra resulting from basic atomic processes: photoionization, electron-ion recombination, bound-bound radiative transitions, and collisional excitation of ions. Precise diagnostics of nebular spectra depend on accurate atomic parameters for these processes. Latest developments in theoretical computations are described, especially under two international collaborations known as the Opacity Project (OP) and the Iron Project (IP), that have yielded accurate and large-scale data for photoionization cross sections, transition probabilities, and collision strengths for electron impact excitation of most astrophysically abundant ions. As an extension of the two projects, a self-consistent and unified theoretical treatment of photoionization and electron-ion recombination has been developed where both the radiative and the dielectronic recombination processes are considered in an unified manner. Results from the Ohio State atomic-astrophysics group, and from the OP and IP collaborations, are presented. A description of the electronic web-interactive database, TIPTOPBASE, with the OP and the IP data, and a compilation of recommended data for effective collision strengths, is given. | astro-ph | astro-ph |
**TITLE**
ASP Conference Series, Vol. **VOLUME**, **PUBLICATION YEAR**
**EDITORS**
Atomic Processes in Planetary Nebulae
Sultana N. Nahar
Department of Astronomy, The Ohio State University, Columbus, OH
43210, USA
Abstract. A hot central star illuminating the surrounding ionized H
II region usually produces very rich atomic spectra resulting from basic
atomic processes: photoionization, electron-ion recombination, bound-
bound radiative transitions, and collisional excitation of ions. Precise
diagnostics of nebular spectra depend on accurate atomic parameters
for these processes. Latest developments in theoretical computations
are described, especially under two international collaborations known
as the Opacity Project (OP) and the Iron Project (IP), that have yielded
accurate and large-scale data for photoionization cross sections, transi-
tion probabilities, and collision strengths for electron impact excitation of
most astrophysically abundant ions. As an extension of the two projects,
a self-consistent and unified theoretical treatment of photoionization and
electron-ion recombination has been developed where both the radiative
and the dielectronic recombination processes are considered in an uni-
fied manner. Results from the Ohio State atomic-astrophysics group,
and from the OP and IP collaborations, are presented. A description
of the electronic web-interactive database, TIPTOPBASE, with the OP
and the IP data, and a compilation of recommended data for effective
collision strengths, is given.
1.
Introduction
A planetary nebula may be thought of as an 'astrophysical laboratory' of atomic
emission-line spectra. The spectra enable determination of its temperature,
density, and abundances of elements. The precise spectral analyis require accu-
rate atomic parameters for the radiative are collisional processes in the nebular
plasma. The four basic atomic processes that dominant the nebular plasma are:
i) Radiative bound-bound transition for excitation or de-excitation:
X +Z + hν ⇀↽ X +Z ∗,
where the ion, X, is of charge, Z.
ii) Photoionization (PI) by absorption of a photon:
X +Z + hν ⇀↽ X +Z+1 + ǫ,
The inverse process is the electron-ion radiative recombination (RR).
iii) Autoionization (AI) and dielectronic recombination (DR):
1
2
Author & Co-author
e + X +Z → (X +Z−1)∗∗ → (cid:26) e + X +Z
AI
X +Z−1 + hν DR
where the intermediate doubly excited state, autoionizing state, introduces res-
onances in the atomic processes. The inverse process of DR is photoionization
via the autoionizing state.
iv) The collisional process of electron-impact excitation (EIE)
e + X +Z → e′ + X +Z ∗∗
is one of the primary processes for spectral formation in astrophysical plasmas.
Each process needs to be treated separately to obtain the relevant atomic
parameters. Sample results for each process, obtained mainly under the OP
(1995, 1996) and the IP (1993) for accurate atomic data and stellar opacities
will be presented later. Astrophysical model applications, such as for plasma
opacities, need a huge amount of atomic data as these deal with a large number
of atomic levels over wide energy ranges. Consistent sets of atomic parameters
can be obtained if the atom or the ion is described by the same wavefunction
expansion in each process. This also reduces the uncertainty in applications
involving different processes and approximations. The Close Coupling (CC) R-
matrix mehtodology employed by the OP and IP enable the computation of such
self-consistent sets of atomic parameters.
The large amount of atomic data and opacities obtained under the OP
and the IP are available electronically through the existing database, TOPbase
(Cunto et al. 1993), and through its planned extension TIPTOPBASE (C.
Mendoza and the OP/IP team).
2. Theory
All calculations for the various atomic parameters of the dominant atomic pro-
cesses are carried out using the accurate and powerful R-matrix method in the
close-coupling approximation (e.g. Burke & Robb 1975, Seaton 1987, Berring-
ton et al. 1987, Berrington et al. 1995). The total wavefunction for a (N+1)
electron system in the CC approximation is described as:
ΨE(e + ion) = A
N
Xi
χi(ion)θi +Xj
cjΦj(e + ion),
(1)
where χi is the target ion or core wavefunction in a specific state SiLiπi or level
Jiπi, θi is the wavefunction of the interacting (N+1)th electron in a channel
labeled as SiLi(Ji)πi k2
i is the incident kinetic energy. Φj is
the correlation functions of (e+ion) system that compensates the orthogonality
condition and short range correlation interations. The complex resonant struc-
tures in photoionization, recombination, and in electron impact excitation are
included through channel couplings.
i ℓi(SLπ or Jπ); k2
Relativistic effects are included through Breit-Pauli approximation in inter-
mediate coupling. The (N+1)-electron Hamiltonian in the Breit-Pauli approxi-
mation, as adopted in the Iron Project, is
APS Conf. Ser. Style
H BP
N +1 = H N R
N +1 + H mass
N +1 + H Dar
N +1 + H so
N +1,
where non-relativisitc Hamiltonian is
H N R
N +1 =
N +1
Xi=1
−∇2
i −
2Z
ri
+
N +1
Xj>i
.
2
rij
3
(2)
(3)
N +1 is the mass correction term, H Dar
H mass
N +1 is the
spin-orbit interaction term. Spin-orbit interaction splits the LS terms into fine-
structure levels labeled by Jπ where J is the total angular momentum. Solutions
of the Schrodinger equation, H BP
N +1Ψ = EΨ which becomes a set of coupled equa-
tions with the CC expansion, give the bound wavefunctions, ΨB, for negative
energies, E < 0, and continuun wavefunction, ΨF , for positive energies, E ≥ 0.
N +1 is the Darwin term, H so
The transition matrix elements for various atomic processes are:
< ΨBDΨF > for photoionization and recombination,
< ΨBDΨB ′ > for oscillator strength,
< ΨF H(e + ion)ΨF ′ > for electron impact excitation,
where D is the dipole operator.
The transition matrix element with the dipole operator can be reduced to
the generalized line strength defined, in either length or velocity form, as
*Ψf
N +1
Xj=1
SL = (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
2
rjΨi+(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
, SV = ω−2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
2
.
(4)
*Ψf
N +1
Xj=1
∂
∂rj
Ψi+(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
where ω is the incident photon energy in Rydberg units, and Ψi and Ψf are the
wave functions representing the initial and final states, respectively.
The oscillator strength fij and the transition probability Aji for the bound-
bound transition can be obtained from S as
fij =
Eji
3gi
S, Aji(a.u.) =
1
2
α3 gi
gj
E2
jifij,
(5)
where Eji is the transition energy, α is the fine structure constant, and gi, gj
are the statistical weight factors of the initial and final states. The lifetime,
τj, of a level j can be obtained from the transition probabilities decaying to
and τ0 =
the lower levels as τj = (Pi Aji(s−1))−1, where Aji(s−1) = Aji(a.u.)
2.4191 × 10−17s.
The photoionization cross section (σP I ) is proportional to the generalized
τ0
line strength (S),
σP I =
4π
3c
1
gi
ωS.
(6)
The calculations of total electron-ion recombination employs a new unified
treatment (Nahar & Pradhan 1994,1995). The method considers the infinite
number of recombining states and incorporates the radiative and dielectronic
recombinations, RR and DR, in a unified manner. The contributions to re-
combination from states with n ≤ 10 are obtained from σP I using principle of
4
Author & Co-author
detailed balance, while the resonant contributions from states with 10 < n ≤ ∞
are obtained from an extension of the DR theory of Bell & Seaton (1985).
The recombination cross section, σRC, is related to σP I through the princi-
ple of detailed balance,
σRC = σP I
gi
gj
h2ω2
4π2m2c2v2 .
The recombination rate coefficient, αRC , is obtained as
αRC (T ) = Z ∞
0
vf (v)σRC dv,
(7)
(8)
where f (v) is the Maxwellian velocity distribution function. The total αRC is
obtained from contributions from infinite number of recombined states.
The collision strength for transition by electron impact excitation from the
initial state of the target ion SiLi to the final state SjLj is given by
Ω(SiLi − SjLj) =
1
2 XSLπXlilj
(2S + 1)(2L + 1)SSLπ(SiLili − SjLjlj)2,
(9)
where S is the scattering matrix. The effective collision strength or the Maxwellian
averaged collision strength can be obtained as
Υ(T ) = Z ∞
0
Ωij(ǫj)e
−ǫj
kT d(ǫj/kT ).
(10)
which give the excitation rate coefficient, qij(T ) = (8.63×10−6/ωiT 1/2)e−Eij /kT Υ(T )
in cm3s−1. where T is in K, Eij = Ej − Ei, Ei < Ej are in Rydbergs (1/kT =
157885/T), and j is the excited upper state.
3. Results and Discussions
Sample results for each atomic process are presented, e.g., for f -values, σP I ,
αR(T ), and Υ(T ), in the following subsections.
3.1. Bound-Bound transitions
Due to fine structure, extensive sets of data for f - and A-values can be obtained
for a large number of bound-bound transitions using the Breit-Pauli R-matrix
method. In contrast to dipole allowed LS multiplets, BPRM method includes
relativistic effects, and consideration of both the dipole allowed and intercombi-
nation transitions. The fine structure energy levels are analysed with quantum
defects to obtain the spectroscopic identification, (Ct St Lt Jt πtnℓ)SLJ π where
(Ct St Lt Jt πt) denotes the core configuration, spin, orbital, total angular mo-
menta and parity, nl is configuration of the outer or valence electron, SL are
the possible spin and orbital angular momenta, and Jπ are the total angular
momentum and parity of the (N+1)-electron system. The fobidden quadrupole
(E2) and magnetic dipole (M1) transitions are treated with atomic structure
calculations using codes such as SUPERSTUCTURE (Eissner et al. 1974).
APS Conf. Ser. Style
5
Accurate fine structure transition probabilities have been obtained for a
number of ions such as:
Li-like: C IV, N V, O VI, F VII, Ne VIII, Na IX, Mg X, Al XI, Si XII, S XIV,
Ar XVI, Ca XVIII, Ti XX, Cr XXII, Ni XXVI
Fe ions: Fe V, Fe XVII, Fe XXI, Fe XXIII, Fe XXIV, Fe XXV
Other ions: C II, C III, O IV, S II, Ar XIII
The first large-scale application of the BPRM method for a complex ion,
Fe V, resulted in 3865 fine structure energy levels, compared to 179 observed, and
about 1.46 million f - and A−values for dipole allowed and intercombination (E1)
transitions and 362 forbidden (E2,M1) transitions (Nahar et al. 2000). Table
I presents a sample with complete spectroscopic identification of the transition
array 3d4(4D) − 3d3(4F )4p(5F o) in Fe V.
Table I: Transition probabilities of Fe V. g=2J+1.
leveli
3d4(5D)
3d4(5D)
3d4(5D)
3d4(5D)
3d4(5D)
3d4(5D)
3d4(5D)
3d4(5D)
3d4(5D)
3d4(5D)
3d4(5D)
3d4(5D)
levelj
gi
gj Ei(Ry) Ej(Ry)
f
3d3(4F )4p(5F o)
3d3(4F )4p(5F o)
3d3(4F )4p(5F o)
3d3(4F )4p(5F o)
3d3(4F )4p(5F o)
3d3(4F )4p(5F o)
3d3(4F )4p(5F o)
3d3(4F )4p(5F o)
3d3(4F )4p(5F o)
3d3(4F )4p(5F o)
3d3(4F )4p(5F o)
3d3(4F )4p(5F o)
1
3
3
5
5
5
7
7
9
7
9
9
3
3
5
3
5
7
5
7
7
9
9
11
5.5132
5.5119
5.5119
5.5094
5.5094
5.5094
5.5058
5.5058
5.5015
5.5058
5.5015
5.5015
3.1644
3.1644
3.1496
3.1644
3.1496
3.1443
3.1496
3.1443
3.1443
3.1391
3.1391
3.1343
2.154E-01
3.790E-04
1.358E-03
4.617E-02
5.967E-02
1.462E-02
6.895E-03
5.889E-02
1.966E-03
3.262E-02
5.139E-02
7.548E-02
A(sec−1)
3.18E+09
1.68E+07
3.65E+07
3.40E+09
2.67E+09
4.69E+08
4.30E+08
2.64E+09
1.13E+08
1.14E+09
2.30E+09
2.78E+09
3d4(5D)
3d3(4F )4p(5F o)
25
35
5.5055
3.1451
1.068E-01
3.42E+09
3.2. Photoionization and Recombination
The photoionization cross sections, σP I, are calculated including autoionizing
resonances that can enhance the background cross sections considerably. Fig. 1
shows the photoionization cross sections of the gound states of Fe I to Fe V
(Bautista & Pradhan 1998). Extensive resonances dominate the cross sections
for these complex ions. The enhancement in the backgound is up to three orders
of magnitude for Fe I, over an order or magnitude for Fe II, and ∼ 50% for Fe
III.
For simpler systems the accuracy of theoretical cross sections can be tested
against experimental data. Very precise measurements of σP I are now being
carried out at a few places, such as the Advanced Light Source at Berkeley (e.g.
Covington et al. 2001), the merged photon-ion beam set-up at Aahus University
(Kjeldsen et al. 1999), and a synchroton based experiment at University of Paris-
Sud (Wuilleumier et al., private communication). An example of the comparison
of ground state photoionization cross sections of C II, with very good agreement
6
Author & Co-author
Fe I
Fe IV
]
b
M
[
n
o
i
t
c
e
s
s
s
o
r
c
n
o
i
t
i
i
a
z
n
o
o
o
h
p
t
0
1
g
o
L
2
1
0
−1
2
1
0
−1
1
2
1
0
−1
2
.6
.8
1
1.2
Fe II
1.4
1.6
1.8
2
2
1.5
1
.5
0
−.5
3
2
1
4
6
8
Fe V
10
12
1.5
2
Fe III
2.5
3
0
4
5
6
7
8
9
10
11
12
2.2
2.4
2.6
2.8
3
Energy [Ry]
Figure 1.
of Fe I - Fe V.
Photoionization cross sections, σP I, of the ground states
)
b
M
(
I
P
σ
30
20
10
0
24
30
)
b
M
(
I
P
σ
20
10
0
24
C II + hν -> C III + e
(b) Expt
25
26
27
28
29
30
31
(b) Theory: BPRM
25
26
27
28
29
30
31
Photon Energy (eV)
Comparison of theoretical and experimental photoion-
Figure 2.
ization cross sections, σP I, of the ground state fine structure levels
2s22p(2P o
1/2,3/2) of C II (Nahar 2002).
with experiment, is shown in Fig. 2. Theoretical calculations include relativistic
fine structure.
An example of the total unified recombination rate coefficient αR(T ) for
O III is presented in Fig. 3 (solid) (Nahar 1998). The earlier results are RR rates
(dashed, Pequignot et al. 1991), low temperature DR rates (dotted, Nussbaumer
& Storey 1983), high temperature DR rates by Badnell & Pindzolla (1990, short-
dash-long-dash) and by Shull & Steenberg (1982, dot-desh). Differences can be
noticed between the present unified values and the sum of the earlier (RR+DR)
results, especially in the low and intermediate temperature. Compared to the
earlier Shull & Steenberg results, autoionization into excited levels in the high
temperature region reduces the recombination rates significantly.
The unified treatment for the total recombination provides rates that are
valid over a wide range of temperatures for all practical purposes in contrast to
addition of RR and DR rates obtained using different approximations for dif-
APS Conf. Ser. Style
7
e + O IV -> O III
10-9
10-10
10-11
10-12
)
1
-
c
e
s
3
m
c
(
R
α
10
100
1000
10000
T (K)
105
106
107
Figure 3.
cients (αR(T ) (solid) with the previous calculations.
Comparison of the unified total recombination rate coeffi-
ferent temperature ranges. The method enables obtaining self-consistent sets of
photoionization and recombination cross sections by using the identical wave-
function expansion for both the inverse processes. Following is the list of atoms
and ions for which self-consistent sets of σP I and αR(T ) have so far been obtained
for over 45 ions (e.g. Nahar & Pradhan 1997, http : //www.astronomy.ohio − state.edu/ ∼pradhan):
Carbon: C I, C II, C III, C IV, C V, C VI
Nitrogen: N I, N II, N II, N IV, N V, N VI, N VI
Oxygen: O I, O II, O III, O IV, O V, O VI, O VII, O VII
Iron: Fe I, Fe II, Fe III, Fe IV, Fe V, Fe XIII, Fe XVII, Fe XXI, Fe XXIV, Fe
XXV, Fe XXVI
C-like: F IV, Ne V, Na VI, Mg VII, Al VIII, Si IX, S XI, Ar XIII, Ca XV
Other ions: Si I, Si II, S II, S III, Ar V, Ca VII, Ni II
The data include the state specific recombination rates for hundreds of bound
levels with n≤ 10 for each ion.
The self-consistent sets of photoionization/recombination data include new
photoionization cross sections that are generally an improvement over the OP
data since more extensive and accurate eigenfunction expansions are employed.
3.3. Electron Impact Excitation
Reviews and compilations of available theoretical data sources are: An evaluated
compilation of theoretical data sources for electron-impact excitation of atomic
ions, (A.K. Pradhan & J.W. Gallagher, Atomic Data and Nuclear Data Tables,
52, 227, 1992), and Electron Collisions with Atomic Ions (A.K. Pradhan & H.L.
Zhang, in LAND OLT-BORNSTEIN Volume Atomic Collisions, Ed. Y. Itikawa,
Springer-Verlag, in press).
A table of recommended data for effective collision strengths and A-values
for nebular ions is available on-line from www.astronomy.ohio-state.edu/∼pradhan
The aim of the IP is to compute collisional data for the iron-peak elements
in various ionization stages. Of the 50 publications in the "Atomic data from
the IRON Project" series in Astronomy and Astrophysics journal, most are on
8
Author & Co-author
Figure 4.
Collision strength for the first transition in Fe VI
Ω(4F3/2 − 4F5/2) from the Iron Project CC R-matrix calculation (Chen
& Pradhan 1999); previous results (• - Nussbaumer & Storey 1978, ×
- Garstang et al. 1978) neglect resonances.
collisional strengths. At present the Iron Project website is maintained by Keith
Butler at: www.usm.uni − muenchen.de/people/ip/iron − project.html
An example of IP calculations for iron ions is the recent work on Fe VI (Chen
& Pradhan 1999). Fig. 4 presents the detailed collision strength Ω(E) showing
extensive resonance structure. Previous calculations neglecting resonances seri-
ously underestimate the Maxwellian averaged effective collision strength Υ(T )
(Eq. 10) by several factors.
4. TIPTOPBASE: atomic radiative and collisional data
The existing electronic archive for the OP data, TOPbase (http : //heasarc.gsf c.nasa.gov
at Goddard, NASA and http : //vizier.u − strasbg.f r/OP.html at CDS) con-
tain (i) photoionization cross sections of bound LS terms for ∼ 200 ions with Z
= 1 - 14, 16, 18, 20, 26, (ii) energy levels and transition probabilities (∼ 107 f -
values), (iii) monochromatic and Rosseland mean opacities (at CDS only)
The new database, TIPTOPbase (under development, C. Mendoza & the
OP/IP team) will have radiative and collisional data from the OP and the IP.
The data include:
(i) All data from TOPbase (and replacement of improved data), (ii) col-
lisional data for iron and iron peak elements, (iii) new elements through all
ionization stages: P, Cl, K, as well as selected ions such as Ni II and Ni III, (iv)
updated sets of radiative data from new CC calculations, (v) "Tail" photoion-
ization cross sections at high energies that include inner-shell ionization, (vi)
total and level specific recombination rate coefficients, (vi) additional data for
f -values including inner-shell excitations in iron ions Fe VIII - XIII ("PLUS"
data) calculated with SUPERSTRUCTURE, (vii) radiative data (σP I and f -
values) for fine structure levels including relativistic effects, (viii) on-line com-
APS Conf. Ser. Style
9
putational facilities for opacities and radiative accelerations for user-specified
mixtures of elements ("customized" opacities and radiative forces) (see IP in
Extended Abstracts).
5. Conclusion
The current status of large-scale ab initio close coupling R-matrix calculations for
radiative and collisional processes is reported. The Iron Project Breit Pauli R-
matrix radiative calculations include large numbers of dipole allowed and inter-
combination transitions. Self-consistent sets of atomic data for photoionization
and unified (electron-ion) recombination (including RR and DR) are obtained,
and should yield more accurate photoionization models. Work is in progress for
heavy ions of the iron group elements.
Acknowledgments. This work is supported partially by the U.S. National
Science Foundation and NASA.
References
Bautista, M.A. & Pradhan, A.K. 1998, ApJ, 492, 650
Berrington, K.A., Burke, P.G., Butler, K., Seaton, M.J., Storey, P.J., Taylor,
K.T., & Yu, Yan, 1987, J.Phys. B, 20, 6379
Berrington, K.A., Eissner, W.B., & Norrington, P.H. 1995, CPC, 92, 290
Burke, P.G. & Robb, W.D. 1975, Adv. At. Mol. Phys., 11, 143
Chen, G.X. & Pradhan, A.K. 2000, ApJS, 147, 111
Covingtion et al. 2001, Phys.Rev.Lett. 87, 243002
Cunto, W., Mendoza, C., Ochsenbein, F., Zeippen, C.J, 1993, A&A, 275, L5
Eissner, W., Jone, S. W., & Nussbaumer, N. 1974, CPC, 8, 270
Hummer, D.G., Berrington, K.A., Eissner, W., Pradhan, A.K., Saraph, H.E., &
Tully, J.A. 1993, A&A, 279, 298
Kjeldsen, H., Folkmann, F., Hensen, J.E., Knudsen, H., Rasmussen, M.S., West,
J.B., & Andersen, T. 1999, ApJ. 524, L143
Nahar S.N. 1999, ApJS, 120, 131
Nahar S.N. 2002, Phys.Rev.A (in press)
Nahar S.N., Delahaye F.,Pradhan A.K., & Zieppen C.J. 2000, A&AS144, 141
Nahar, S.N. & Pradhan, A.K. 1994, Phys.Rev.A, 49, 1816; 1995, ApJ, 447, 966;
Nahar, S.N 1996, Phys.Rev.A, 53, 1545
Nahar, S.N., & Pradhan, A.K. 1997, ApJS, 111, 339
Pequignot, D., Petitjean, P., & Boisson, C. 1991, A&A, 251, 680; Nussbaumer,
H. & Storey, P.J. 1983, A&A, 126, 75; Badnell, N.R. & Pindzola, M.S.
1989, Phys.Rev.A, 39, 1690; Shull, J.M. & van Steenberg, M. 1982, ApJS,
48, 95
Seaton, M.J. 1987, J.Phys.B, 20, 6363
The Opacity Project 1 and 2, compiled by the Opacity Project team (Institute
of Physices, London, UK, 1995, 1996)
|
astro-ph/9308044 | 1 | 9308 | 1993-08-30T18:35:00 | Reconstructing the Inflaton Potential---Perturbative Reconstruction to Second Order | [
"astro-ph",
"hep-ph"
] | One method to reconstruct the scalar field potential of inflation is a perturbative approach, where the values of the potential and its derivatives are calculated as an expansion in departures from the slow-roll approximation. They can then be expressed in terms of observable quantities, such as the square of the ratio of the gravitational wave amplitude to the density perturbation amplitude, the deviation of the spectral index from the Harrison--Zel'dovich value, etc. Here, we calculate complete expressions for the second-order contributions to the coefficients of the expansion by including for the first time corrections to the standard expressions for the perturbation spectra. As well as offering an improved result, these corrections indicate the expected accuracy of the reconstruction. Typically the corrections are only a few percent. | astro-ph | astro-ph | SUSSEX -- AST 93/8-2
FNAL -- PUB -- 93/258 -- A
astro-ph/9308044
(August 1993)
3
9
9
1
g
u
A
0
3
1
v
4
4
0
8
0
3
9
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
Reconstructing the inflaton potential --
perturbative reconstruction to second-order
Edmund J. Copeland∗
School of Mathematical and Physical Sciences,
University of Sussex, Brighton BN1 9QH, U. K.
Edward W. Kolb†
NASA/Fermilab Astrophysics Center
Fermi National Accelerator Laboratory, Batavia, IL 60510, and
Department of Astronomy and Astrophysics, Enrico Fermi Institute
The University of Chicago, Chicago, IL 60637
Andrew R. Liddle‡
Astronomy Centre, School of Mathematical and Physical Sciences,
University of Sussex, Brighton BN1 9QH, U. K.
James E. Lidsey§
NASA/Fermilab Astrophysics Center
Fermi National Accelerator Laboratory, Batavia, IL 60510
One method to reconstruct the scalar field potential of inflation is a
perturbative approach, where the values of the potential and its derivatives
are calculated as an expansion in departures from the slow-roll approxima-
tion. They can then be expressed in terms of observable quantities, such
as the square of the ratio of the gravitational wave amplitude to the den-
sity perturbation amplitude, the deviation of the spectral index from the
Harrison -- Zel'dovich value, etc. Here, we calculate complete expressions
for the second-order contributions to the coefficients of the expansion by
including for the first time corrections to the standard expressions for the
perturbation spectra. As well as offering an improved result, these cor-
rections indicate the expected accuracy of the reconstruction. Typically
the corrections are only a few percent.
PACS numbers: 98.80.-k, 98.80.Cq, 12.10.Dm
∗Electronic mail: [email protected]
†Electronic mail: [email protected]
‡Electronic mail: [email protected]
§Electronic mail: [email protected]
I. INTRODUCTION
An intriguing prospect raised by recent large-scale structure observations, par-
ticularly those of the Cosmic Background Explorer (COBE) satellite [1], is that ob-
servations may soon provide rather detailed information regarding the nature of the
vacuum energy driving inflation [2,3,4]. In most models this vacuum energy is iden-
tified as the self-interaction potential of a scalar inflaton field, and its precise form is
determined by some particle physics model. Since there currently exist many possible
models [5], it is of great interest to investigate whether observations can select which,
if any, of these models is correct. As well as having detailed implications for the
initial conditions for structure formation in the Universe, the energy scale of inflation
provides a link with particle physics at high energies, and may be a useful method of
probing models of unification.
The prime observational consequences of inflation derive from the stochastic spec-
tra of density (scalar) perturbations and gravitational wave (tensor) modes generated
during inflation. Each stretches from scales of order centimeters to scales well in ex-
cess of the size of the presently observable Universe. Once within the Hubble radius,
gravitational waves redshift away and so their main influence is on the large-scale
microwave background anisotropies, such as those probed by COBE [6]. Advanced
gravitational wave detectors such as the proposed beam-in-space experiments may be
able to detect the gravitational waves on a much shorter (about 1014cm) wavelength
range [7]. The density perturbations are thought to lead to structure formation in
the Universe. They produce microwave background anisotropies across a much wider
range of angular scales than do the tensor modes, and constraints on the scalar spec-
trum are also available from the clustering of galaxies and galaxy clusters, peculiar
velocity flows and a host of other measurable quantities [4].
Recently, we provided a formalism which allows one to reconstruct the inflaton
potential V (φ) directly from a knowledge of these spectra [8,9]. This developed an
original but incomplete analysis by Hodges and Blumenthal [10]. An important result
that follows from our formalism is that knowledge of the scalar spectrum alone is in-
sufficient for a unique reconstruction. Reconstruction from only the scalar spectrum
leaves an arbitrary integration constant, and since the reconstruction is nonlinear, dif-
ferent choices of this constant lead to different functional forms for the potential. A
minimal knowledge of the tensor spectrum, say its amplitude at a single wavelength,
is sufficient to lift this degeneracy. With further information the problem becomes
overdetermined, providing powerful consistency relations which would exclude infla-
tion if not satisfied.
The most ambitious aim of reconstruction is to employ observational data to
deduce the complete functional form of the inflaton potential over the range cor-
responding to large-scale structure. The observational situation is some way from
providing the quality of data that this would require, and at present a more realistic
approach is to attempt a reconstruction of the potential about a single point φ0 [9].
1
For this one requires such information as the amplitudes of both scalar and tensor
modes and also the spectral index of the scalar perturbations at a single scale. It
is possible that such information can be deduced from a combination of microwave
background experiments that span a range of angular scales [11]. In this paper it is
our aim to provide an improved calculation of the coefficients of such a perturbative
reconstruction.
To some extent all inflationary calculations rely on the use of the slow-roll approx-
imation. In the form we present here, the slow-roll approximation is an expansion in
terms of quantities defined from derivatives of the Hubble parameter H. In general
there are an infinite hierarchy of these which can in principle all enter at the same
order in an expansion. However, to calculate V (φ0) one needs only the first deriva-
tive of H, for V ′(φ0) one needs up to the second derivative of H, and so on. These
parameters can be converted into more observationally related quantities, as we shall
see.
The slow-roll approximation arises in two separate places. The first is in simpli-
fying the classical inflationary dynamics of expansion, and the lowest-order approxi-
mation ignores the contribution of the inflaton's kinetic energy to the expansion rate.
The second is in the calculation of the perturbation spectra; the standard expres-
sions are true only to lowest-order in slow-roll. In our earlier work [8,9], we utilized
the Hamilton-Jacobi approach [12] to treat the dynamical evolution exactly, but were
forced for analytic tractability to retain the lowest-order approximation for the pertur-
bation spectra. Technically therefore, the results were accurate only to lowest-order,
though in models close to the power-law inflation limit this hybrid approach offers
substantial improvements for certain quantities.
Until recently further improvements have not been possible, but a very elegant cal-
culation of the perturbation spectra to next order in slow-roll has now been provided
by Stewart and Lyth [13]. This does not permit analytic progress in functional recon-
struction, but their results can be combined with the Hamilton-Jacobi approach to
generate the complete second-order term in perturbative reconstruction. The purpose
of this work is to calculate this correction. This serves two useful purposes. Firstly,
the results allow a more accurate reconstruction to be performed, and secondly the
relative size of lowest-order and second-order contributions provides a useful (though
not rigorous) measure of the theoretical error in reconstruction. As we shall see, even
the lowest-order results are typically accurate to within a few percent.
II. TO SECOND-ORDER IN SLOW-ROLL
We shall use the notation and philosophy of our earlier paper [9], except that the
perturbation spectra shall be given by expressions improving on Eq. (3.4) of that
work. The Hamilton-Jacobi equations arise when one uses the scalar field φ as a time
2
variable, and writes the Hubble parameter H = a/a, where a is the scale factor, as a
function of φ. The field equations are [12]
[H ′(φ)]2 −
3
2
κ2H 2(φ) = −
1
2
κ4V (φ) ,
κ2 φ = −2H ′ ,
(2.1)
(2.2)
where dots are time derivatives, primes are φ derivatives, κ2 = 8π/m2
P l and mP l is the
Planck mass. Without loss of generality we may assume φ > 0, so that H ′(φ) < 0.
Where square roots appear later this choice is used to fix the sign of the prefactor.
The slow-roll approximation can be specified by parameters defined from deriva-
tives of H(φ). There are in general an infinite number of these as each derivative
is independent, but usually only the first few enter into any expressions. We shall
require the first three, which are all of the same order when defined by1
ǫ(φ) =
η(φ) =
ξ(φ) =
2
H(φ)#2
κ2 " H ′(φ)
,
2
κ2
2
κ2
H ′′(φ)
H(φ)
H ′′′(φ)
H ′(φ)
= ǫ −
= η −
ǫ′
,
√2κ2ǫ
2η′
√2κ2ǫ
.
(2.3)
The slow-roll approximation applies when these slow-roll parameters are small in
comparison to unity. The condition for inflation, a > 0, is precisely equivalent to
ǫ < 1.
The lowest-order expressions for the scalar (AS) and tensor (AG) amplitudes as-
sume {ǫ, η, ξ} are negligible compared to unity. Improved expressions for the scalar
and tensor amplitudes for finite but small {ǫ, η, ξ} were found by Stewart and Lyth
[13]:
H 2
H ′
√2κ2
8π3/2
κ
4π3/2 H [1 − (C + 1)ǫ] ,
AS ≃ −
AG ≃
[1 − (2C + 1)ǫ + Cη] ,
(2.4)
(2.5)
where C = −2 + ln 2 + γ ≃ −0.73 is a numerical constant, γ ≈ 0.577 being the Euler
constant. The right hand sides of these expressions are evaluated when the scale in
question crosses the Hubble radius during inflation, 2π/λ = aH. The spectra can
equally well be considered to be functions of wavelength or of the scalar field value.
Eq. (2.7) below allows one to move from one to the other.
The standard results to lowest-order are given by setting the square brackets to
unity. Historically it has been common even for this result to be written as only
1Let us stress that our choice φ > 0 implies √ǫ = −p2/κ2 H ′/H; one needs to be careful with
the signs to reproduce our results.
3
an approximate equality (the ambiguity arising primarily because of a vagueness in
defining the precise meaning of the density perturbation), though the precise nor-
malization to lowest-order was established some time ago by Lyth [14] (see also the
discussion in [4]).
The improved expressions for the spectra in Eqs. (2.4) and (2.5) are accurate in
so far as ǫ and η are sufficiently slowly varying functions that they can be treated
adiabatically as constants while a given scale crosses outside the Hubble radius. Cor-
rections to this would enter at next order. This differs from the usual situation in
which H is treated adiabatically. For the standard calculation to be strictly valid H
must be constant, but provided it varies sufficiently slowly (characterized by small ǫ
and η), it can be evaluated separately at each epoch. This injects a scale depen-
dence into the spectra. There is a special case corresponding to power-law inflation
for which ǫ and η are precisely constant and equal to each other. In this case the
above expressions for the perturbation spectra are exact [13,15]. Furthermore, the
corrections to each spectrum are the same and they cancel when the ratio is taken.
In the general case ǫ and η may be treated as different constants if it is assumed that
the timescale for their evolution is much longer than the timescale for perturbations
to be imprinted on a given scale. This assumption worsens as η is removed from ǫ,
which would be characterized by the next order terms becoming large.
Throughout we shall be quoting results which feature a leading term and a cor-
rection term linear in the slow-roll parameters. We shall utilize the symbol "≃"
to indicate this level of accuracy throughout. The correction terms shall be placed
in square brackets, so the lowest-order equations can always be obtained by setting
the square brackets equal to one. A useful relationship can be obtained from Eqs.
(2.3) -- (2.5):
ǫ ≃
A2
G
A2
S
[1 − 2C(ǫ − η)] .
(2.6)
As we shall see, η is encoded in the spectral index of the scalar perturbations, whose
deviation from unity must also be small for slow-roll to apply.
A key equation in Refs. [8,9] is the consistency equation, which connects the scalar
spectrum to the tensor spectrum and its derivative. The spectra as given in Eqs. (2.4)
and (2.5) are functions of the value of φ when the fluctuations crossed the Hubble
radius during inflation. This is converted into a dependence on wavelength λ with
the relation [9]
dλ
dφ
= λ
H
H ′
κ2
2
[1 − ǫ] .
Differentiation of Eq. (2.5) with respect to φ implies that
d ln AG
dφ ≃ −s κ2
2
AG
AS
[1 + (C + 2)ǫ − (C + 2)η] ,
4
(2.7)
(2.8)
and it follows that
λ
AG
dAG
dλ
=
A2
G
A2
S
[1 + 3ǫ − 2η] .
(2.9)
As expected this agrees with the expansion of the corresponding expression in Ref.
[9] for the special case ǫ = η. This equation is interesting in its own right, but for
perturbative reconstruction its use is restricted to the removal of derivatives of the
tensor spectrum.
III. PERTURBATIVE RECONSTRUCTION TO SECOND ORDER
The aim now is to obtain expressions for the potential and its derivatives about
a single point φ0, given information regarding the spectra at the scale λ0 which left
the horizon at φ = φ0. The four main quantities of observational interest are the am-
plitudes and spectral indices of the two spectra. However, in view of the consistency
equation, Eq. (2.9), only three of these are independent. We shall concentrate on the
two amplitudes and the scalar spectral index, since these are probably the easiest to
measure. The scalar spectral index n is defined by
1 − n =
S(λ)
d ln A2
d ln λ
.
To lowest-order in slow-roll one can show that the spectral index is given by [9]
1 − n = 4ǫ − 2η ,
(3.1)
(3.2)
which provides the route to determining η. Conceptually we are passing from these
three observables to the three parameters that describe the potential, which are the
slow-roll parameters ǫ and η, and the overall normalization. In terms of the observ-
ables, therefore, the slow-roll approximation amounts to an expansion in both A2
G/A2
S
and (1−n), which give corrections to the same order. We shall see that the correction
term for V ′′(φ0) requires the introduction of the third slow-roll parameter ξ, requiring
a new independent observable to determine it.
One obtains directly from the field equation Eq. (2.1) and the definitions of the
spectra in Eqs. (2.4) and (2.5) an expression for the amplitude of the potential:
V (φ0) ≃
≃
48π3
κ4 A2
48π3
κ4 A2
G(λ0)"1 +(cid:18)5
G(λ0)"1 + 0.21
3
G(λ0)
S(λ0)# ,
A2
+ 2C(cid:19) A2
S(λ0)# .
A2
A2
G(λ0)
(3.3)
(3.4)
In Refs. [8,9], we gave the numerical factor on the second-order term as −1/3, which
incorporated only the dynamical slow-roll corrections. In fact, the spectral corrections
5
to V (φ0) dominate the dynamical ones for any inflation model, reversing the sign
of the correction, which may be significant if the tensors are important. However
the relative contribution of the scalar and tensor modes to large angle microwave
anisotropies with our spectral normalization is given approximately by R [9], where
R =
2A2
S
25A2
G
.
(3.5)
Even if the contributions to the anisotropies from scalar and tensor modes are equal,
the correction term in the potential is only 2%. This is a powerful indication that
even the lowest-order perturbative reconstruction promises to be very accurate.
To obtain V ′(φ0) we need the scalar spectral index at λ0, denoted by n0. Differ-
entiation of the potential with respect to φ, followed by some straightforward algebra
gives
V ′(φ0) ≃ −
≃ −
≃ −
96π3
√2κ3
96π3
√2κ3
96π3
√2κ3
A3
G(λ0)
AS(λ0)
A3
G(λ0)
AS(λ0)
A3
G(λ0)
[1 + (C + 2)ǫ + (C − 1/3)η] ,
[1 + 1.27ǫ − 1.06η] ,
AS(λ0) "1 − 0.85
A2
A2
G(λ0)
S(λ0)
+ 0.53(1 − n0)# .
Note that for power-law inflation, which has [6]
(1 − n0) ≃
25
4π
A2
A2
G(λ0)
S(λ0)
,
(3.6)
(3.7)
the corrections in the square brackets nearly cancel, but other models [16,17] can
feature larger corrections, e.g. 16 % for natural inflation with n0 = 0.7.
The calculation for V ′′(φ0) is much more involved. One can show a precise rela-
tionship
V ′′(φ0)
H 2(φ0)
= 3(ǫ + η) −(cid:16)η2 + ǫξ(cid:17) .
(3.8)
A new observable will be needed to determine ξ, the easiest example being the rate of
change of the scalar spectral index. This would be substantially harder to measure,
and it is fortunate that it only enters at second-order.
[It would however enter at
leading order in V ′′′(φ0)]. From Eqs. (2.5) and (3.8), we can obtain the second-order
correction to V ′′(φ0) in terms of the slow-roll parameters
V ′′(φ0) ≃
48π3
κ2 A2
G(λ0) (ǫ + η)"1 + (2C + 2)ǫ −
η2 + ǫξ
3(ǫ + η)# .
(3.9)
It is however harder to convert the prefactor into observables, because there are now
lowest-order terms in both ǫ and η. To generate the correct second-order term, it is
6
not enough to use the first-order expression for η in terms of the spectral index. One
must instead use the second-order result, as given by Stewart and Lyth [13]
1 − n ≃ 4ǫ − 2η + 8(1 + C)ǫ2 − (6 + 10C)ǫη + 2Cǫξ ,
(our η being the negative of their δ), which leads to
(3.10)
4C + 4
5C + 3
ǫ + η ≃ 3ǫ(cid:20)1 +
S (cid:20)1 +
≃ 3
A2
G
A2
ǫ −
3
4 − 2C
3
η +
3
C − 3
3
ǫ +
η +
C
3
ξ(cid:21) −
ξ(cid:21) −
C
3
,
1 − n0
2
1 − n0
2
.
(3.11)
Substituting this into Eq. (3.9) yields
V ′′(φ0) ≃
144π3
A4
A2
G(λ0)
4C + 10
C − 3
S(λ0) (cid:20)1 +
G(λ0) (1 − n0) [1 + (2C + 2)ǫ] −
ǫ +
3
3
η +
C
3
ξ(cid:21)
16π3
κ2 A2
κ2
24π3
κ2 A2
−
G(η2 + ǫξ) ,
(3.12)
where the last term is entirely second-order. Note that there are two lowest-order
terms. An interesting case is η = −ǫ, corresponding to H ∝ φ1/2, for which the
lowest-order term vanishes identically and the final term of Eq. (3.9) is the only one
to contribute. The second derivative of the potential is the lowest derivative at which
it is possible for the expected lowest-order term to vanish.
The final step is to convert the second-order terms into the observables. As they
are already second-order, one only needs the lowest term in their expansion to convert.
From the expression for the spectral index, one finds to lowest-order that
+ 5η − 4ǫ .
(3.13)
1
2ǫ
ξ ≃
dn
d ln λ(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)λ0
Note that the derivative of the spectral index is of order ǫ2. To lowest-order we also
have
η ≃ 2ǫ −
1
2
(1 − n0);
ǫ ≃
A2
G
A2
S
.
Progressively substituting all these into Eq. (3.12) yields
3
(1 − n0) + (36C + 2)
2
A2
G(λ0)
A2
S(λ0)
(1 − n0)
A4
A4
G(λ0)
S(λ0)
V ′′(φ0) ≃
G(λ0)"9
16π3
A2
G(λ0)
κ2 A2
S(λ0) −
A2
1
(1 − n0)2 − (12C − 6)
4
3C − 1
−
+
2
dn
d ln λ(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)λ0# ,
7
(3.14)
(3.15)
where the first two terms are lowest-order and the remainder are second-order.2
For power-law models, the last term is zero and the remaining correction terms
nearly cancel each other, though they are not small individually. For natural inflation
models the correction terms are all individually small, with an overall correction of
about 4% at n0 = 0.7. (In natural inflation, (dn/d ln λ)λ0 ≃ η2 ≃ (1 − n0)2/16 [13].)
IV. DISCUSSION AND CONCLUSIONS
To conclude, we have calculated the full second-order corrections to the perturba-
tive reconstruction of the inflaton potential. The first-order terms agree with those
we found previously [8,9], while the second-order terms offer an improvement. They
serve to quantify the expected errors in the perturbative reconstruction, and in general
these errors are small. Even in cases where tensors provide a substantial contribu-
tion to the large angle microwave background anisotropies and/or the spectral index
deviates significantly from unity, the corrections are typically only a few percent.
Consequently, example figures based on plausible data sets that we presented in our
earlier papers remain valid. One then has some degree of confidence that one can use
our lowest-order expressions as was done in [18].
ACKNOWLEDGMENTS
EJC, ARL and JEL are supported by the Science and Engineering Research Coun-
cil (SERC) UK. EWK and JEL are supported at Fermilab by the DOE and NASA
under Grant NAGW -- 2381. ARL acknowledges the use of the Starlink computer sys-
tem at the University of Sussex. EJC and ARL thank the Aspen Center for Physics
for hospitality during the completion of this work. We would like to thank David
Lyth for helpful discussions.
[1] G. F. Smoot et al., Astrophys. J. Lett. 396, L1 (1992); E. L. Wright et al.,
Astrophys. J. Lett. 396, L13 (1992).
[2] A. Guth, Phys. Rev. D23, 347 (1981); A. Albrecht and P. J. Steinhardt, Phys.
Rev. Lett. 48, 1220 (1982); A. D. Linde, Phys. Lett. 108B, 389 (1982); A. D.
Linde, Phys. Lett. 129B, 177 (1983).
2Factoring out the lowest-order terms as in previous expressions leads to a very complicated
result, so we break with our convention regarding the use of square brackets.
8
[3] E. W. Kolb and M. S. Turner, The Early Universe, (Addison-Wesley, New York,
1990).
[4] A. R. Liddle and D. H. Lyth, "The Cold Dark Matter Density Perturbation,"
to be published, Phys. Rep. (1993).
[5] See Refs. 3 and 4 for surveys of different models.
[6] L. M. Krauss and M. White, Phys. Rev. Lett. 69, 869 (1992); R. L. Davis, H.
M. Hodges, G. F. Smoot, P. J. Steinhardt and M. S. Turner, Phys. Rev. Lett.
69, 1856 (1992); A. R. Liddle and D. H. Lyth, Phys. Lett. 291B, 391 (1992); J.
E. Lidsey and P. Coles, Mon. Not. R. astr. Soc. 258, 57P (1992); D. S. Salopek,
Phys. Rev. Lett. 69, 3602 (1992); F. Lucchin, S. Matarrese, and S. Mollerach,
Astrophys. J. Lett. 401, 49 (1992).
[7] A. R. Liddle, "Can the Gravitational Wave Background from Inflation be De-
tected Locally?" Sussex report SUSSEX-AST 93/7-3 (1993).
[8] E. J. Copeland, E. W. Kolb, A. R. Liddle and J. E. Lidsey, Phys. Rev. Lett.
71, 219 (1993).
[9] E. J. Copeland, E. W. Kolb, A. R. Liddle and J. E. Lidsey, to appear, Phys.
Rev. D, (Sept 15th 1993).
[10] H. M. Hodges and G. R. Blumenthal, Phys. Rev. D42, 3329 (1990).
[11] R. Crittenden, J. R. Bond, R. L. Davis, G. P. Efstathiou and P. J. Steinhardt,
Phys. Rev. Lett. 71, 324 (1993).
[12] D. S. Salopek and J. R. Bond, Phys. Rev. D42, 3936 (1990); J. E. Lidsey, Phys.
Lett. 273B, 42 (1991).
[13] E. D. Stewart and D. H. Lyth, Phys. Lett. 302B, 171 (1993).
[14] D. H. Lyth, Phys. Rev. D31, 1792 (1985).
[15] D. H. Lyth and E. D. Stewart, Phys. Lett. 274B, 168 (1992).
[16] F. C. Adams, J. R. Bond, K. Freese, J. A. Frieman and A. V. Olinto, Phys.
Rev. D47, 426 (1993).
[17] J. E. Lidsey, "Tilting the Primordial Power Spectrum with Bulk Viscosity,"
Fermilab report 93/099 -- A (1993).
[18] M. S. Turner, "Recovering the Inflationary Potential," Fermilab report 93/182-
A (1993).
9
|
astro-ph/9802154 | 1 | 9802 | 1998-02-12T08:49:21 | Gamma-Ray Burst Spectra and Time Histories from 2 to 400 keV | [
"astro-ph"
] | The Gamma-Ray burst detector on Ginga consisted of a proportional counter to observe the x-rays and a scintillation counter to observe the gamma-rays. It was ideally suited to study the x-rays associated with gamma-ray bursts (GRBs). Ginga detected 120 GRBs and 22 of them had sufficient statistics to determine spectra from 2 to 400 keV. Although the Ginga and BATSE trigger criteria were very similar, the distribution of spectral parameters was different. Ginga observed bend energies in the spectra down to 2 keV and had a larger fraction of bursts with low energy power law indexes greater than zero. The average ratio of energy in the x-ray band (2 to 10 keV) compared to the gamma-ray band (50 to 300 keV) was 24%. Some events had more energy in the x-ray band than in the gamma-ray band. One Ginga event had a period of time preceding the gamma rays that was effectively pure x-ray emission. This x-ray ``preactivity'' might be due to the penchant for the GRB time structure to be broader at lower energy rather than a different physical process. The x-rays tend to rise and fall slower than the gamma rays but they both tend to peak at about the same time. This argues against models involving the injection of relativistic electrons that cool by synchrotron radiation. | astro-ph | astro-ph |
Gamma-Ray Burst Spectra and Time Histories From 2 to 400 keV
E. E. Fenimore
MS D436, Los Alamos National Laboratory, Los Alamos, NM 87544
The Gamma-Ray burst detector on Ginga consisted of a proportional counter to observe the x-rays and a
scintillation counter to observe the gamma-rays. It was ideally suited to study the x-rays associated with gamma-
ray bursts (GRBs). Ginga detected ∼ 120 GRBs and 22 of them had sufficient statistics to determine spectra
from 2 to 400 keV. Although the Ginga and BATSE trigger criteria were very similar, the distribution of spectral
parameters was different. Ginga observed bend energies in the spectra down to 2 keV and had a larger fraction
of bursts with low energy power law indexes greater than zero. The average ratio of energy in the x-ray band
(2 to 10 keV) compared to the gamma-ray band (50 to 300 keV) was 24%. Some events had more energy in the
x-ray band than in the gamma-ray band. One Ginga event had a period of time preceding the gamma rays that
was effectively pure x-ray emission. This x-ray "preactivity" might be due to the penchant for the GRB time
structure to be broader at lower energy rather than a different physical process. The x-rays tend to rise and fall
slower than the gamma rays but they both tend to peak at about the same time. This argues against models
involving the injection of relativistic electrons that cool by synchrotron radiation.
1. INTRODUCTION
2. INSTRUMENTAL DETAILS
Gamma-ray bursts (GRBs) are aptly name,
most of their power is usually emitted in the 100
keV to 1 MeV energy range. Early results indi-
cated that only a few percent of the energy of a
Gamma-Ray burst (GRB) occurs in the 2-to-10
keV x-ray range [1,2], although the X-ray emis-
sion might outlast the main GRB event in some
bursts (so called X-ray tails). Based on these
intriguing results, the gamma-ray burst detec-
tor (GBD) flown aboard the Ginga satellite was
specifically designed to investigate burst spectra
in the X-ray regime [3]. Ginga was launched in
February of 1987, and the GBD was operational
from March, 1987, until the reentry of the space-
craft in October, 1991.
More recently, the BeppoSax satellite has ob-
served several bursts in both x-rays and gamma-
rays [4 -- 6] and has discovered soft x-ray afterglows
[7]. This latter discovery has opened the way for
the long-sought GRB counterparts, including one
with a measured redshift [8].
Here, we summarize x-ray results from the
Ginga experiment.
The GBD aboard Ginga consisted of a propor-
tional counter (PC) sensitive to photons in the 2-
to-25 keV range and a scintillation counter (SC)
recording photons with energies between 15-to-
400 keV. Each detector had an ≈ 60 cm2 effective
area. The PC and SC provided 16 and 32 channel
spectra, respectively, over their indicated energy
ranges. The detectors were uncollimated except
for the presence of shielding to reduce backside
illumination and the mechanical support for the
PC window. The field of view was effectively π
steradian for both the PC and the SC.
Each detector produced both spectral and time
history-data for GRBs. Photons within the en-
ergy range of each detector were accumulated
into time samples. The hardware ensured that
the time samples for both detectors started and
stopped at the same time. This facilitates mea-
suring the relative timing between the x-rays
and gamma-rays. The temporal resolution of
the time-history data depended on the teleme-
try mode. An on-board trigger system checked
the time histories for a significant increase (11σ
on either 1/8, 1/4, or 1 s time scales). Upon such
a trigger, special high resolution temporal data
(called "Memory Read Out" [MRO]) would be
2
produced. The MRO data has time resolution of
0.03125 s. The PC MRO time history extends
from 32 s before the trigger to 96 s after the trig-
ger. The SC MRO time history extends from 16
s before the trigger to 48 s after the trigger. Dur-
ing the period when both detectors have MRO
time histories, the MRO samples were in phase
with each other. In addition to the MRO data,
"Real Time" data were often continuously avail-
able. Three different telemetry modes resulted in
PC and SC time histories with either 0.125, 1.0,
or 4.0 s time resolution.
In burst mode, the GBD recorded spectral data
from the PC and SC at 0.5 s intervals for 16 s be-
fore the burst trigger time and for 48 s after the
trigger. The PC and SC had 16 and 32 energy
channels, respectively. The MRO data were used
for many of the spectral fits described here. In
the event that MRO data were not available for a
burst, we utilized the spectral data from the "real
time" telemetry modes. For these bursts, spectral
data were available with either 2, 16, or 64 s accu-
mulations. For the longer accumulations, spectral
studies were not generally feasible. See [3] and [9]
for more information concerning the instrument.
The Ginga GBD was in operation from March
1987 to October 1991. During this time ≈ 120
γ-ray bursts were identified [10,11]. From this
group, a sample of 22 events with statistically
significant spectral data were analyzed [9]. These
events occurred within the forward, π steradian
field of view of the detectors (front-side events).
Front-side events are easily identified because
they have consistent fluxes in the energy range
observed by both instruments (15 to 25 keV). Ex-
cluded bursts usually show strong rolloffs below
25 keV and spectral fits with an absorption com-
ponent due to aluminum (a principal spacecraft
material) are consistent with this interpretation.
Because the incidence angle was known for just
four of the events, we had to assume an incidence
angle for the remaining events. We selected 37
degrees for the incidence angle when the angle
was unknown. This is a typical angle consider-
ing that the mechanical support for the window
on the PC acts as a collimator limiting the field
of view to an opening angle of ∼ ±60 degrees.
Extensive simulations [9] demonstrated that our
conclusions were not affected by the uncertainty
in the incidence angle.
We have adopted the spectral model employed
by [12] because of its relative simplicity and abil-
ity to accurately characterize a wide range of
spectral continua. In addition, this choice facili-
tates direct comparison of our results with those
obtained from bursts seen by the Burst and Tran-
sient Source Experiment (BATSE). This model
has the form
N (E) = AEα exp(−E/E0)
(α − β)E0 ≥ E
N (E) = A [(α − β)E0]α−β exp(β − α)Eβ
if
and
if
(α − β)E0 ≤ E.
Here, A is an overall scale factor, α is the low-
energy spectral index, β is the high-energy spec-
tral index, and E0 is the exponential cutoff or
bend energy. The model parameters are adjusted
iteratively until a minimum in χ2 is obtained.
3. X-RAY CHARACTERISTICS OF
GINGA BURSTS
Of particular interest is the behavior of the
Ginga sample at x-ray energies. About 40% of
the bursts in the sample show a positive spectral
number index below 20 keV (i.e., α > 0), with
the suggestion of rolloff toward lower energies in
a few of the bursts (α as large as +1.5) [9]. Un-
fortunately, the lack of data below 1 keV, and the
often weak signal below 5-to-10 keV precludes es-
tablishing the physical process (photoelectric ab-
sorption, self-absorption) that may be involved
in specific bursts. Observations of the low-energy
asymptote can place serious constraints on several
GRB models, most notably the synchrotron shock
model which predicts that α should be between
-3/2 and -1/2 [13]. Crider et al. [14] uses exam-
ples from BATSE to argue that some GRB vio-
late these limits during some time-resolved sam-
ples. We find violations of these limits in the
time-integrated events.
In our sample of bursts, we find that α can be
both less than zero and greater than zero. Nega-
tive α's are often seen in time-integrated BATSE
spectra. Positive α's for which the spectrum rolls
over at low energies are usually only seen in time-
resolved BATSE spectra [14].
Our distribution of parameters (α, β, E0) is
broader than that found by BATSE. For exam-
ple, 40% of our events had α greater than 0 com-
pared to only 15% of BATSE events, and we had
bend energies that occurred as low as 1.7 keV,
wheras the lowest BATSE bend energy was 15
keV. The Ginga trigger range (50 to 400 keV)
was virtually the same as BATSE's. Thus, we
do not think we are sampling a different popu-
lation of bursts, yet we get a different range of
fit parameters. One possible explanation might
be that GRBs have two break energies, one often
in the 50 to 500 keV range and the other near 5
keV. Both BATSE and Ginga fit with only a sin-
gle break energy so BATSE tends to find breaks
near the center of its energy range, and we tend
to find breaks in our energy range. Without good
high energy observations of bursts with low E0,
it is difficult to know whether they also have a
high-energy bend.
4. X-RAY EMISSION RELATIVE TO
GAMMA-RAY EMISSION
Early measurements of the x-rays associated
with gamma-rays were fortuitous observations by
collimated x-ray detectors that just happen to
catch a GRB in their field of view. From a few
events it appeared that the amount of energy in x-
rays was only a few percent (Laros et al., 1984 [1])
confirming that GRBs were, indeed, a gamma-ray
phenomena. The Ginga experiment was designed
with a wide field of view to detect a sufficient
number of events to determine the range of x-ray
characteristics. Early reports from Ginga events
indicated that sometimes a much larger fraction
of the emitted energy was contained in the x-rays.
For example, by comparing the signal in the pro-
portional counter (roughly 2 to 25 keV) to that
of the scintillator (roughly 15 to 400 keV), we re-
ported an x-ray to gamma-ray emission ratio up
to ∼46% (Yoshida et al., 1989 [15]). Such a ratio
3
Figure 1.
The distribution of the ratio of the
energy emitted in x-rays relative to that emitted
in gamma-rays. The x-ray bandpass is defined to
be from 2 to 10 keV, and the gamma-ray bandpass
is the BATSE range of 50 to 300 keV. Note there
are some examples of equal energy in the x-rays
and the gamma-rays. The average ratio of the
energy in the x-rays to the energy in the gamma-
rays is 24%. From Strohmayer et al., 1998.
2 EN (E)dE/ R 300
depends on the bandpass for which it is evaluated.
Using the best fit α, β, and E0 from [9], one can
find the ratio of emission for a typical x-ray band-
pass (2 to 10 keV) compared to the BATSE en-
ergy range (50 to 300 keV). The ratio is defined to
be Rx/g = R 10
50 EN (E)dE. Fig-
ure 1 presents the distribution for the 22 events
analyzed in [9]. Although the ratio is often a
few percent, for some events the ratio is near
(or larger than) unity.
Some GRBs actually
have more energy in the x-ray bandpass than the
gamma-ray bandpass. The simple average of the
22 values is 24%. This large value arises because
of the few events with nearly equal energy in the
x-ray and gamma-ray bandpass. However, even
the logarithmic average is 7%.
4
5. X-RAY PREACTIVITY OR PULSE
SPREADING?
It is common for the x-ray emission to last
longer than the gamma-ray emission [4 -- 6,15 -- 17].
There is one clear example of x-ray activity pre-
ceding the GRB [16]. Figure 2 presents the time
history of GB900126 in several energy bands.
From top to bottom, one sees an x-ray hard-
ness ratio, the count rate in the PC (1-to-28
keV, 0.03125 s resolution), and five time histo-
ries based on the PC and SC pulse height analy-
sis (PHA) data (0.5 s resolution). The horizontal
arrow in the bottom panel indicates the period
that is effectively pure x-rays. During the period
marked by the arrow, there is no detectable emis-
sion above 7 keV, but there is significant emission
below 7 keV (see Murakami et al. [16]).
It would be misleading to refer to the x-ray
phase as a "precursor" because it is not a sep-
arate peak. "Preactivity" gives the connotation
of a separate cause for the emission although not
necessarily as distinct as "precursor" might imply.
However, the x-ray phase in Figure 2 could be due
to the penchant for peaks to be wider at low en-
ergy. Figure 3 shows the average autocorrelation
of GRB time histories in four energy bands. At
higher energy, the autocorrelation is wider. This
spreading of the time structure with energy is also
seen in the average rise and fall times of individ-
ual pulses [18]. On average, the peak temporal
width scales as E−0.45. The spreading in Figure
2 is larger but that could reflect that some bursts
have larger spreading than the average.
Perhaps there is not a separate cause for the x-
ray emission, but rather, the physics responsible
for the spreading as a function of energy causes
the x-rays to appear to turn on before the gamma-
rays.
The physics responsible for the spreading is un-
clear. Kazanas, Titarchuk, & Hua (1998) [19]
have suggested two processes, synchrotron cool-
ing of injected relativistic electrons and Compton
downscattering of injected photons. Synchrotron
cooling of ejected electrons can produce a time
structure that scales as E−1/2. However, such
cooling should cause the emission at lower en-
ergy to peak later. In most GRBs, the rise and
the fall are slower at low energy (causing the
E−0.45 spreading) but the peak of the emission
is not delayed substantially. In the calculations
of Kazanas, Titarchuk, & Hua, the low energy
emission (e.q., 25-to-50 keV) starts to rise after
the high energy emission (e.q., 300-t0-1000 keV)
has completely fallen. Because GRBs never seem
to do this, we think it strongly argues against syn-
chrotron cooling as the mechanism that produces
the peaks in GRBs.
REFERENCES
1. Laros, J. G., et al., 1984, ApJ, 286, 681.
2. Katoh, M., et al., 1984, in High Energy Tran-
sients in Astrophysics, AIP Conf. 115, ed. S.
E. Woosley (AIP New York), 390.
3. Murakami, T., et al., 1989, Publ. Astron. Soc.
Jap., 41, 405.
4. Piro, L., et al., 1997, A&A, in press, astro-
ph/9707215.
5. Piro, L., et al., 1998, A&A, in press, astro-
ph/9710334.
6. Frontera, F., et al., 1998, ApJ, in press, astro-
ph/9711279.
7. Costa, E., et al., 1997, Nature, 387, 783,
astro-ph/9706065.
8. Metgzer, M., et al., 1997, Nature, 387, 878.
9. Strohmayer, T., et al., 1998, ApJ, in press,
astro-ph/9712332.
10. Ogasaka, Y., et al., 1991, ApJ, 383, L61.
11. Fenimore, E. E., et al., 1993, in AIP Confer-
ence Proceeding 280, Compton Gamma-Ray
Observatory, ed. M. Friedlander, N. Gehrels,
& D. Macomb, (New York:AIP), 917.
12. Band, D., et al., 1993, ApJ, 413, 281.
13. Katz, J. L., 1994, ApJ, 432, L107.
14. Crider, A., et al., 1997, 479, L93.
15. Yoshida, A., et al., 1989, PASJ, 41, 509.
16. Murakami, T., et al., 1991, Nature, 350, 592.
Inoue, H., van Paradijs,
17. Murakami, T.,
J., Fenimore, E., & Yoshida, A., 1992,
in
Gamma-Ray Bursts, ed. C. Ho, R. I. Epstein
& E. E. Fenimore, (Cambridge: Cambridge
University Press), 239.
18. Fenimore, E. E., et al., 1995, ApJ, 448, L101.
19. Kazanas, D., Titarchuk, L. G., & Hua, X.-M.,
1998, ApJ, in press, astro-ph/9709180.
5
Figure 2. The temporal evolution of GB900126
demonstrating the x-ray "preactivity." The top
panel is the x-ray hardness and the next panel is
the total count rate in the proportional counter.
The next 5 panels use the energy-resolved PHA
data in 5 energy bands from 1 to 370 keV. Note
that the peak widths become wider with lower
energy but they are not substantially shifted in
time. Synchrotron cooling of ejected electrons is
expected to produce substantial shifts at lower en-
ergy. In the bottom panel, the time period mark
with a horizontal arrow is the period of x-ray pre-
activity during which the emission occurs only at
energies less than 7 keV. (From Murakami et al.,
1989.)
Figure 3. Average autocorrelation of 45 bright
BATSE gamma-ray bursts in four energy chan-
nels. At higher energy, GRBs have shorter
timescales. The solid curves are fits of the sum
of two exponentials to the autocorrelation his-
togram. (From Fenimore et al. 1995.)
|
astro-ph/9903375 | 1 | 9903 | 1999-03-24T17:49:15 | SNRs in the Galactic Center region observed with ASCA | [
"astro-ph"
] | We report the ASCA results of the supernova remnants (SNRs) and their candidates in the Galactic Center region. We found apparent X-ray emission from G359.1-0.5 and G0.9+0.1, and made marginal detection for G359.1+0.9, but found no significant X-ray from the other cataloged SNRs: G359.0-0.9, Sgr A East (G0.0+0.0), G0.3+0.0, Sgr D SNR (G1.0-0.1) (Green 1998). The emission from G359.1-0.5 is found to be thermal with multi temperature structures whereas that from G0.9+0.1 is quite hard, probably non-thermal. We discovered two new candidates of SNRs: G0.0-1.3 (AX J1751-29.6) and G0.56-0.01 (AX J1747.0-2828). The former, G0.0-1.3, shows the extended emission with a thin thermal plasma. The latter, G0.56-0.01, shows quite strong 6.7-keV line with the equivalent width of 2 keV, which resembles that of the Galactic Center plasma. We discuss the nature of those SNRs, relating with the origin of the Galactic Center hot plasma. | astro-ph | astro-ph |
SNRs in the Galactic Center region observed with ASCA
M. Sakano1,3, J. Yokogawa1, H. Murakami1, K. Koyama1,5 Y. Maeda2,4 and The ASCA Galactic
Plane/Center Survey team
1 Department of Physics, Kyoto University, Sakyo, Kyoto 606-8502, Japan
2 Department of Astronomy and Astrophysics, Pennsylvania State University, University Park PA 16802-6305 U.S.A.
3 Research Fellow of the Japan Society for the Promotion of Science
4 Postdoctctoral Fellow for Research Abroad of JSPS
5 CREST: Japan Science and Technology Corporation (JST)
Abstract. We report the ASCA results of the super-
nova remnants (SNRs) and their candidates in the Galac-
tic Center region. We found apparent X-ray emission
from G359.1−0.5 and G0.9+0.1, and made marginal
detection for G359.1+0.9, but found no significant X-
ray from the other cataloged SNRs: G359.0−0.9, Sgr
A East (G0.0+0.0), G0.3+0.0, Sgr D SNR (G1.0−0.1)
(Green 1998). The emission from G359.1−0.5 is found to
be thermal with multi temperature structures whereas
that from G0.9+0.1 is quite hard, probably non-thermal.
We discovered two new candidates of SNRs: G0.0−1.3
(AX J1751−29.6) and G0.56−0.01 (AX J1747.0−2828).
The former, G0.0−1.3, shows the extended emission with a
thin thermal plasma. The latter, G0.56−0.01, shows quite
strong 6.7-keV line with the equivalent width of 2 keV,
which resembles that of the Galactic Center plasma. We
discuss the nature of those SNRs, relating with the origin
of the Galactic Center hot plasma.
1. Introduction
The hard X-ray structure of the Galactic Center region is
quite unusual as well as those of the other wavelengths.
In particular, thin thermal hot plasma prevails all over
the region (Koyama et al. 1989; Yamauchi et al. 1990;
Koyama et al. 1996; Maeda 1998). It has several surpris-
ing natures. First, the temperature is quite high, about
10 keV. Second, the total energy is also high, about 1054
erg. Third, the plasma wide-spreads in the region bigger
than 1◦×1◦, which corresponds to about 150pc×150pc for
the distance of the Galactic Center. Fourth, the spectrum
of the plasma is surprisingly uniform from field to field
except for the slight change of the surface brightness. The
origin of the plasma is still an enigma.
Supernova remnants (SNRs) are one of the possi-
ble candidates for the origin of the hot plasma (e.g.,
Koyama et al. 1986), as well as the other candidates;
the past activity of the central massive black hole (e.g.,
Koyama et al. 1996), unresolved many cataclysmic vari-
ables (e.g., Mukai & Shiokawa 1993), and the global mag-
netic activity which heats the interstellar matter (e.g.,
Yokoyama et al. 1998).
If the SNRs explain all the hot plasma, first, the to-
tal energy (∼ 1054 erg) requires about 103 supernovae
in the narrow region with the scale of a few hundred pc
in the last 5 104 years, which is the age of the plasma
(Koyama et al. 1996). It would lead to the much higher
supernova rate than the common rate of a few supernovae
in a century in all the Galaxy. However, the possible star-
burst activity in the Galactic Center region has been often
discussed. In particular, recent COMPTEL observations of
26Al 1.8 MeV line suggest that 105 supernovae occurred
in the last 4 106 years (Chen et al. 1995; Hartmann 1995),
which is consistent with the required supernova rate.
Second, the supernova origin hypothesis requires the
mean temperature of about 10 keV. It is much higher than
the usual temperature of SNRs (kT ≤ a few keV). How-
ever, if SNRs densely exist in the region, the temperature
may heat up to about 10 keV by the mutual interaction
of the shocks. Thus, SNRs are still one of the candidates
for the origin of the Galactic Center plasma.
In this paper we concentrate on the X-ray natures
of the SNRs in the Galactic Center region from the ob-
servational point of view. The ASCA capability of imag-
ing spectroscopy with 0.5 -- 10 keV (Tanaka et al. 1994) is
quite suitable for the study. In fact, owing to the hard
X-ray sensitivity of ASCA, particularly above 2 keV, we
can possibly detect new SNRs which have been hidden
by the heavy interstellar absorption, hence obtain some
information whether the SNRs are valid for the origin of
the Galactic Center plasma.
Using all the available ASCA pointing observations
in this region (−1◦< l <1◦, −1◦< b <1◦) and a part
of the data of the ongoing ASCA Galactic Center sur-
vey, we studied the X-ray emission from the known SNRs
(Green 1998), and searched new SNRs. We assume the
distance to the Galactic Center to be 8.5 kpc.
2
Sakano et al.: SNRs in the Galactic Center region observed with ASCA
2. Results on the cataloged SNRs
We searched X-ray emission from the cataloged SNRs
(Green 1998), and then investigated each nature where
significant X-ray emission is detected.
2.1. G359.1−0.5
Fig. 1 (upper panel) shows the ASCA GIS image of
G359.1−0.5 with 1.6 -- 2.1 keV band. We detected center-
filled X-rays from this source, whereas the radio image
shows a clear shell-like structure (e.g., Uchida et al. 1992).
The spectrum is found to have emission lines from
highly ionized ion (Fig. 1 (lower panel)), hence the X-
rays come from thin thermal plasma. The most dis-
tinct two lines are Kα lines from helium-like silicon
and hydrogen-like sulfur. They imply that the plasma
has multi-temperature structure. The absorption column
density is estimated at NH∼ 8 1022H cm−2, suggest-
ing that G359.1−0.5 is located at near the Galactic
Center, which is consistent with the radio observations
(Uchida et al. 1992). This column density is larger by a
factor of 2 or 3 than that with the ROSAT measurement
(Egger & Sun 1998). However, we are still confident of our
estimation because the ASCA energy band is the most
suitable for measuring such heavy absorption. Details of
the analysis are given in Yokogawa et al. (1999).
2.2. G0.9+0.1
Fig. 2 (upper panel) shows the GIS image of G0.9+0.1
with 3 -- 10 keV band. We detected significant X-ray emis-
sion from this source in this hard energy band, whereas
no significant X-ray is detected in the softer energy band.
The X-ray emitting region is compact and not resolved
with ASCA GIS. The radio size of the source is 2′ in diam-
eter (Helfand & Becker 1987), hence it is consistent with
the X-ray image.
The spectrum is found to be hard; kT >2 keV
(the best-fit of 40 keV) in thermal model or Γ ∼1.5
in power-law model, with heavy absorption of NH∼
1023H cm−2, although the statistics are not good. This
hardness may imply the emission to be non-thermal ori-
gin. The flux is 2 10−12 erg s−1 cm−2 in 2 -- 10 keV
band. It is consistent with the upper limit by Einstein
(Helfand & Becker 1987), but significantly lower than the
SAX result (Mereghetti et al. 1999).
2.3. G359.1+0.9
Fig. 3 (upper panel) shows the GIS image around
G359.1+0.9. X-ray emission from the position correspond-
ing to G359.1+0.9 is detected with the significance of 9.8σ
in 0.7 -- 3 keV band (AX J1739.6−2911 in Fig. 3), whereas
the significance in 3 -- 10 keV band is 2.1σ.
Faint extended emission around AX J1739.6−2911
also can be seen in the 0.7 -- 3 keV band image. We
made the radial profile with the center at the peak
of AX J1739.6−2911 and fitted it with the model of
the point spread function (PSF) and the background
(NXB+CXB(cosmic X-ray background)) where the nor-
malizations of the PSF and the background were allowed
to be free. The model is found to be rejected with the con-
fidence of 97.4%, i.e., the slightly extended emission with
the radius of 4′∼5′ was marginally detected with the sig-
nificance of 2.2σ (Fig. 3 lower panel). The size is consistent
with the radius of the radio shell, r ∼5′.
the
central
region
The
spectrum of
of
AX J1739.6−2911 is
found to be well fitted with
the thin thermal plasma model with kT ∼0.7 keV
(χ2/d.o.f.=2.67/5) and NH∼ 0, although the statistics
are not good. The flux is 3 10−13 erg s−1 cm−2 in total
energy band, converted to the luminosity of 2 1033 erg s−1
if we assume the distance to be 8.5 kpc. Note that the soft
spectrum with no absorption may imply that this source
is not located in the Galactic Center region, but is a
foreground source, hence the above estimated luminosity
may be reduced by several factors or more.
The extended emission with positional coincidence
with the radio structure strongly supports that it is an
X-ray counter part of the SNR, if the detection of the
extended emission is true. The soft spectrum also sup-
ports it. The luminosity may be rather low for an SNR.
However, ASCA Plane survey has failed to detect many
radio SNRs, suggesting their quite dim X-ray luminosities
(Yamauchi & ASCA Galactic Plane Survey team 1999).
Therefore, the low X-ray luminosity of this source may
be acceptable for an SNR. Future observations with high
sensitivity will be encouraged.
2.4. The other cataloged SNRs
For the other cataloged SNRs (G359.0−0.9, Sgr A East
(G0.0+0.0), G0.3+0.0, Sgr D SNR (G1.0−0.1)), we de-
tected no significant X-ray emission associated with the
SNRs. It is possibly because no strong X-ray is actu-
ally emitted. However we should be conservative for no
X-ray emission from those sources because the detected
positions of those SNRs are heavily contaminated from
nearby strong X-ray sources; SLX 1744−299/300 near
G359.0−0.9, Sgr A and AX J1745.6−2901 near Sgr A
East, 1E 1743.1−2843 near G0.3+0.0, and GX3+1 near
G1.0−0.1.
3. Results on the new candidates of SNRs
We discovered two new candidates of SNRs. In this sec-
tion, we report the preliminary results on them.
Sakano et al.: SNRs in the Galactic Center region observed with ASCA
3
3.1. G0.0−1.3 (AX J1751−29.6)
Fig. 4 (upper panel) shows the GIS image of G0.0−1.3
(AX J1751−29.6) in 0.7 -- 3 keV band. We discovered the
clearly extended emission with the scale of about 40′×10′.
The spectrum is found to have the emission lines from
highly ionized ions, hence is the thermal origin (Fig. 4
lower panel). In fact, the spectrum is well fitted with a
thin thermal plasma model with kT = 0.5 ± 0.08 keV and
NH= (1.3 ± 0.2) 1022 H cm−2. The X-ray flux is ∼ 10−11
erg s−1 cm−2 in 0.5 -- 3 keV band.
The column density suggests that this source may be
in front of the Galactic Center region according to Sakano
et al. (1999), and be located at the distance of about 4 kpc
if we assume the mean interstellar density of 1 H cm−3.
Then, the obtained flux is converted to the luminosity
of 3 1035 erg s−1 under the assumption of the distance of
4 kpc. Even in the case of the quite small distance of 1 kpc,
the luminosity is larger than 1034 erg s−1.
We now consider the classification of the source. The
clearly extended emission of the thin thermal plasma with
kT ∼0.5 keV implies that this source is an SNR or a star
forming region. The luminosity is also within the typical
range of SNRs but much higher than that of a star forming
region. Therefore, AX J1751−29.6 is a strong candidate
for a new SNR.
3.2. G0.56−0.01 (AX J1747.0−2828)
Fig. 5 (upper panel) shows the GIS image of G0.56−0.01
(AX J1747.0−2828) in 6 -- 7 keV band, where this source
can be seen the most significantly. The X-ray emitting
region is compact and not resolved with GIS.
The background subtracted spectrum is given in
Fig. 5 (lower panel). We accumulated the source X-
ray photons from the 3′-radius circular region around
AX J1747.0−2828, and the background photons, from an
elliptical region with the major axis parallel to the Galac-
tic Plane, excluding the 3′-radius circular regions around
AX J1747.0−2828 itself and Sgr B2 (see Murakami et al.
(1999)).
The spectrum is found to be characterized with a quite
strong line at between 6 -- 7 keV, which is also implied from
the X-ray image. We tried to fit the spectrum with the
thermal bremsstrahlung and a Gaussian line. Then we
found the equivalent width of the line to be quite large,
∼2 keV, and the center energy of the Gaussian to be
6.63±0.06 keV, being consistent with Kα line from helium-
like iron. Hence, the spectrum is definitely a thin thermal
origin with high temperature of several keV or higher.
We then fitted the spectrum with a thin thermal
plasma model. The model well represents the total spec-
tral shape. The best-fit temperature is kT = 6.0+1.9
−1.5 keV,
the abundance, Z > 2 solar, and the hydrogen column
density, NH= (6.1+1.6
−1.1) 1022 H cm−2. The flux is 1.6 10−12
erg s−1 cm−2 in 0.7 -- 10 keV band.
The large column density suggests this source to be
located at near the Galactic Center. Thus, the absorp-
tion corrected X-ray luminosity is estimated at ∼3.6 1034
erg s−1 under the assumption of the distance of 8.5 kpc.
This high temperature and the overabundance suggest
AX J1747.0−2828 to be a possible new candidate of a
young SNR. The luminosity is also within the range of
that of the typical SNR. Although the supernova rem-
nant is the most probable source, the other possibility, for
example, a cataclysmic variable, still cannot be excluded
(e.g., Terada et al. 1999). In any case, it is a new class
object in the Galactic Center region, which is a good can-
didate for the origin of the Galactic Center plasma.
4. Discussion
SNRs are one of the possible candidates for the origin of
the Galactic Center hot plasma.
With ASCA, we completely surveyed the region of
l < 1◦and b < 0.3◦, where a large portion of SNRs in
this region are expected to exist, and partially surveyed
some areas with larger galactic latitude. The number of
the detected SNRs with ASCA is two or three in 7 cata-
loged SNRs, and two for the new candidates in the Galac-
tic Center region. On the other hand, the required number
to be responsible for the Galactic Center plasma is about
103 in the region of l ≤ 1◦and b ≤ 0.5◦. Therefore, the
number of the detected SNRs is quite insufficient.
From recent
line
radio molecular
observations,
Hasegawa et al. (1998) found over 300 shell-like struc-
tures, possibly SNRs, in the region of l ≤ 0.5◦, which
correspond to about 20 shell-like structures along any line
of sight in the region. Therefore, the number of SNRs may
be possibly too much to resolve with ASCA. Future obser-
vations with higher angular resolution would be required
to solve this problem.
The spectra of four of the detected five SNRs (or can-
didates) are relatively soft (G359.1−0.5, G359.1+0.9, and
G0.0−1.3 (AX J1751−29.6)) or have no line-like feature
(G0.9+0.1). Thus they cannot explain the spectrum of the
Galactic Center plasma. On the other hand, a new SNR
candidate G0.56−0.01 (AX J1747.0−2828) is notable for
the strong iron line similar to that of the Galactic Center
plasma. Such class of objects may be a key to understand
the origin of the plasma. Search for such objects will be
strongly encouraged.
Acknowledgements. The authors express their thanks to all the
members of the ASCA team. We are grateful to Drs. J. P.
Hughes and P. Slane for their valuable comments. MS and
YM acknowledge the support from Research Fellowships of the
Japan Society for the Promotion of Science.
References
Chen, W., Gehrels, N., Diehl, R., 1995, ApJ 440, L57
4
Sakano et al.: SNRs in the Galactic Center region observed with ASCA
250
Egger, R., Sun, X., 1998, In: "The Local Bubble and Be-
yond", Breitschwerdt D., Freyberg M.J., Trumper J. (eds.),
Springer, IAU Colloq. 166, 417
nants", http://www.mrao.cam.ac.uk/surveys/snrs/
Green, D.A., 1998, In: "A Catalog of Galactic Supernova Rem-
200
Hartmann, D.H., 1995, ApJ 447, 646
Hasegawa, T., Oka, T., Sato, F., In: The Physics and Chem-
istry of the Interstellar Medium (Abstract Book), Aachen
V.O.S.-V.
150
Helfand, D.J., Becker, R.H., 1987, ApJ 314, 203
Koyama, K., Awaki, H., Kunieda, H., et al., 1989, Nature 339,
603
Koyama, K., Ikeuchi, S., Tomisaka, K., 1986, PASJ 38, 503
100
Koyama, K., Maeda, Y., Sonobe, T., et al., 1996, PASJ 48, 249
Maeda, Y., 1998, Ph.D thesis, Kyoto University
Mereghetti, S., Sidoli, L., Israel, G.L., 1998, A&A 331, L77
Mukai, K., Shiokawa, K., 1993, ApJ 418, 863
Murakami, H., Koyama, K., Sakano, M., Maeda, Y., 1999, this
50
Sakano, M., Koyama, K., Nishiuchi, M., et al., 1999, Advances
volume
in Space Research, in press
Tanaka, Y., Inoue, H., Holt, S.S., 1994, PASJ 46, L37
Terada, Y., Kaneda, H., Makishima, K., et al., 1999, preprint
Uchida, K.I., Morris, M., Yusef-Zadeh, F., 1992, AJ 104, 1533
Yamauchi, S., ASCA Galactic Plane Survey team, 1999, this
volume
Yamauchi, S., Kawada, M., Koyama, K., et al., 1990, ApJ 365,
532
Yokogawa, J., Sakano, M., Koyama, K., Yamauchi, S., 1999,
Advances in Space Research, in press
Yokoyama, T., Tanuma, S., Kudoh, T., Shibata, K., 1998, In:
The central regions of the Galaxy and galaxies, Sofue Y.
(ed.), Kluwer Academic Publishers, IAU Symp. 184, 355
1.6-2.1keV GIS
A1742-294
+0(cid:176) 20¢
-0(cid:176) 20¢
-0(cid:176) 40¢
1E1740.7-2942
+0(cid:176) 00¢
source region
-1(cid:176) 00¢
50
-0(cid:176) 40¢
100
the Snake
background region
radio shell
150
-0(cid:176) 20¢
200
250
-2
10
-3
10
V
e
k
/
s
/
t
n
u
o
c
-4
10
2
0
-2
1
2
Energy (keV)
5
Fig. 1.
(Upper) GIS contour map with 1.6 -- 2.1 keV
band superposed on a schematic diagram of the radio
structures; the radio shell of G359.1−0.5 and the ra-
dio non-thermal filament, the Snake. Contour level
is
linearly spaced and is saturated for A1742−294 and
1E 1740.7−2942. The accumulated regions for the spectra
are also noted. (Lower) The background-subtracted GIS
spectrum. We also show the best-fit model where we fit the
spectrum with the model of the thermal bremsstrahlung
and two narrow Gaussians with interstellar absorption.
These two figures are adopted from Yokogawa et al.
(1999).
c
Sakano et al.: SNRs in the Galactic Center region observed with ASCA
5
220
200
180
160
140
120
100
80
60
40
+1(cid:176) 20¢
+1(cid:176) 00¢
+0(cid:176) 20¢
G0.9+0.1
250
-0(cid:176) 20¢
PSR B1736-29
200
AX J1740.3-2904
+1(cid:176) 40¢
AX J1738.4-2903
SNR G359.1+0.9
150
-0(cid:176) 40¢
AX J1739.6-2911
100
+1(cid:176) 20¢
AX J1739.3-2924
50
-1(cid:176) 00¢
-0(cid:176) 20¢
+0(cid:176) 40¢
+0(cid:176) 00¢
+0(cid:176) 20¢
+0(cid:176) 40¢
+1(cid:176) 00¢
-1(cid:176) 40¢
40
60
80
100
120
140
160
180
200
220
50
100
150
200
250
Fig. 2.
(Upper) GIS2+GIS3 contour map of G0.9+0.1
with 3 -- 10 keV band. The image is smoothed with a Gaus-
sian filter of σ = 0.75′, and corrected for exposure, vi-
gnetting and the GIS grid structure after subtraction of
non-X-ray background (NXB). Contour level is linearly
spaced. The coordinate is in galactic (lII, bII), and the
north is up. (Lower) The same as Fig. 1 lower panel, but
of G0.9+0.1. The fitting model is an absorbed power-law.
)
s
t
n
u
o
c
(
40
35
30
25
20
15
10
5
0
0
2
4
6
8
10
Radius (arcmin)
of
same
as Fig.
2
the
0.7 -- 3.0
but
The
(Upper) The
of G359.1+0.9 with
positions
upper
Fig. 3.
keV
panel,
detected X-ray
band.
sources
(AX J1738.4−2903, AX J1739.3−2924,
AX J1739.6−2911, AX J1740.3−2904), the radio shell
of G359.1+0.9, and the radio pulsar PSR B1736−29 are
indicated. (Lower) The radial profile with the center at
the peak of AX J1739.6−2911. The profile is fitted with
the point spread function (see text). The best-fit model is
given with the dashed line, whereas the dotted line shows
only the background component in the best-fit model.
6
Sakano et al.: SNRs in the Galactic Center region observed with ASCA
-50
-100
-150
-200
G0.0-1.3
(AXJ1751-29.6)
-250
-300
-350
-400
(b)
-1(cid:176) 00¢
-1(cid:176) 20¢
-1(cid:176) 40¢
-2(cid:176) 00¢
200
180
160
140
120
100
80
60
+0(cid:176) 40¢
(l)
+0(cid:176) 40¢
+0(cid:176) 20¢
+0(cid:176) 00¢
-0(cid:176) 19¢ 59†
-0(cid:176) 40¢
Sgr B2 cloud
+0(cid:176) 20¢
AX J1747.0-2828
+0(cid:176) 20¢
1E1743.1-2843
+0(cid:176) 00¢
-50
0
50
100
150
200
250
300
60
80
100
120
140
160
180
200
Fig. 4.
(Upper) The same as Fig. 2 upper panel, but
of G0.0−1.3 (AX J1751−29.6) with 0.7 -- 3.0 keV band.
(Lower) The same as Fig. 1 lower panel, but of G0.0−1.3.
For the background spectrum, we accumulated the pho-
tons from the elliptical region surrounding the source, ex-
cluding the source region, in the same GIS field of view.
The fitting model is the thin thermal plasma model with
absorption.
Fig. 5.
(Upper) The same as Fig. 2 upper panel, but
of G0.56−0.01 (AX J1747.0−2828) with 6.0 -- 7.0 keV
band, which is dominated by iron Kα line. Note that
the image was corrected only for exposure, NXB was
not subtracted, and the region around a bright source
1E 1743.1−2843 was excluded before the smoothing in
order to reduce the contamination from 1E 1743.1−2843.
The positions of the X-ray reflection nebula Sgr B2 cloud
(e.g., Murakami et al. 1999) and 1E 1743.1−2843 are also
indicated. (Lower) The same as Fig. 1 lower panel, but of
G0.56−0.01. The fitting model is the thin thermal plasma
model and 6.4-keV narrow line, both with interstellar ab-
sorption.
|
0710.4274 | 1 | 0710 | 2007-10-23T14:49:42 | Large-scale Propagation of Very Light Jets in Galaxy Clusters | [
"astro-ph"
] | We performed MHD simulations of very light bipolar jets with density contrasts down to 10^-4 in axisymmetry, which were injected into a medium of constant density and evolved up to 200 kpc (200 r_j) full length. These jets show weak and roundish bow shocks as well as broad cocoons and thermalize their kinetic energy very efficiently. We argue that very light jets are necessary to match low-frequency radio observations of radio lobes as well as the bow shocks seen in X-rays. Due to the slow propagation, the backflows and their turbulent interaction in the midplane are important for a realistic global appearance. | astro-ph | astro-ph |
Extragalactic Jets: Theory and Observation from Radio to Gamma Ray
ASP Conference Series, Vol. ??, 2007
T. A. Rector and D. S. De Young (eds.)
Large-scale Propagation of Very Light Jets in Galaxy
Clusters
V. Gaibler, M. Camenzind
Landessternwarte, ZAH, Konigstuhl 12, 69117 Heidelberg, Germany
M. Krause
Astrophysics Group, Cavendish Laboratory, Madingley Road, Cambridge
CB3 0HE, United Kingdom
Abstract. We performed MHD simulations of very light bipolar jets with
density contrasts down to 10−4 in axisymmetry, which were injected into a
medium of constant density and evolved up to 200 kpc (200 rj) full length.
These jets show weak and roundish bow shocks as well as broad cocoons and
thermalize their kinetic energy very efficiently. We argue that very light jets are
necessary to match low-frequency radio observations of radio lobes as well as
the bow shocks seen in X-rays. Due to the slow propagation, the backflows and
their turbulent interaction in the midplane are important for a realistic global
appearance.
1.
Introduction
During the last years, simulations of extragalactic jets with reasonable resolu-
tion and realistic sizes became computationally feasible, which makes compar-
isons between simulated and observed properties possible (Saxton et al. 2002;
Zanni et al. 2003; Carvalho et al. 2005; Krause 2005; O'Neill et al. 2005). Un-
fortunately, the direct physical variables and the observed properties are rather
hard to link, which leaves simulations with a wide range of parameters. Sim-
ulations are mainly governed by the initial setup of the density ratio between
jet and ambient gas, the Mach number and the magnetic field. If the magnetic
field is not dynamically dominant (though important), the density contrast is
the most dominant parameter, but may be one of the hardest to measure. The
thermal jet pressure has turned out to be of little importance in the very light jet
limit (Krause 2003). As the (kinetic) power of a jet can be estimated from ener-
gies in X-ray bubbles, typical values of velocity, lifetime, jet radius and cluster
gas densities indicate that density contrasts of 10−2 to 10−4 (or even lower) are
necessary to describe real sources. Parameter studies support this further, if the
global jet/cocoon/bow shock properties are compared. Thus, we concentrate on
the very light jets with magnetic fields as another important ingredient.
2. Numerical Method and Setup
We examine the evolution of the jets in axisymmetric simulations using the non-
relativistic MHD code NIRVANA (Ziegler & Yorke 1997) and evolve the mag-
1
2
Gaibler, Camenzind, Krause
netic fields using the constrained transport method, which conserves ∇ · B to
machine roundoff errors.
Simulations of very light hydro and MHD jets were performed with density
contrasts η = ρj/ρa between 10−4 and 10−1 where the jet density is ρj and the
ambient gas has a constant density of ρa. We will focus on the MHD jets, as
their hydro counterparts are only for comparison. The bipolar jet was injected
along the Z axis in cylindrical coordinates with a jet radius of rj = 1 kpc, the
jet speed and the sound speed were fixed at 0.6 c and 0.1 c respectively (which
gives internal Mach number 6). The ambient gas has a density of 0.01 mp/cm3
and a temperature of 5 × 107 K. Fully ionized hydrogen was assumed for both
the jet and the external medium. The MHD simulations have an initial dipolar
field in the whole domain with ∼ 20 µG at the jet boundary and a temporally
constant toroidal field with ∼ 15 µG which is confined to the nozzle. For the
10−3 and especially the 10−4 jet, the magnetic fields thus become dynamically
important and influence the appearance. The simulations were run until they
reach the boundary of the grid which has (4000 × 800) or (4000 × 1600) cells
(depending on the density contrast) and the jet radius is resolved with 20 cells.
3. Morphology
Density and temperature distribution for a η = 10−3 jet is shown in Fig. 1.
The jet backflow blows up a pronounced cocoon, surrounded by a thick shell of
shocked ambient matter. Ambient gas is mixed into the cocoon in finger-like
structures due to Kelvin-Helmholtz instabilities at the contact surface. Near
the jet heads, this instability is suppressed by the magnetic field, which leads to
a smoother appearance there. In purely hydrodynamic simulations, this stabi-
lization is absent. As observations at low radio frequencies show quite smooth
contact discontinuities, this indicates the importance of magnetic fields there.
Figure 1. Density and temperature for a bipolar jet with η = 10−3 after 15
Myrs. The upper panel shows the density in units of 10−28 g/cm3, the lower
one shows the logarithm of temperature in units of 1010 K.
Very Light Jets on Large Scales
3
Figure 2.
(as Fig. 3). The bow shock is located at R ≈ 37 kpc.
Pressure slice through the η = 10−3 jet at Z = 0 after 15 Myr
The cocoon is highly turbulent and vortices hitting the jet beam can easily
destabilize and disrupt it for low jet densities. The Mach numbers quickly
decrease and there is no classical "Mach disk" anymore -- the terminal shock
moves back and forth and isn't well-defined.
Because very light jets only propagate slowly, the backflow is strong and the
turbulence makes the interaction between both jets in the midplane important.
These jets have to be simulated bipolarly to get the lateral expansion and hence
the global appearance right.
If only one jet was simulated, the result would
strongly depend on the boundary condition in the equatorial plane (as shown in
Saxton et al. 2002).
Outside of the contact surface is the shocked ambient gas, which is pushed
outwards by the cocoon pressure. The bow shock for very light jets is different
in its shape and strength from that of heavier jets (see section 5.). It is addition-
ally changed by a density profile in the external medium (Krause 2005), which
increases the aspect ratio with time and shows cylindrical cocoons.
As example, a radial pressure slice at Z = 0 is shown for the 10−3 jet
(Fig. 2). The pressure jump at R ≈ 37 kpc is the bow shock and is pretty weak
compared to bow shocks in heavier jets. Shock speed, pressure and density
jump, consistently with the shock jump conditions, give a Mach number of 1.4.
Observations of bow shocks (e.g. Hercules A in Nulsen et al. 2005), which
are possible with modern X-ray telescopes, show low Mach numbers and low
ellipticity, thus supporting the necessity for very light jet parameters. To find
the right cocoon shapes, for comparison low-frequency radio observations have to
be chosen, because at higher frequencies only a small part of the cocoon is visible
as radio lobes (cooled-down electron population in the backflow is invisible at
these frequencies).
4
Gaibler, Camenzind, Krause
Pressure maps for a η = 10−1 and a 10−3 jet with the same
Figure 3.
lengths. The pressure is shown logarithmically in dyne/cm2. The heavier jet
is still much overpressured with respect to the ambient gas and has a much
more elongated bow shock compared to the elliptically-shaped bow shock for
the lighter jet.
Pressure vs. density histogram for a 10−3 jet after 2 Myr (left
Figure 4.
panel) and 15 Myr (right panel). Cell counts, pressure and density are shown
logarithmically using cgs units.
4. Pressure Evolution
The pressure slice shows many variations in Fig. 2, which is not surprising if
one looks at the turbulent motion and the mixing inside the cocoon in Fig. 1.
Strong pressure waves travel through the cocoon and try to find pressure balance.
This process is much more effective for very light jets due to the much slower
jet head propagation and it leads to a rather spherical expansion of the bow
shock, just like an overpressured bubble. The cocoon of the 10−1 jet in Fig. 3 is
overpressured by a factor of 20 with respect to the ambient gas, while being a
factor of only 1.5 for the 10−3 jet (and 4.9 for this jet at t = 2.5 Myr).
This can also be seen in the pressure -- density diagrams in Fig. 4. The
ambient gas is described by the patch near (−26, −10), the jet nozzle by the
cells around (−29, −10). Adiabatic compression and expansion leads to the
oblique and longish features present at different positions. Top right of the jet
Very Light Jets on Large Scales
5
Figure 5.
Evolution of the forward and sideways bow shock strength for
density contrasts η = 10−1 (dotted) and 10−3 (solid) as a function of the
monotonically increasing axial bow shock radius.
nozzle position are the cocoon grid points, which spread over a large range of
density to the right because of mixing with shocked ambient gas, which is the
elongated feature top right of the ambient gas position. Comparing the two
different simulation snapshots, we find that the pressure distribution is quickly
adjusting towards the external pressure, in agreement to the findings in Krause
(2003).
5. Bow Shock and Cocoon
The quick decrease in cocoon pressure naturally affects the strength of the bow
shock as it is this pressure that drives the shock sideways. Fig. 5 shows the tem-
poral evolution of the bow shock strength, in terms of external Mach numbers,
for the forward (R = 0) direction as well as the sideways (Z = 0) direction for
jets with different density contrasts. For easier comparison with observations,
the axial bow shock radius is used for the abscissa instead of time (but both
increase monotonically).
The bow shocks in forward direction are always stronger than the sideways
shocks due to the direct impact of the jet onto the ambient gas. The lighter jet
has a much weaker bow shock in all directions and the differences between the
two directions shown are much less pronounced. The axial diameter of the bow
shock increases proportionally to t0.68 after a slower growth during the first 2
Myr, the width grows similarly as t0.64. An exponent of 0.6 is expected for the
blast wave expansion with constant jet power (Krause 2003), while the slower
growth rate in the initial phase behaves more like a Sedov blast wave (fixed
initial energy amount, exponent is 0.4).
The cocoons, measured by their full (bipolar) lengths and their full Z-
averaged widths, grow like the bow shock in axial direction (∝ t0.71), but much
slower in width (∝ t0.38). This leads to a continuously increasing bow shock vs.
cocoon width ratio (Fig. 6) with a very thick layer of shocked ambient gas. This
effect is weak for heavier jets, but more proncounced the lighter the jet is.
6
Gaibler, Camenzind, Krause
Figure 6.
Bow shock / cocoon width ratio over time for a 10−3 jet.
Figure 7. Aspect ratios over full jet length. Dotted line: 10−1 jet, solid:
10−3 jet. In each case, the lower lines refer to the bow shock and the upper
ones to the cocoon.
The aspect ratios (length/width) for the 10−1 and 10−3 jets are plotted
in Fig. 7. The bow shock for the lighter jet starts with a roughly spherical
shape and only slightly increases its aspect ratio (length/width) to a constant
value of 1.4 (Fig. 7). The heavier jet behaves similarly but approaches a much
higher aspect ratio of 2.6. Thus the aspect ratio of the bow shock may be a
good property to compare with observations. The aspect ratio of the cocoons in
contrast continues to increase, with the heavier jet being on much higher values
at all times.
6. Thermalization
From the quick adjustment of pressure towards an average value, one might
expect a strong conversion of the (kinetic) jet power to thermal energy. This,
in fact, is measured for our simulations. Already for the heavy 10−1 jet, on
Very Light Jets on Large Scales
7
average 67 % of the total energy input is measured as thermal energy increase
and only 33 % as kinetic energy increase. For the 10−2 jet this is 72 % vs.
27 %, and for the 10−3 jet already 81 % of the jet power appear as thermal
energy increase with 16 % going into kinetic energy. Here, already 2 % go into
an increase in magnetic energy, because with lower density the magnetic fields of
constant values become more and more important. The 10−4 simulation showed
a thermalization efficiency of 94 %, but the fractions for kinetic and magnetic
energy are now governed by the strong magnetic pressure and will be examined
in the future.
The overall trend to very efficient thermalization for low-density jets nicely
suits the increasingly spherical bow shock shape due to (isotropic) cocoon pres-
sure. It also provides the cluster with a huge amount of thermal energy and
high-entropy plasma, which may be relevant for the problem of cluster heating
and cooling flows (eg. Magliocchetti & Bruggen 2007).
Acknowledgments. This work was also supported by the Deutsche For-
schungsgemeinschaft (Sonderforschungsbereich 439).
References
Carvalho, J. C., Daly, R. A., Mory, M. P., & O'Dea, C. P. 2005, ApJ, 620, 126
Krause, M. 2003, A&A, 398, 113
Krause, M. 2005, A&A, 431, 45
Magliocchetti, M., & Bruggen, M. 2007, MNRAS, 528
Nulsen, P. E. J., Hambrick, D. C., McNamara, B. R., Rafferty, D., Birzan, L., Wise, M.
W., & David, L. P. 2005, ApJ, 625, L9
O'Neill, S. M., Tregillis, I. L., Jones, T. W., & Ryu, D. 2005, ApJ, 633, 717
Saxton, C. J., Sutherland, R. S., Bicknell, G. V., Blanchet, G. F., & Wagner, S. J. 2002,
A&A, 393, 765
Zanni, C., Bodo, G., Rossi, P., Massaglia, S., Durbala, A., & Ferrari, A. 2003, A&A,
402, 949
Ziegler, U., & Yorke, H. W. 1997, Computer Physics Communications, 101, 54
|
astro-ph/0501096 | 1 | 0501 | 2005-01-06T15:43:08 | Galactic Positrons From Localized Sources | [
"astro-ph"
] | The anomalous bump in the cosmic ray positron to electron ratio at $10 GeV$ can be explained as being a component from a point source that was originally harder than the primary electron background and degrades due to synchrotron and inverse Compton losses in the Galaxy while propagating to the Earth's vicinity. The fit is better than can be obtained with homogeneous injection and is attributed to a minimum age threshold. Annihilating neutralinos can provide a fair fit to the data if they have a mass just above 1/2 the mass of the $Z^o$ and if they annihilate primarily in distant density concentrations in the Galaxy. A possible observational consequence of this scenario would be intense inverse Comptonization of starlight at the Galactic center, with a sharp energy cutoff in the emergent photons as a possible signature of the neutralino mass. | astro-ph | astro-ph |
Galactic Positrons From Localized Sources
David Eichler1, Irit Maor2
ABSTRACT
The anomalous bump in the cosmic ray positron to electron ratio at 10 GeV
can be explained as being a component from a point source that was originally
harder than the primary electron background and degrades due to synchrotron
and inverse Compton losses in the Galaxy while propagating to the Earth's vicin-
ity. The fit is better than can be obtained with homogeneous injection and is
attributed to a minimum age threshold. Annihilating neutralinos can provide
a fair fit to the data if they have a mass just above 1/2 the mass of the Z o
and if they annihilate primarily in distant density concentrations in the Galaxy.
A possible observational consequence of this scenario would be intense inverse
Comptonization of starlight at the Galactic center, with a sharp energy cutoff in
the emergent photons as a possible signature of the neutralino mass.
Subject headings: cosmology: dark matter -- diffusion -- elementary particles --
galaxy: center
1.
Introduction
The possibility that weakly interacting dark matter particles (WIMP's) could annihilate
into detectable cosmic radiation was suggested by Silk & Srednicki (1984). Tylka & Eichler
(1987) noted a reported positron excess, curiously localized near 10 GeV [Mueller & Tang
(1985, 1987); Barwick et al.
(1997, 1998)] and considered whether it could be due to the
annihilation of photinos (as a simple example of neutralinos) in the tens of GeV mass range.
The difficulty was that this process, given the laboratory constraints on the neutralinos,
seemed to fall short of providing enough positrons, and the results were not published.
Various papers on this excess eventually appeared [Tylka (1989); Eichler (1989); Turner &
Wilczek (1990); Coutu et al. (1999)], and some noted that the potential for positron excess
1Physics Department, Ben-Gurion University, Beer-Sheva, Israel, [email protected]
2Institute of Astronomy, University of Cambridge, Madingley Road, Cambridge CB3 0WA, UK,
[email protected]
-- 2 --
could be bolstered by clumpiness in the annihilating dark matter or by decay of weakly
unstable dark matter particles.
The approach usually found in present literature is to try and fit the overall e+/(e+ +e−)
ratio, without giving special attention to the curious behaviour at 10 GeV , see Hooper &
Silk (2004). Baltz & Edsjo (2001) considered a whole class of minimal standard supersym-
metric models and failed to get any non-monotonicity in the e+/(e+ + e−) ratio. Eichler &
Maor (2004) (henceforth Paper I) considered annihilation of particles through the Z o-channel
(which we shall henceforth refer to as virtual Z o-decay) noted that non-relativistic virtual
Z o decay (i.e. when the rest mass of the annihilating dark matter particle is slightly above
1/2 the Z o mass) provide a remarkably good fit to the observed e+/(e+ + e−) ratio below
10 GeV mainly due to the positrons that emerge from decaying muons. At higher energies,
however, the predicted e+/(e+ +e−) ratio rose above the observed values within conventional
assumptions about the injection and propagation. In particular, it was assumed in Paper I
that the positrons and primary electrons are each injected with the same spatial profile, and
that their propagation in the Galaxy is identical. The reason for this rise is that some Z o's
decay directly into high energy e+e− pairs so that the e+ energy is half the Z o mass, and
this gives rise to a high energy bump in the e+/(e+ + e−) ratio at about 50 GeV . While this
bump can be partially washed out by losses and escape, it was found that the high energy
e+/(e+ + e−) ratio is nevertheless apparently too high to fit the observations to within 1σ
error bars. As discussed in Eichler (1989) this is a generic problem for any positron source
that is significantly harder than the primary electrons above 10 GeV .
However, the dark matter annihilation scenario for explaining the positron excess in
any case requires that clumping of the dark matter, and a likely place for this is near the
Galactic center. This means that the positrons in our neighborhood that are dark matter
annihilation products would have a minimum age, i.e. the time needed to diffuse from the
source to our neighborhood, and the age distribution of the positrons that make it to the
Earth's vicinity contain fewer young positrons than the age distribution that one associates
with the standard leaky box model. In this letter we consider that the positrons are injected
by an effectively point source at a finite distance, and show that it greatly improves the
fit over that obtained in Paper I. The positron bump at ∼ 10 GeV can be attributed to
halo-type age ∼ 3× 107 yr for the positrons, for over such a lifetime, positrons losing energy
by synchrotron and inverse Comptonization would end up at about this energy.
We will find that obtaining a good fit from a single source with a single diffusion coeffi-
cient is difficult. However, it is well known that below a certain intensity, e.g. far enough up
front in a diffusion front, cosmic rays can freely stream. Such behavior is observed upstream
of the Earth's bow shock. Theoretical reasons for such free streaming include the difficulty
-- 3 --
of resonantly scattering cosmic rays through 90 degree pitch angle in the linear wave am-
plitude regime. Also, a larger, counterstreaming , finitely stable component of cosmic rays
would stabilize the smaller free streaming positron component. We therefore allow for the
possibility that a small fraction of the positrons freely stream, and arrive at the Earth's
vicinity much younger than the rest. We find that this improves the fit still further. We
find that the low energy non-monotonicity, which appears in the injection spectrum from
non-relativistic Z o decay Paper I, can be produced also by using a simple power law for
the injected spectrum (using the same propagation model). It can thus be produced by the
combination of a harder (but monotonic) spectrum of injected positrons and propagation
effects. We conclude that it is still too early to unambiguously interpret the low energy
behavior of the spectrum as a signature of self annihilating dark matter.
2. Equations and Results
The steady state diffused equation for the particle number density, n(x, r), is
∂n
∂t
= 0 = Dn − Rn +
1
mZ
∂
∂x (cid:18)mZ
dx
dt
n(cid:19) + I(x)δ (r/L)
(1)
dx
x = E/mZ, D is a diffusion operator, R = Bx0.5 is the escape rate with B ∼ f ew×10−151/s,
dt = Ax2 with A = 8.5 × 10−16 erg/s is the Compton loss rate, corresponding to an
mZ
electromagnetic energy density in the Galaxy of 10−12 erg/cm3, L is a distance scale, and
I(x) is the spectrum injected by a point source.
We assume as in Paper I that the primary electrons and background positrons are
injected homogeneously, Dnb = 0:
nb(x) =
mZ
Ax2 exp(cid:20)−
2mZB
A√x (cid:21)Z ∞
x
Ib(x′)exp(cid:20) 2mZB
A√x′ (cid:21) dx′
Ib,e−(x) = Cx−2
Ib,e+(x) = Dx−2.8
f or background e−
f or background e+
(2)
(3)
For the Z o decay injected spectrum we take the diffusion to be one-dimensional with
a diffusion coefficient D, D = D ∂2
∂rr=L = 0
(conserving the number of particles except for the escape term). With these boundary
conditions, the solution to eq. (1) is
∂x2 , and with boundary conditions such that ∂n
nZ(x, r) =
mZ
Ax2 exp(cid:20)−
2mZB
A√x (cid:21) ×
-- 4 --
Z ∞
x
IZ(x′)exp(cid:20) 2mZB
A√x′ (cid:21) ∞
X−∞
cosh πnr
L i exp(cid:20)−(cid:16)πn
L (cid:17)2 mZD
A (cid:18) 1
x −
1
x′(cid:19)(cid:21) dx′(4)
While D (the diffusion coefficient) and L (the size of the leaky box) are free parameters, we
took r = 8 Kpc, the distance to the galactic centre. K ≡ D/r2 gives the inverse time for
diffusion.
We have chosen a one dimensional diffusion because it gives somewhat better results than
3 dimensional diffusion. This is physically plausible if one considers magnetic fields which
will confine the movement of the charged particles. So the geometry is tube-like, with an
effective cross section such that the total volume is the galactic volume, (20 kpc)3.
IZ(x) is the Z decay products,
IZ(x) = N (cid:16) 0.0344Ie(x) + 0.0344Iµ(x) + 0.0069Iτ (x) + 0.6916Ih(x)(cid:17)
Ie(x) = δ(cid:18)x −
2(cid:19)
1
(5)
Iµ(x) =
Iτ (x) =
+
Ih(x) =
2
2
3 (cid:16) 5 − 36x2 + 32x3 (cid:17)
3 (cid:16) 5 − 36x2 + 32x3 (cid:17)
9 (cid:20)−
9 Z 1
95
3 − 108x2 +
10ak −bk ¯x
d¯x
¯x
14
2
28
9 x
1408
3
x3 −(cid:0)25 + 324x2 + 128x3(cid:1) ln (2x)(cid:21)
a1 = 3, b1 = 10
0 < ¯x < 0.1
a2 = 2, b2 = 4
0.1 < ¯x < 1
Each Ich describes the ch channel of decay, and the pre-factors correspond to the branch-
ing ratios. The calculation was done in zeroth order, assuming 3 massless families and
neglecting the top quark, for details see Paper I. Following the discussion there, we take
N = 1.3 × 10−29 1/(cm3 s) as the annihilation rate per unit volume.
Fig.
(1) shows a fit with a single point source of Z o decay and a single diffusion
coefficient. The good fit to the low energies from Paper I is still present, but at the price
that the excess in energies toward x = 1/2 is now is suppressed by the finite age effect. As
the figure shows, we are now facing a scenario which is opposite to Paper I; the finite age
effect tends to suppress the high energy excess at the price of killing it off altogether.
However, there are several possibilities that avoid this problem: There may be more than
-- 5 --
one source, and there may be more than one route (roughly guided by magnetic field lines)
by which the particles diffuse or freely stream from the source to our vicinity. High energy
particles diffuse much less than low energy ones because they are fewer in number and create
less waves. So their self-generated scattering is less efficient. Thus, the fraction of free
streaming particles should be higher at higher energy. Fig. (2) shows a combination of two
Z o decay components, an older, larger one that arrives via diffusion, and a younger, smaller
component that has managed more free streaming. This figure illustrates that if one takes an
age distribution into account, the flexibility in adjusting the high energy spectrum becomes
much larger, and can be fitted to the data.
For sake of comparison, we also include a power law injected spectrum, fig. (3) shows
various power laws, and fig. (4) shows two components with different ages. We find that as
long as the injected power law is hard enough, one can produce a low energy (5−10 GeV ) dip.
The quality of the fit is almost as good for a power law as for virtual Z o decay. We consider
the low energy dip to have qualitatively more significance than the higher points and have
emphasized those data points accordingly in choosing the best fit. We have deliberately not
quantified this with the standard statistical measures. Trying to get the statistically best
parameters (for either power law or Z o decay as injected spectrums) would wash out the low
energy behavior that we are focusing on.
Although we can reproduce the 7 GeV dip, the peak at E ∼ 15 GeV is still too big for
the HEAT data (though too small for the earlier data). This seems to be a generic feature
of our results, regardless of whether the injection source is virtual Z o decay or a power law.
The problem would be worse if the virtual Z o had an energy well above mZ.
3. Possible Observational Consequences
The hypothesis that the neutralino mass mχ is only slightly more than half the Z o mass
is motivated by several factors: The annihilation cross section can be resonantly enhanced
by a factor of 2 or 3 more than that during annihilation in the early universe, when the ther-
mal broadening of the Z resonance somewhat exceeded its natural width, Greist and Seckel
(1991). Moreover, assuming the smallest allowable mass allows the greatest annihilation rate
since the annihilation reaction rate is fixed by the condition that it allows a given cosmic
dark matter contribution. (Although dark matter clumping can enhance the annihilation
rate, a plausible level of such enhancement is limited by observational constraints on dark
matter clumping that are set by stellar distributions in galactic centers.) Making mχ just
above mZ causes the Z o resonance to be asymmetric, but this would be hard to measure
-- 6 --
experimentally because of the weak coupling of the emerging neutralinos at CM collision
energies above 2mχ. On the other hand, that the annihilation cross the virtual Z o must be
close to its mass shell if it is to provide a decent fit suggests that its loop corrections would
be large and it might be discernable or falsifiable with particle collider data on processes
that depend on such loop corrections.
In an astronomical context, a possible observational consequence of a point source of
positrons at the Galactic center could be inverse Comptonization of starlight, which is far
more intense than at a typical point in the Galaxy. The profile of Galactic starlight near
the Galactic center is given by Kent
(1992). The derived photon energy density is then
U(r) = 4.3 × 10−9(r/pc)−0.85 erg/s. Assuming the positrons are produced within a typical
radius r of the Galactic Center, they produce a minimum of EIC(γ) ≡ R 100pc
γ 2σT U(r)dr
in inverse Compton (IC) scattered starlight before escaping the central 100 pc region, and
the luminosity L(≥ γmec2) above γ 2ǫph, where ǫph is the typical energy of the pre-scattered
starlight photons, is R ∞
γ I(γ ′mec2)EIC(γ ′)dγ ′. The most energetic e+e− pairs alone, that is
those which result directly from the Z o decay (Ie(x) in eq. 6), will produce IC luminosity of
3.8 × 1035 erg/s.
r
Fig.
(5) shows the logarithmic derivative of the IC luminosity due to the positrons
only, −γ 2Lγ2 = −dL/d ln(γ 2), as a function of the square of the positron Lorentz factor,
γ 2. Shown in the figure is the minimum IC luminosity as a function of the frequency scaled
to the frequency of the pre-scattered photon (x axis). The minimum luminosity assumes
that the positrons emerge from the central region in a straight line. If the mean free path
λ is less than 100 pc, then the predicted luminosity goes up by roughly a factor of (100pc/λ).
For dark matter annihilation that yields direct monochromatic e+e− pairs at Lorentz fac-
tor γo, this would translate into a sharp cutoff in the IC gamma rays of γ 2
o eV .
In our particular example mχ ≃ mZ/2, this would lead to a cutoff at 1010ǫph ∼ 30 GeV . This
would in principle be detectable by MAGIC, see for example Cortina (2004), if the location
were suitable for observing the Galactic center. Alternatively, it could be detected by HESS
if the energy threshold could be pushed to below 30 GeV . This scenario would not explain
the T eV photons from the Galactic center recently reported by the HESS collaboration,
Aharonian et al. (2004). If, however, there is annihilation in the Galactic Center of heavier
dark matter particles, then direct e+e− pairs might be detectable via such a cutoff in the
T eV gamma ray spectrum at mχ/2.
o ǫph or about 3γ 2
In conclusion, we find that the cosmic ray positron data can be fit with more than one
-- 7 --
hard source of positrons provided that a) they have a chance to lose energy before escaping
the Galaxy and b) they have a minimum age (e.g. they come from discrete, distant sources),
unlike the background primary electrons. They need not be from dark matter annihilation,
but a best case scenario for this is not confidently ruled out by existing data. Detection of
inverse Compton radiation with good energy resolution can in principle provide information
as to the spectrum of the positrons and their point of origin.
We thank T. Alexander, E. Baltz, and A. Dekel for helpful discussions. DE gratefully
acknowledges the support from the Israel-U.S. Binational Science Foundation, the Israel
Science Foundation and the Arnow Chair of Physics at Ben Gurion University. IM gratefully
acknowledges the support from the Leverhulme Trust.
REFERENCES
Aharonian, F. et al. [The HESS Collaboration], arXiv:astro-ph/0408145.
Baltz, E. & Edsjo, J. astro-ph/0109318 (2001)
Barwick, S. W. et al. (HEAT collaboration) Astroph. J. 482, L191 (1997)
Barwick, S. W. et al. (HEAT collaboration) Astroph. J. 498, 779 (1998)
Cortina, J. [the MAGIC Collaboration], arXiv:astro-ph/0407475.
Coutu. S. et al. (HEAT collaboration) Astropart. Phys. 11, 429 (1999)
Eichler, D. Phys. Rev. Lett. 63, 2440 (1989)
Eichler, D & Maor, I. Astropart. Phys. 21 (2004) 195.
Greist, K. and Seckel, D. Phys. Rev. D 43, 3191
Hooper, D. & Silk, J. arXiv:hep-ph/0409104.
Kent, S. M. Astroph. J. 387, 181 (1992)
Mueller, D. & Tang, J. Proc. of Nineteenth International Cosmic Ray Conference, La Jolla,
California, NASA Conf. Pub. No. 2376 (U.S. GPO, Washington, DC), 2, 378 (1985)
Mueller, D. & Tang, J. Astroph. J, 312, 183 (1987)
Silk, J. & Srednicki, M. Phys. Rev. Lett., 53, 624 (1984)
-- 8 --
Turner, M. S. & Wilczek, F. Phys. Rev. D 42, 1001, (1990)
Tylka, A. J. & Eichler, D. University of Maryland preprint (1987)
Tylka, A. J. Phys. Rev. Lett. 63, 40 (1989)
This preprint was prepared with the AAS LATEX macros v5.2.
-- 9 --
0.1
0.2
E/mZ
0.3
0.4
0.5
0.25
0.2
0.15
0.1
0.05
0
0
)
−
e
+
+
e
(
/
+
e
Fig. 1. -- The e+/(e+ + e−) as a function of x = E/mZ, for a single source Z decay injected
spectrum. A = 8.5×10−16 erg
cm3 s ,
and K = 1.9× 10−16 1
s . Data taken from Barwick et al. (1997) (black) and Mueller & Tang
(1987) (grey).
cm3 s , D = 1.3×10−31
s , B = 7.1×10−15 1
s , C = 4.0×10−29
1
1
)
−
e
+
+
e
(
/
+
e
0.25
0.2
0.15
0.1
0.05
0
0
both
K1
K2
0.1
0.2
E/mZ
0.3
0.4
0.5
Fig. 2. -- The e+/(e+ + e−) as a function of x = E/mZ, for a combination of 2 sources of
Z decay injected spectrum. A = 8.5 × 10−16 erg
s , C = 4.9 × 10−29
cm3 s ,
D = 1.3 × 10−31
s . The ratio between the
two components is 1 : 5. Data taken from Barwick et al.
(1997) (black) and Mueller &
Tang (1987) (grey).
s , B = 7.6 × 10−15 1
s and K2 = 2.8 × 10−16 1
cm3 s , K1 = 2.8 × 10−14 1
1
1
-- 10 --
w=−0.2
w=−0.3
w=−0.4
w=−0.5
0.1
0.2
E/mZ
0.3
0.4
0.5
0.25
0.2
0.15
0.1
0.05
0
0
)
−
e
+
+
e
(
/
+
e
Fig. 3. -- The e+/(e+ + e−) as a function of x = E/mZ, for various power laws, Nxw, as
the injected spectrum. A = 8.5 × 10−16 erg
cm3 s ,
D = 1.1 × 10−31
(1997)
(black) and Mueller & Tang (1987) (grey).
cm3 s , and K = 6.6 × 10−17 1
s . Data taken from Barwick et al.
s , B = 4.4 × 10−15 1
s , C = 1.7 × 10−29
1
1
)
−
e
+
+
e
(
/
+
e
0.25
0.2
0.15
0.1
0.05
0
0
both
K1
K2
0.1
0.2
E/mZ
0.3
0.4
0.5
Fig. 4. -- The e+/(e+ + e−) as a function of x = E/mZ, for a combination of 2 sources
of power law (w = −0.3) injected spectrum. A = 8.5 × 10−16 erg
s , B = 4.4 × 10−15 1
s ,
s and K2 = 1.2 × 10−16 1
C = 1.7 × 10−29
s .
The ratio between the two components is 10 : 1. Data taken from Barwick et al.
(1997)
(black) and Mueller & Tang (1987) (grey).
cm3 s, K1 = 7.2 × 10−17 1
cm3 s, D = 1.1 × 10−31
1
1
-- 11 --
x 1036
3
2.5
2
1.5
1
0.5
0
108
109
γ2
/
)
s
g
r
e
(
)
2
γ
l
(
n
d
L
d
/
Fig. 5. -- Differential IC luminosity due to the positrons only, −dL/d ln(γ 2) × (20 kpc)3/V
as a function of γ 2.
|
astro-ph/9610078 | 2 | 9610 | 1996-10-24T22:28:22 | The Best Theory of Cosmic Structure Formation is Cold + Hot Dark Matter (CHDM) | [
"astro-ph"
] | I have been asked to make the case here that CHDM is the best theory of cosmic structure formation, and indeed I believe that it is the best of all those I have considered if the cosmological matter density is near critical and if the expansion rate is not too large (i.e. $h \equiv H_0/(100 \kmsMpc) \lsim 0.6$). I discuss CHDM together with its chief competitor among CDM variants, low-$\Omega_0$ CDM with a cosmological constant ($\Lambda$CDM). $\Lambda$CDM with $\Omega_0 \sim 0.3$ has the possible virtue of allowing a higher expansion rate H_0 for a given cosmic age t_0, but the defect of predicting too much fluctuation power on small scales. Also, except for the $H_0-t_0$ problem, there is not a shred of evidence in favor of a nonzero cosmological constant, only increasingly stringent observational upper bounds on it. CHDM has less power on small scales, in good agreement with data, although it remains to be seen whether it predicts early enough galaxy formation to be compatible with the latest high-redshift data. I also briefly compare CHDM to other CDM variants such as Warm Dark Matter (WDM) and tilted CDM. CHDM has the advantage among $\Omega=1$ CDM-type models of requiring little or no tilt, which appears to be an advantage in fitting recent small-angle CMB anisotropy data. The presence of a hot component that clusters less than cold dark matter lowers the effective $\Omega_0$ that would be measured on small scales, which appears to be in accord with observations, and it may also avoid the discrepancy between the high central density of dark matter halos from CDM simulations compared to evidence from rotation curves of dwarf spiral galaxies. | astro-ph | astro-ph |
The Best Theory of Cosmic Structure Formation
is Cold + Hot Dark Matter (CHDM) ∗
Joel R. Primack
University of California, Santa Cruz, CA 95064
[email protected]
January 18, 2018
1 Summary
The fact that the simplest modern cosmological theory, standard Cold Dark
Matter (sCDM), almost fits all available data has encouraged the search for
variants of CDM that can do better. I have been asked to make the case here
that CHDM is the best theory of cosmic structure formation, and indeed I
believe that it is the best of all those I have considered if the cosmological
matter density is near critical (i.e., Ω0 ≈ 1) and if the expansion rate is not too
large (i.e. h ≡ H0/(100 km s−1 Mpc−1) <∼ 0.6). But I think it will be helpful to
discuss CHDM together with its chief competitor among CDM variants, low-Ω0
CDM with a cosmological constant (ΛCDM). While the predictions of COBE-
normalized CHDM and ΛCDM both agree reasonably well with the available
data on scales of ∼ 10 to 100 h−1 Mpc, each has potential virtues and defects.
ΛCDM with Ω0 ∼ 0.3 has the possible virtue of allowing a higher expansion rate
H0 for a given cosmic age t0, but the defect of predicting too much fluctuation
power on small scales. CHDM has less power on small scales, so its predictions
appear to be in good agreement with data on the galaxy distribution, although
it remains to be seen whether it predicts early enough galaxy formation to be
compatible with the latest high-redshift data. Also, several sorts of data suggest
that neutrinos have nonzero mass, and the variant of CHDM favored by this data
-- in which the neutrino mass is shared between two species of neutrinos -- also
seems more compatible with the large-scale structure data. Except for the H0 −
t0 problem, there is not a shred of evidence in favor of a nonzero cosmological
constant, only increasingly stringent upper bounds on it from several sorts of
measurements. Two recent observational results particularly favor high cosmic
∗To appear in Critical Dialogues in Cosmology, ed. N. Turok (World Scientific, 1996).
1
density, and thus favor Ω = 1 models such as CHDM over ΛCDM -- (1) the
positive deceleration parameter q0 > 0 measured using high-redshift Type Ia
supernovae, and (2) the low primordial deuterium/hydrogen ratio measured
in two different quasar absorption spectra.
If confirmed, (1) means that the
cosmological constant probably cannot be large enough to help significantly with
the H0 − t0 problem; while (2) suggests that the baryonic cosmological density
is at the upper end of the range allowed by Big Bang Nucleosynthesis, perhaps
high enough to convert the "cluster baryon crisis" for Ω = 1 models into a crisis
for low-Ω0 models. I also briefly compare CHDM to other CDM variants such as
Warm Dark Matter (WDM) and tilted CDM. CHDM has the advantage among
Ω = 1 CDM-type models of requiring little or no tilt, which appears to be an
advantage in fitting recent small-angle cosmic microwave background anisotropy
data. The presence of a hot component that clusters less than cold dark matter
lowers the effective Ω0 that would be measured on small scales, which appears to
be in accord with observations, and it may also avoid the discrepancy between
the high central density of dark matter halos from CDM simulations compared
to evidence from rotation curves of dwarf spiral galaxies.
2
Introduction
"Standard" Ω = 1 Cold Dark Matter (sCDM) with h ≈ 0.5 and a near-
Zel'dovich spectrum of primordial fluctuations [1] until a few years ago seemed
to many theorists to be the most attractive of all modern cosmological mod-
els. But although sCDM normalized to COBE nicely fits the amplitude of the
large-scale flows of galaxies measured with galaxy peculiar velocity data [2], it
does not fit the data on smaller scales:
it predicts far too many clusters [3]
and does not account for their large-scale correlations [4], and the shape of the
power spectrum P (k) is wrong [5, 6]. Here I discuss what are perhaps the two
most popular variants of sCDM that might agree with all the data: CHDM and
ΛCDM. The linear matter power spectra for these two models are compared in
Figure 1 (from Ref.
[7]) with the real-space galaxy power spectrum obtained
from the two-dimensional APM galaxy power spectrum [5]. The ΛCDM and
CHDM models essentially bracket the range of power spectra in currently pop-
ular cosmological models which are variants of CDM.
CHDM cosmological models have Ω = 1 mostly in cold dark matter but
with a small admixture of hot dark matter, light neutrinos contributing Ων =
mν,tot/(92h2eV) ≈ 0.2, corresponding to a total neutrino mass of mν,tot ≈ 5 eV
for h = 0.5. CHDM models are a good fit to much observational data [8, 9]
-- for example, correlations of galaxies and clusters and direct measurements
of the power spectrum P (k), velocities on small and large scales, and other
statistics such as the Void Probability Function (probability P0(r) of finding
no bright galaxy in a randomly placed sphere of radius r). My colleagues and
I had earlier shown that CHDM with Ων = 0.3 predicts a VPF larger than
2
Figure 1: Power spectrum of dark matter for ΛCDM and CHDM models consid-
ered in this paper, both normalized to COBE, compared to the APM galaxy real-
space power spectrum. (ΛCDM: Ω0 = 0.3, ΩΛ = 0.7, h = 0.7, thus t0 = 13.4
Gy; CHDM: Ω = 1, Ων = 0.2 in Nν = 2 ν species, h = 0.5, thus t0 = 13 Gy;
both models fit cluster abundance with no tilt, i.e. np = 1. From Ref. [7].)
observations indicate [10], but new results based on our Ων = 0.2 simulations
in which the neutrino mass is shared equally between two neutrino species [8]
show that the VPF for this model is in excellent agreement with observations.
However, our simulations [12] of COBE-normalized ΛCDM with h = 0.7 and
Ω0 = 0.3 lead to a VPF which is too large to be compatible with the data [11].
Moreover, there is mounting astrophysical and laboratory data suggesting
that neutrinos have non-zero mass [8, 13]. The analysis of the LSND data
through 1995 [14] strengthens the earlier LSND signal for ¯νµ → ¯νe oscillations.
Comparison with exclusion plots from other experiments implies a lower limit
µe ≡ m(νµ)2 − m(νe)2 >∼ 0.2 eV2, implying in turn a lower limit mν >∼
∆m2
0.45 eV, or Ων >∼ 0.02(0.5/h)2. This implies that the contribution of hot dark
matter to the cosmological density is larger than that of all the visible stars
(Ω∗ ≈ 0.004 [15]). More data and analysis are needed from LSND's νµ →
νe channel before the initial hint [16] that ∆m2
µe ≈ 6 eV2 can be confirmed.
Fortunately the KARMEN experiment has just added shielding to decrease its
3
background so that it can probe the same region of ∆m2
µe and mixing angle,
with sensitivity as great as LSND's within about two years. The Kamiokande
data [17] showing that the deficit of E > 1.3 GeV atmospheric muon neutrinos
increases with zenith angle suggests that νµ → ντ oscillations [18] occur with
an oscillation length comparable to the height of the atmosphere, implying that
∆m2
τ µ ∼ 10−2 eV2 [17] -- which in turn implies that if either νµ or ντ have
large enough mass (>∼ 1 eV) to be a hot dark matter particle, then they must
be nearly degenerate in mass, i.e. the hot dark matter mass is shared between
these two neutrino species. The much larger Super-Kamiokande detector is
now operating, and we should know by about the end of 1996 whether the
Kamiokande atmospheric neutrino data that suggested νµ → ντ oscillations will
be confirmed and extended [19]. Starting in 1997 there will be a long-baseline
neutrino oscillation disappearance experiment to look for νµ → ντ with a beam
of νµ from the KEK accelerator directed at the Super-Kamiokande detector,
with more powerful Fermilab-Soudan, KEK-Super-Kamiokande, and possibly
CERN-Gran Sasso long-baseline experiments in subsequent years.
Evidence for non-zero neutrino mass evidently favors CHDM, but it also dis-
favors low-Ω models. Because free streaming of the neutrinos damps small-scale
fluctuations, even a little hot dark matter causes reduced fluctuation power on
small scales and requires substantial cold dark matter to compensate; thus evi-
dence for even 2 eV of neutrino mass favors large Ω and would be incompatible
with a cold dark matter density Ωc as small as 0.3 [8]. Allowing Ων and the tilt
to vary, CHDM can fit observations over a somewhat wider range of values of the
Hubble parameter h than standard or tilted CDM [20]. This is especially true if
the neutrino mass is shared between two or three neutrino species [8, 21, 22, 23],
since then the lower neutrino mass results in a larger free-streaming scale over
which the power is lowered compared to CDM; the result is that the cluster
abundance predicted with Ων ≈ 0.2 and h ≈ 0.5 and COBE normalization (cor-
responding to σ8 ≈ 0.7) is in reasonable agreement with observations without
the need to tilt the model[24] and thereby reduce the small-scale power further.
(In CHDM with a given Ων shared between Nν = 2 or 3 neutrino species, the
linear power spectra are identical on large and small scales to the Nν = 1 case;
the only difference is on the cluster scale, where the power is reduced by ∼ 20%
[21, 8].)
Another consequence of the reduced power on small scales is that structure
formation is more recent in CHDM than in ΛCDM. This may conflict with obser-
vations of damped Lyman α systems in quasar spectra, and other observations
of protogalaxies at high redshift, although the available evidence does not yet
permit a clear decision on this (see below). While the original Ων = 0.3 CHDM
model [25, 26] certainly predicts far less neutral hydrogen in damped Lyman
α systems (identified as protogalaxies with circular velocities Vc ≥ 50 km s−1)
than is observed [27, 28], lowering the hot fraction to Ων ≈ 0.2 dramatically
improves this [28, 29]. Also, the evidence from preliminary data of a fall-off of
4
the amount of neutral hydrogen in damped Lyman α systems for z >∼ 3 [30] is
in accord with predictions of CHDM [28].
However, as for all Ω = 1 models, h >∼ 0.55 implies t0 <∼ 12 Gyr, which con-
flicts with age estimates from globular cluster [31] and white dwarf cooling [32].
The only way to accommodate both large h and large t0 within the standard
FRW framework of General Relativity is to introduce a positive cosmological
constant (Λ > 0) [33, 34].
ΛCDM flat cosmological models with Ω0 = 1 − ΩΛ ≈ 0.3, where ΩΛ ≡
Λ/(3H 2
0 ), were discussed as an alternative to Ω = 1 CDM since the beginning
of CDM [1, 35]. They have been advocated more recently [36] both because
they can solve the H0 − t0 problem and because they predict a larger fraction
of baryons in galaxy clusters than Ω = 1 models. Early galaxy formation
also is often considered to be a desirable feature of these models. But early
galaxy formation implies that fluctuations on scales of a few Mpc spent more
time in the nonlinear regime, as compared with CHDM models. As has been
known for a long time, this results in excessive clustering on small scales. My
colleagues and I have found that a typical ΛCDM model with h = 0.7 and Ω0 =
0.3, normalized to COBE on large scales (this fixes σ8 ≈ 1.1 for this model),
is compatible with the number-density of galaxy clusters[24], but predicts a
power spectrum of galaxy clustering in real space that is much too high for
wavenumbers k = (0.4 − 1)h/Mpc [12]. This conclusion holds if we assume
either that galaxies trace the dark matter, or just that a region with higher
density produces more galaxies than a region with lower density. One can see
immediately from Figure 1 that there will be a problem with this ΛCDM model,
since the APM power spectrum is approximately equal to the linear power
spectrum at wavenumber k ≈ 0.6h Mpc−1, so there is no room for the extra
power that nonlinear evolution certainly produces on this scale (see Figure 1 of
Ref. [12] and further discussion below). The only way to reconcile the model
with the observed power spectrum is to assume that some mechanism causes
strong anti-biasing -- i.e., that regions with high dark matter density produce
fewer galaxies than regions with low density. While theoretically possible, this
seems very unlikely; biasing rather than anti-biasing is expected, especially on
small scales [37]. Numerical hydro+N-body simulations that incorporate effects
of UV radiation, star formation, and supernovae explosions [44] do not show
any antibias of luminous matter relative to the dark matter.
Our motivation to investigate this particular ΛCDM model was to have H0
as large as might possibly be allowed in the ΛCDM class of models, which in
turn forces Ω0 to be rather small in order to have t0 >∼ 13 Gyr. There is little
room to lower the normalization of this ΛCDM model by tilting the primordial
power spectrum Pp(k) = Aknp (i.e., assuming np significantly smaller than the
"Zel'dovich" value np = 1), since then the fit to data on intermediate scales will
be unacceptable -- e.g., the number density of clusters will be too small [12].
Tilted ΛCDM models with higher Ω0, and therefore lower H0 for t0 >∼ 13 Gyr,
appear to have a better hope of fitting the available data, based on comparing
5
quasi-linear calculations to the data [12, 38]. But all cosmological models with
a cosmological constant Λ large enough to help significantly with the H0 − t0
problem are in trouble with new observations providing strong upper limits on
Λ [39]: gravitational lensing [40], HST number counts of ellptical galaxies [41],
and especially the preliminary results from measurements using high-redshift
Type Ia supernovae [42]. The analysis of the data from the first 7 of the Type
Ia supernovae from the LBL group [43] gave Ω0 = 1 − ΩΛ = 0.94+0.34
−0.28, or
equivalently ΩΛ = 0.06+0.28
−0.34 (< 0.51 at the 95% confidence level).
It is instructive to compare the Ω0 = 0.3, h = 0.7 ΛCDM model that we have
been considering with standard CDM and with CHDM. At k = 0.5h Mpc−1,
Figs. 5 and 6 of Ref. [45] show that the Ων = 0.3 CHDM spectrum and that
of a biased CDM model with the same σ8 = 0.67 are both in good agreement
with the values indicated for the power spectrum P (k) by the APM and CfA
data, while the CDM spectrum with σ8 = 1 is higher by about a factor of two.
As Figure 2 shows, CHDM with Ων = 0.2 in two neutrino species [8] also gives
nonlinear P (k) consistent with the APM data.
3 Cluster Baryons
I have recently reviewed the astrophysical data bearing on the values of the
fundamental cosmological parameters, especially Ω0 [39]. One of the arguments
against Ω = 1 that seemed hardest to answer was the "cluster baryon crisis" [46]:
for the Coma cluster the baryon fraction within the Abell radius is
fb ≡
Mb
Mtot
≥ 0.009 + 0.050h−3/2,
(1)
where the first term comes from the galaxies and the second from gas. If clusters
are a fair sample of both baryons and dark matter, as they are expected to
be based on simulations, then this is 2-3 times the amount of baryonic mass
expected on the basis of BBN in an Ω = 1, h ≈ 0.5 universe, though it is just
what one would expect in a universe with Ω0 ≈ 0.3. The fair sample hypothesis
implies that
Ω0 =
Ωb
fb
= 0.33(cid:18) Ωb
0.05(cid:19)(cid:18) 0.15
fb (cid:19) .
(2)
A review of the quantity of X-ray emitting gas in a sample of clusters [47]
finds that the baryon mass fraction within about 1 Mpc lies between 10 and
22% (for h = 0.5; the limits scale as h−3/2), and argues that it is unlikely that
(a) the gas could be clumped enough to lead to significant overestimates of
the total gas mass -- the main escape route considered in [46] (cf. also [48]).
If Ω = 1, the alternatives are then either (b) that clusters have more mass
than virial estimates based on the cluster galaxy velocities or estimates based
on hydrostatic equilibrium [49] of the gas at the measured X-ray temperature
6
CHDM, 2nu, Pcold(k)
Figure 2: Comparison of APM galaxy power spectrum (triangles) with nonlinear
cold particle power spectrum from CHDM model considered in this paper (upper
solid curve). The dotted curves are linear theory; upper curves are for z = 0,
lower curves correspond to the higher redshift z = 9.9. (From Ref. [7].)
(which is surprising since they agree [50]), (c) that the usual BBN estimate
Ωb ≈ 0.05(0.5/h)2 is wrong, or (d) that the fair sample hypothesis is wrong [51].
Regarding (b), it is interesting that there are indications from weak lensing [52]
that at least some clusters may actually have extended halos of dark matter --
something that is expected to a greater extent if the dark matter is a mixture
of cold and hot components, since the hot component clusters less than the
cold [53, 54].
If so, the number density of clusters as a function of mass is
higher than usually estimated, which has interesting cosmological implications
(e.g., σ8 is a little higher than usually estimated). It is of course possible that the
solution is some combination of alternatives (a)-(d). If none of the alternatives
is right, then the only conclusion left is that Ω0 ≈ 0.33. The cluster baryon
problem is clearly an issue that deserves very careful examination.
It has recently been argued [55] that CHDM models are compatible with
7
the X-ray data within observational uncertanties of both the BBN predictions
Indeed, the rather high baryon fraction Ωb ≈ 0.1(0.5/h)2
and X-ray data.
implied by recent measurements of low D/H in two high-redshift Lyman limit
systems [56, 57] helps resolve the cluster baryon crisis for all Ω = 1 models --
it is escape route (c) above. With the higher Ωb implied by the low D/H, there
is now a "baryon cluster crisis" for low-Ω0 models! Even with a baryon fraction
at the high end of observations, fb <∼ 0.2(h/0.5)−3/2, the fair sample hypothesis
with this Ωb implies Ω0 >∼ 0.5(h/0.5)−1/2.
4 Warm Dark Matter vs. CHDM
It will be instructive to look briefly at Warm Dark Matter (WDM), both to see
that some variants of CDM have less success than others in fitting cosmological
observations, and also because there is renewed interest in WDM. Although
CHDM and WDM are similar in the sense that both are intermediate models
between CDM and HDM, CHDM and WDM are quite different in their impli-
cations.
The problems with a pure Hot Dark Matter (HDM) adiabatic cosmology are
well known: free-streaming of the hot dark matter completely destroys small-
scale fluctuations, so that the first structures that can form are on the mass
scale of clusters or superclusters, and galaxies must form by fragmentation of
these larger structures; but observations show that galaxies are much older than
superclusters, which have low overdensity and are still forming. Moreover, with
the COBE upper limit to the normalization of HDF, hardly any structure at all
will form even by the present epoch.
WDM is a simple modification of HDM, obtained by changing the assumed
average number density n of the particles. In the usual HDM, the dark mat-
ter particles are neutrinos, each species of which has nν = 108 cm−3, with a
corresponding mass of m0 = Ωνρ0/nν = Ων92h2 eV, with Ων = 1 − Ωb ≈ 1.
In WDM, there is a new parameter, m/m0, the ratio of the mass of the warm
particle to the above neutrino mass; correspondingly, the number density of
the warm particles is reduced by the inverse of this factor, so that their total
contribution to the cosmological density is unchanged. Pagels and I [58] pro-
posed perhaps the first WDM particle candidate, a light gravitino, which was
the lightest supersymmetric particle (LSP) in the earliest version of supersym-
metric phenomenology (which was subsequently largely abandoned in favor of a
hidden-sector sypersymmetry breaking scheme, but which is now being recon-
sidered: cf. [59]). Olive & Turner [60] proposed left-handed neutrinos as WDM.
In both cases, these particles interact much more weakly than neutrinos, de-
couple earlier from the hot big bang, and thus have diluted number density
compared to neutrinos since they do not share in the entropy released by the
subsequent annihilation of species such as quarks. This is analogous to (but
more extreme than) the neutrinos themselves, which have lower number density
8
today than photons because the neutrinos decouple before e+e− annihilation
(and also because they are fermions).
In order to investigate the cosmological implications of any dark matter can-
didate, it is necessary to work out the gravitational clustering of these particles,
first in linear theory, and then after the amplitude of the fluctuations grows into
the nonlinear regime. Colombi, Dodelson, & Widrow [61] recently did this for
WDM, and Figure 1 in their paper compares the square of the linear transfer
functions for WDM and CHDM.
One often can study large scale structure just on the basis of such linear
calculations, without the need to do computationally expensive simulations of
the non-linear gravitational clustering. Such studies have shown that matching
the observed cluster and galaxy correlations on scales of about 20-30 h−1 Mpc
in CDM-type theories requires that the "Excess Power" EP ≈ 1.3, where
EP ≡
σ(25 h−1 Mpc)/σ(8 h−1 Mpc)
[σ(25 h−1 Mpc)/σ(8 h−1 Mpc)]sCDM
,
(3)
and as usual σ(r) = (δρ/ρ)(r) is the rms fluctuation amplitude in randomly
placed spheres of radius r. The EP parameter was introduced in the COBE-
DMR interpretation paper [62], and in a recent paper [24] we have shown that
EP is related to the spectrum shape parameter Γ [64] by Γ ≈ 0.5(EP )−3.3.
For CHDM and other models, this is a useful generalization since the cluster
correlations do seem to be a function of this generalized Γ, with Γ ≈ 0.23 to
match cluster correlation data [24]. Peacock & Dodds [63] have shown that
Γ ≈ 0.23 also is required to match large scale galaxy clustering data.
Since calculating σ(r) is a simple matter of integrating the power spectrum
times the top-hat window function,
σ(r) = Z ∞
0
P (k)W (kr)k2dk
(4)
the linear calculations immediately allow determination of EP for WDM and
CHDM. Ref. [61] shows that for WDM to give the required EP , the parameter
value m/m0 ≈ 1.5 − 2, while for CHDM the required value of the CHDM
parameter is Ων ≈ 0.3. But for WDM with m/m0 >∼ 2, the spectrum lies
a lot lower than the CDM spectrum at k >∼ 0.3h−1 Mpc (length scales λ <∼
20 h−1 Mpc), which in turn implies that formation of galaxies, corresponding to
the gravitational collapse of material in a region of size ∼ 1 Mpc, will be strongly
suppressed compared to CDM. Thus WDM will not be able to accommodate
simultaneously the distribution of clusters and galaxies. But CHDM will do
much better.
Probably the only way to accommodate WDM in a viable cosmological model
is as part of a mixture with hot dark matter, which might even arise naturally
in a supersymmetric model [65] of the sort in which the gravitino is the LSP
[59]. Cold plus "volatile" dark matter is a related possibility [66].
9
There are many more parameters needed to describe the presently available
data on the distribution of galaxies and clusters and their formation history
than the few parameters needed to specify a CDM-type model. Thus it should
not be surprising that at most a few CDM variant theories can fit all this data.
Once it began to become clear that standard CDM was likely to have problems
accounting for all the data, after the discovery of large-scale flows of galaxies
was announced in early 1986 [67], I advised Jon Holtzman in his dissertation
research to work out the linear theory for a wide variety of CDM variants [21]
so that we could see which ones would best fit the data [22]. The clear winners
were CHDM with Ων ≈ 0.3 if h ≈ 0.5, and ΛCDM with Ω0 ≈ 0.2 if h ≈ 1.
Variants of both these models remain perhaps the best bets still.
5 CHDM: Early Structure Troubles?
Aside from the possibility mentioned at the outset that the Hubble constant
is too large and the universe too old for any Ω = 1 model to be viable, the
main potential problem for CHDM appears to be forming enough structure
at high redshift. Although, as I mentioned above, the prediction of CHDM
that the amount of gas in damped Lyman α systems is starting to decrease at
high redshift z >∼ 3 seems to be in accord with the available data, the large
velocity spread of the associated metal-line systems may indicate that these
systems are more massive than CHDM would predict (see e.g., [68, 69]). Also,
results from a recent CDM hydrodynamic simulation [70] in which the amount
of neutral hydrogen in protogalaxies seemed consistent with that observed in
damped Lyman α systems led the authors to speculate that CHDM models
would produce less than enough; however, since the regions identified as damped
Lyman α systems in the simulations were not actually resolved, this will need
to be addressed by higher resolution simulations for all the models considered.
Finally, Steidel et al. [71] have found objects by their emitted light at red-
shifts z = 3 − 3.5 apparently with relatively high velocity dispersions (indicated
by the equivalent widths of absorption lines), which they tentatively identify
as the progenitors of giant elliptical galaxies. Assuming that the indicated ve-
locity dispersions are indeed gravitational velocities, Mo & Fukugita (MF) [72]
have argued that the abundance of these objects is higher than expected for the
COBE-normalized Ω = 1 CDM-type models that can fit the low-redshift data,
including CHDM, but in accord with predictions of the ΛCDM model consid-
ered here.
(In more detail, the MF analysis disfavors CHDM with h = 0.5
and Ων >∼ 0.2 in a single species of neutrinos. They apparently would argue
that this model is then in difficulty since it overproduces rich clusters -- and
if that problem were solved with a little tilt np ≈ 0.9, the resulting decrease in
fluctuation power on small scales would not lead to formation of enough early
objects. However, if Ων ≈ 0.2 is shared between two species of neutrinos, the
resulting model appears to be at least marginally consistent with both clusters
10
and the Steidel objects even with the assumptions of MF. The ΛCDM model
with h = 0.7 consistent with the most restrictive MF assumptions has Ω0 >∼ 0.5,
hence t0 <∼ 12 Gyr. ΛCDM models having tilt and lower h, and therefore more
consistent with the small-scale power constraint discussed above, may also be
in trouble with the MF analysis.) But in addition to uncertainties about the
actual velocity dispersion and physical size of the Steidel et al. objects, the
conclusions of the MF analysis can also be significantly weakened if the grav-
itational velocities of the observed baryons are systematically higher than the
gravitational velocities in the surrounding dark matter halos, as is perhaps the
case at low redshift for large spiral galaxies [73], and even more so for elliptical
galaxies which are largely self-gravitating stellar systems in their central regions.
Given the irregular morphologies of the high-redshift objects seen in the
Hubble Deep Field [74] and other deep HST images, it seems more likely that
they are relatively low mass objects undergoing starbursts, possibly triggered by
mergers, rather than galactic protospheroids. Since the number density of the
brightest of such objects may be more a function of the probability and duration
of such starbursts rather than the nature of the underlying cosmological model,
it may be more useful to use the star formation or metal injection rates [75]
indicated by the total observed rest-frame ultraviolet light to constrain models
[76]. The available data on the history of star formation [77, 78, 75] suggests that
most of the stars and most of the metals observed formed relatively recently,
after about redshift z ∼ 1; and that the total star formation rate at z ∼ 3
is perhaps a factor of 3 lower than at z ∼ 3, with yet another factor of ∼ 3
falloff to z ∼ 4 (although the rates at z >∼ 3 could be higher if most of the
star formation is in objects too faint to see). This is in accord with indications
from damped Lyman α systems [79] and expectations for Ω = 1 models such as
CHDM, but not with the expectations for low-Ω0 models which have less growth
of fluctuations at recent epochs, and therefore must form structure earlier. But
this must be investigated using more detailed modelling, including gas cooling
and feedback from stars and supernovae [76], before strong conclusions can be
drawn.
There is another sort of constraint from observed numbers of high-redshift
protogalaxies that would appear to disfavor ΛCDM. The upper limit on the
number of z >∼ 4 objects in the Hubble Deep Field (which presumably corre-
spond to smaller-mass galaxies than most of the Steidel objects) is far lower
than the expectations in low-Ω0 models, especially with a positive cosmological
constant, because of the large volume at high redshift in such cosmologies [80].
Thus evidence from high-redshift objects cuts both ways, and it is too early to
tell whether high- or low-Ω0 models will ultimately be favored.
11
6 Advantages of Mixed CHDM Over Pure CDM
Models
There are three basic reasons why a mixture of cold plus hot dark matter works
better than pure CDM without any hot particles: (1) the power spectrum shape
P (k) is a better fit to observations, (2) there are indications from observations
for a more weakly clustering component of dark matter, and (3) a hot compo-
nent may help avoid the too-dense central dark matter density in pure CDM
dark matter halos. I will discuss each in turn.
(1) Spectrum shape. As I explained in discussing WDM vs. CHDM
above, the pure CDM spectrum P (k) does not fall fast enough on the large-k
side of its peak in order to fit indications from galaxy and cluster correlations and
power spectra. The discussion there of "Excess Power" is a way of quantifying
this. This is also related to the overproduction of clusters in pure CDM. The
obvious way to prevent Ω = 1 sCDM normalized to COBE from overproducing
clusters is to tilt it a lot (the precise amount depending on how much of the
COBE fluctuations are attributed to gravity waves, which can be increasingly
important as the tilt is increased). But a constraint on CDM-type models
that is likely to follow both from the high-z data just discussed and from the
preliminary indications on cosmic microwave anisotropies at and beyond the
first acoustic peak from the Saskatoon experiment [81] is that viable models
cannot have much tilt, since that would reduce too much both their small-
scale power and the amount of small-angle CMB anisotropy. As I have already
explained, by reducing the fluctuation power on cluster scales and below, COBE-
normalized CHDM naturally fits both the CMB data and the cluster abundance
without requiring much tilt. The need for tilt is further reduced if a high
baryon fraction Ωb >∼ 0.1 is assumed [82], and this also boosts the predicted
height of the first acoustic peak. No tilt is necessary for Ων = 0.2 shared
between Nν = 2 neutrino species with h = 0.5 and Ωb = 0.1.
Increasing
the Hubble parameter in COBE-normalized models increases the amount of
small-scale power, so that if we raise the Hubble parameter to h = 0.6 keeping
Ων = 0.2 and Ωb = 0.1(0.5/h)2 = 0.069, then fitting the cluster abundance
in this Nν = 2 model requires tilt 1 − np ≈ 0.1 with no gravity waves (i.e.,
T /S = 0; alternatively if T /S = 7(1 − np) is assumed, about half as much tilt
is needed, but the observational consequences are mostly very similar, with a
little more small scale power). The fit to the small-angle CMB data is still good,
and the predicted Ωgas in damped Lyman α systems is a little higher than for
the h = 0.5 case. The only obvious problem with h = 0.6 applies to any Ω = 1
model -- the universe is rather young: t0 = 10.8 Gyr.
(2) Need for a less-clustered component of dark matter. The fact
that group and cluster mass estimates on scales of ∼ 1 h−1 Mpc typically give
values for Ω around 0.1-0.2 [84], while larger-scale estimates give larger values
around 0.3-1 [2] suggests that there is a component of dark matter that does
12
not cluster on small scales as efficiently as cold dark matter is expected to do.
In order to quantify this, my colleagues and I have performed the usual group
M/L measurement of Ω0 on small scales in "observed" Ω = 1 simulations of both
CDM and CHDM [83]. We found that COBE-normalized Ων = 0.3 CHDM gives
ΩM/L = 0.12 − 0.18 compared to ΩM/L = 0.15 for the CfA1 catalog analyzed
exactly the same way, while for CDM ΩM/L = 0.34 − 0.37, with the lower
value corresponding to bias b = 1.5 and the higher value to b = 1 (still below
the COBE normalization). Thus local measurements of the density in Ω = 1
simulations can give low values, but it helps to have a hot component to get
values as low as observations indicate. We found that there are three reasons
why this virial estimate of the mass in groups misses so much of the matter in the
simulations: (1) only the mass within the mean harmonic radius rh is measured
by the virial estimate, but the dark matter halos of groups continue their roughly
isothermal falloff to at least 2rh, increasing the total mass by about a factor of 3
in the CHDM simulations; (2) the velocities of the galaxies are biased by about
70% compared to the dark matter particles, which means that the true mass
is higher by about another factor of 2; and (3) the groups typically lie along
filaments and are significantly elongated, so the spherical virial estimator misses
perhaps 30% of the mass for this reason. Our visualizations of these simulations
[53] show clearly how extended the hot dark matter halos are. An analysis of
clusters in CHDM found similar effects, and suggested that observations of the
velocity distributions of galaxies around clusters might be able to discriminate
between pure cold and mixed cold + hot models [54]. This is an area where
more work needs to be done -- but it will not be easy since it will probably
be necessary to include stellar and supernova feedback in identifying galaxies in
simulations, and to account properly for foreground and background galaxies in
observations.
(3) Preventing too dense centers of dark matter halos. Flores and
I [85] pointed out that dark matter density profiles with ρ(r) ∝ r−1 near the
origin from high-resolution dissipationless CDM simulations [87] are in serious
conflict with data on dwarf spiral galaxies (cf. also Ref.
[86]), and in possible
conflict with data on larger spirals [88] and on clusters (cf. [89, 90]). Navarro,
Frenk, & White [73] agree that rotation curves of small spiral galaxies such as
DDO154 and DDO170 are strongly inconsistent with their universal dark mat-
ter profile ρN F W (r) ∝ 1/[r(r + a)2]. I am at present working with Stephane
Courteau, Sandra Faber, Ricardo Flores, and others to see whether ρN F W is
consistent with data from high- and low-surface-brightness galaxies with mod-
erate to large circular velocities are consistent with this universal profile. The
failure of simulations to form cores as observed in dwarf spiral galaxies either
is a clue to a property of dark matter that we don't understand, or is telling
us the simulations are inadequate. It is important to discover whether this is a
serious problem, and whether inclusion of hot dark matter or of dissipation in
the baryonic component of galaxies can resolve it. It is clear that including hot
dark matter will decrease the central density of dark matter halos, both because
13
the lower fluctuation power on small scales in such models will prevent the early
collapse that produces the highest dark matter densities, and also because the
hot particles cannot reach high densities because of the phase space constraint
[91, 54]. But this may not be enough.
7 Best Bet CDM-Type Models
As I said at the outset, I think CHDM is the best bet if Ω0 turns out to be near
unity and the Hubble parameter is not too large, while ΛCDM is the best bet
if the Hubble parameter is too large to permit the universe to be older than its
stars with Ω = 1.
Both theories do seem less "natural" than sCDM. But although sCDM won
the beauty contest, it doesn't fit the data. CHDM is just sCDM with some light
neutrinos. After all, we know that neutrinos exist, and there is experimental
evidence -- admittedly not yet entirely convincing -- that at least some of these
neutrinos have mass, possibly in the few-eV range necessary for CHDM.
Isn't it an unnatural coincidence to have three different sorts of matter --
cold, hot, and baryonic -- with contributions to the cosmological density that
are within an order of magnitude of each other? Not necessarily. All of these
varieties of matter may have acquired their mass from (super?)symmetry break-
ing associated with the electroweak phase transition, and when we understand
the nature of the physics that determines the masses and charges that are just
adjustable parameters in the Standard Model of particle physics, we may also
understand why Ωc, Ων, and Ωb are so close. In any case, CHDM is certainly
not uglier than ΛCDM.
In the ΛCDM class of models, the problem of too much power on small
scales that I discussed at some length for Ω0 = 0.3 and h = 0.7 ΛCDM implies
either that there must be some physical mechanism that produces strong, scale-
dependent anti-biasing of the galaxies with respect to the dark matter, or else
that higher Ω0 and lower h are preferred, with a significant amount of tilt to get
the cluster abundance right and avoid too much small-scale power [12]. Higher
Ω0 >∼ 0.5 also is more consistent with the evidence summarized above against
large ΩΛ and in favor of larger Ω0, especially in models such as ΛCDM with
Gaussian primordial fluctuations. But then h <∼ 0.63 for t0 >∼ 13 Gyr.
Among CHDM models, having Nν = 2 species share the neutrino mass gives
a better fit to COBE, clusters, and small-scall data than Nν = 1, and moreover
it appears to be favored by the available experimental data [8]. But it remains
to be seen whether CHDM models can fit the data on structure formation at
high redshifts.
Acknowledgments. This work was partially supported by NASA and NSF
grants at UCSC. I thank my collaborators, especially Anatoly Klypin, for many
helpful discussions of the material presented here.
14
References
[1] G.R. Blumenthal, S. Faber, J.R. Primack, & M.J. Rees 1994, Nature, 311,
517.
[2] A. Dekel 1994, Ann. Rev. Astron. Astroph., 32, 371; and Dekel & Burstein,
these proceedings.
[3] S.D.M. White, G. Efstathiou, & C.S. Frenk 1993, MNRAS, 262, 1023.
[4] E.g., S. Olivier, J. Primack, G.R. Blumenthal, & A. Dekel 1993, ApJ 408,
17.
[5] C.M. Baugh & G. Efstathiou 1994, MNRAS, 267, 32; see also Efstathiou's
contribution in this volume.
[6] S. Zaroubi, A. Dekel, Y. Hoffman, T. Kolatt 1996, astro-ph/9603068; cf.
T. Kolatt & A. Dekel 1966, astro-ph/9512132, ApJ submitted, and Dekel's
contribution in this volume.
[7] J.R. Primack & A. Klypin 1996, in Proc. Internat. Conf. on Sources and
Detection of Dark Matter in the Universe, UCLA, February 1996, D. Cline
and D. Sanders, eds, Nucl. Phys. B Proc. Suppl., in press.
[8] J.R. Primack, J. Holtzman, A. Klypin, & D.O. Caldwell, 1995, Phys. Rev.
Lett., 74, 2160.
[9] D. Pogosyan, & A.A. Starobinsky 1995, ApJ, 447, 465; A. Liddle, D.H.
Lyth, R.K. Schaefer, Q. Shafi, & P.T.P. Viana 1996, MNRAS, 281, 531,
and references therein.
[10] S. Ghigna, S. Borgani, S. Bonometto, L. Guzzo, A. Klypin, J.R. Primack,
R. Giovanelli, & M. Haynes 1994, ApJ, 437, L71.
[11] S. Ghigna, S. Borgani, M. Tucci, S. Bonometto A. Klypin, & J.R. Primack
1996, ApJ, submitted.
[12] A. Klypin, J.R. Primack, & J. Holtzman, ApJ, 466, 1.
[13] G.M. Fuller, J.R. Primack, & Y.-Z. Qian 1995, Phys. Rev. D, 52, 1288.
[14] C. Athanassopoulos et al. 1996, nucl-ex/9605001, Phys. Rev. submitted,
and nucl-ex/9605003, Phys. Rev. Lett., submitted (also available at http:
//nu1.lampf.lanl.gov/ lsnd/). Using only the data for Eν = 36 − 60 MeV,
for which the background is lower than the larger neutrino energy range
used in last figure in these preprints, lowers the lower limit on sin2 2θ and
thus increases the range of allowed ∆m2
µe (private communications from D.
Caldwell & S. Yellin, May 1996; D.O. Caldwell, in Neutrino'96, Helsinki,
June 1996).
15
[15] P.J.E. Peebles, Physical Cosmology (Princeton University Press, 1993), eq.
(5.150).
[16] D.O. Caldwell 1995, in Trends in Astroparticle Physics, Stockholm, Sweden
22-25 September 1994, eds. L. Bergstrom, P. Carlson, P.O. Hulth, & N.
Snellman, Nucl. Phys. B, Proc. Suppl., 43, 126.
[17] Y. Fukuda, et al. 1994, Phys. Lett. B, 280, 146.
[18] The Kamiokande data is consistent with atmospheric νµ oscillating to any
other neutrino species with a large mixing angle. But (see discussion and
references in, e.g., [8, 13]) νµ oscillating to νe with a large mixing angle
is probably inconsistent with reactor and other data, and νµ oscillating to
a sterile neutrino νs (i.e., one that does not interact via the usual weak
interactions) with a large mixing angle is inconsistent with the usual Big
Bang Nucleosynthesis constraints.
[19] Y. Suzuki, at Neutrino'96, Helsinki, June 1996, and private communication
June 1996.
[20] D. Pogosyan, & A.A. Starobinsky 1995, ApJ, 447, 465; A. Liddle, et al.,
Ref. [9].
[21] J. Holtzman 1989, ApJS, 71, 1.
[22] J. Holtzman & J.R. Primack 1993, ApJ, 405, 428.
[23] D. Pogosyan, & A.A. Starobinsky 1995, astro-ph/9502019; K.S. Babu, R.K.
Schaefer, & Q. Shafi 1996, Phys. Rev. D, 53, 606.
[24] S. Borgani, L. Moscardini, M. Plionis, K.M. G´orski, J. Holtzman, A.
Klypin, J.R. Primack, C.L. Smith, & R. Stompor 1996, submitted to New
Astronomy.
[25] M. Davis, F. Summers, & D. Schlegel 1992, Nature, 359, 393.
[26] A. Klypin, J. Holtzman, J.R. Primack, & E. Regos 1993, ApJ, 416, 1.
[27] H.J. Mo & J. Miralda-Escude 1994, ApJ, 430, L25; G. Kauffmann & S.
Charlot 1994, ApJ, 430, L97; C.-P. Ma & E. Bertschinger 1994, ApJ, 434,
L5.
[28] A. Klypin, S. Borgani, J. Holtzman, & J.R. Primack 1995, ApJ, 444, 1.
[29] C.-P. Ma 1995, in Dark Matter, AIP Conference Proceedings 336, p. 420.
[30] L.J. Storrie-Lombardi, R.G. McMahon, & M.J.
Irwin 1996, astro-
ph/9608147, MNRAS, in press. Cf. Art Wolfe's contribution in these pro-
ceedings.
16
[31] Chaboyer, B., Demarque, P., Kernan, P.J., & Krauss, L.M. 1996, Science,
271, 957.
[32] T.D. Oswalt, J.A. Smith, & M.A. Wood 1996, Nature, 382, 692.
[33] O. Lahav, P. Lilje, J.R. Primack, & M.J. Rees 1991, MNRAS, 251, 128.
[34] S.M. Carroll, W.H. Press, & E.L. Turner 1992, Ann. Rev. Astron. Astro-
phys, 30, 499.
[35] P.J.E. Peebles 1984, ApJ, 284, 439.
[36] G. Efstathiou, W.J. Sutherland & S.J. Maddox 1990, Nature 348, 705; L.A.
Kofman, N.Y. Gendin, & N.A. Bahcall 1993, ApJ, 413, 1; R. Cen, N.Y.
Gendin & J.P. Ostriker 1993, ApJ, 417, 387; R.A. Croft & G. Efstathiou
1994, MNRAS, 267, 390; J.P. Ostriker & P.J. Steinhardt 1995, Nature, 377,
600; L.M. Krauss, & M.S. Turner 1995, General Relativity & Gravitation,
27, 1137.
[37] G. Kauffmann, A. Nusser, & M. Steinmetz, astro-ph/9512009.
[38] A.R. Liddle, D.H. Lyth, P.T.P. Viana, & M. White, 1996, MNRAS, 282,
281.
[39] J.R. Primack, astro-ph/9604184, to appear in International School of
Physics "Enrico Fermi", Course CXXXII: Dark Matter in the Universe,
Varenna, eds. S. Bonometto, J.R. Primack, & A. Provenzale (IOS Press,
Amsterdam, in press). A popular version is J. Roth & J.R. Primack 1996,
Sky & Telescope, 91(1), 20.
[40] C. Kochanek 1996, ApJ, 466, 638.
[41] S.P. Driver, R.A. Windhorst, S. Phillipps, & P.D. Bristow 1996, ApJ 461,
525.
[42] S. Perlmutter et al 1996., astro-ph/9602122, to appear in Thermonuclear
Supernovae (NATO ASI), eds. R. Canal, P. Ruiz-LaPuente, & J. Isern.
[43] S. Perlmutter et al. 1996, astro-ph/9608192, submitted to ApJ.
[44] G. Yepes, R. Kates, A. Khokhlov & A. Klypin 1996, astro-ph/9605182.
[45] A. Klypin, R. Nolthenius, & J.R. Primack 1995, astro-ph/9502062, ApJ,
in press (Jan 10, 1997).
[46] S.D.M. White, & C.S. Frenk 1991, ApJ, 379, 52; S.D.M. White et al. 1993,
Nature, 366, 429.
[47] D.A. White, & A.C. Fabian 1995, MNRAS, 273, 72.
17
[48] K.F. Gunn & P.A. Thomas 1995, astro-ph/9510082.
[49] C. Balland & A. Blanchard 1995, astro-ph/9510130.
[50] N.A. Bahcall & L.M. Lubin 1994, ApJ, 426, 513; M. Bartelmann & R.
Narayan 1995, in Dark Matter, AIP Conference Proceedings 336, p. 307.
[51] M. Lowenstein & R.F. Mushotzky 1996, astro-ph/9608111, discuss two poor
clusters with similar total mass distributions but baryon fractions differing
by a factor ∼ 2.
[52] G. Squires, N. Kaiser, A. Babul, G. Fahlman, D. Woods, M. Neumann, &
H. Bohringer 1996, ApJ, 461, 572. Cf. G. Luppino, M. Lowenstein, in Proc.
Internat. Conf. on Sources and Detection of Dark Matter in the Universe,
UCLA, February 1996, D. Cline & D. Sanders, eds, Nucl. Phys. B Proc.
Suppl., in press.
[53] D. Brodbeck, D. Hellinger, R. Nolthenius, J.R. Primack, & A. Klypin 1997,
ApJ, in press (with accompanying video). Still and Mpeg excerpts are avail-
able from Primack web page at http: //physics.ucsc.edu
[54] L. Kofman, A. Klypin, D. Pogosyan, & J.P. Henry 1995, astro-ph/9509145.
[55] R.W. Strickland & D.N. Schramm, astro-ph/9511111.
[56] D. Tytler, X.-M. Fan & S. Burles 1996, Nature, 381, 207; S. Burles &
D. Tytler 1996, astro-ph/9603070; D. Tytler & S. Burles 1996, astro-
ph/9606110, in Proc. Origin of Matter and Evolution of Galaxies in the
Universe, in press.
[57] See Hogan's contribution to this volume for a dissenting view. The only
high-redshift deuterium detections that are secure are those by Tytler et
al., but Hogan questions whether Tytler et al. might have overestimated
the hydrogen content in the same systems, and thereby underestimated
D/H; Tytler & Burles answer this objection in detail in the last ref. in [56].
[58] H. Pagels & J.R. Primack 1982, Phys. Rev. Lett., 48, 223.
[59] M. Dine, A. Nelson, Y. Nir, & Y. Shirman 1996, Phys. Rev. D53, 2658; S.
Dimopoulos, M. Dine, S. Raby, & S. Thomas 1996, Phys. Rev. Lett., 76,
3494.
[60] K.A. Olive & M.S. Turner 1982, Phys. Rev. D25, 213.
[61] S. Colombi, S. Dodelson, & L.M. Widrow 1996, ApJ, 458, 1.
[62] E.L. Wright et al. 1992, ApJ, 396, L13.
18
[63] J.A. Peacock & S.J. Dodds 1994, MNRAS, 1020; -- 1996, MNRAS, 280,
L19.
[64] G. Efstathiou, J.R. Bond, & S.D.M. White 1992, MNRAS, 258, 1p. Cf.
J.M. Bardeen, J.R. Bond, N. Kaiser, & A.S. Szalay 1986, ApJ, 304, 15. For
CDM and the ΛCDM family of models, Γ = Ωh.
[65] S. Borgani, A. Masiero, & M. Yamaguchi 1996, hep-ph/9605222.
[66] E. Pierpaoli, P. Coles, S. Bonometto, & S. Borgani 1996, astro-ph/9603150,
ApJ, in press, and references therein.
[67] D. Burstein 1986, in Galaxy Distances and Deviations from Universal Ex-
pansion, eds B.F. Madore & R.B. Tully, Dordrecht.
[68] L. Lu, W.L.W. Sargent, D.S. Womble, & T.A. Barlow 1996, ApJ, 457, L1.
[69] A. Wolfe, these proceedings. Note however that the effects of dust extinction
(cf. Y. Pei & S.M. Fall 1995, ApJ, 454, 69; M. Pettini et al. 1996, astro-
ph/9607093) and gravitational lensing (M. Bartelmann & A. Loeb 1996,
ApJ, 457, 529) should be taken into account in modelling damped Lyman
α systems.
[70] N. Katz, D.H. Weinberg, L. Hernquist, & J. Miralda-Escude 1996, ApJ,
457, L57.
[71] C.A. Steidel et al. 1996, ApJ, 462, L17.
[72] H.J. Mo & M. Fukugita 1996, ApJ, 467, L9.
[73] J.F. Navarro, C.S. Frenk, & S.D.M. White 1996, ApJ, 462, 563. Also, D.
Zaritsky & S.D.M. White 1994, ApJ, 435, 599 (and work in preparation)
find little correlation between the rotation velocity of optical galaxies and
that of their extended dark matter halos. I thank M. Steinmetz for empha-
sizing the relevance of these facts.
[74] S. van den Bergh et al. 1996, AJ, 112, 359; R.G. Abraham et al. 1996,
MNRAS, 279, L47.
[75] P. Madau et al. 1996, astro-ph/9607172.
[76] R. Somerville and I are doing this for all currently popular models, using a
semi-analytic merging hierarchy model of the sort pioneered by G. Kauff-
mann, S.D.M. White, & B. Guiderdoni, 1993, MNRAS, 264, 201. We have
related work underway with G. Larsen and S.M. Fall.
[77] J. Gallego, J. Zamorano, A. Aragon-Salamanca, & M. Rego 1996, ApJ,
455, L1.
19
[78] S.J. Lilly, O. Le Fevre, F. Hammer, & D. Crampton 1996, ApJ, 460, L1.
[79] S.M. Fall, S. Charlot, & Y.C. Pei 1996, ApJ, 464, L43.
[80] K.M. Lanzetta, A. Yahil, & A. Fern´andez-Soto 1996, Nature, 381, 759; A.
Yahil, K.M. Lanzetta, A. Campos, & A. Fern´andez-Soto 1996, in prepara-
tion.
[81] C.B. Netterfield et al. 1996, ApJ, 445, L69. Cf. J. Silk, these proceedings.
[82] M. White, P.T.P. Viana, A.R. Liddle, & D. Scott 1996, astro-ph/9605057.
[83] R. Nolthenius, A. Klypin, J. Primack 1997, ApJ, in press.
[84] N. Bahcall, in these proceedings.
[85] R. Flores & J.R. Primack 1994, ApJ, 427, L1.
[86] B. Moore 1994, Nature, 370, 629.
[87] J. Dubinski & R.G. Carlberg 1991, ApJ, 378, 496; M.S. Warren, P.J. Quinn,
J.K. Salmon, & W.H. Zurek 1992, ApJ, 399, 405; M. Crone, A. Evrard, &
D. Richstone 1994, ApJ, 434, 402.
[88] R. Flores, J.R. Primack, G.R. Blumenthal, & S.M. Faber 1993, ApJ, 412,
443.
[89] J. Miralda-Escud´e 1995, ApJ, 438, 514.
[90] R. Flores & J.R. Primack 1996, ApJ, 457, L5.
[91] S. Tremaine & J.E. Gunn 1979, Phys. Rev. Lett., 42, 407.
20
|
astro-ph/9905193 | 1 | 9905 | 1999-05-15T07:24:30 | The Luminosity Function of M3 | [
"astro-ph"
] | We present a high precision, large sample luminosity function (LF) for the Galactic globular cluster M3. With a combination of ground based and Hubble Space Telescope data we cover the entire radial extent of the cluster. The observed LF is well fit by canonical standard stellar models from the red giant branch (RGB) tip to below the main sequence turnoff point. Specifically, neither the RGB LF-bump nor subgiant branch LF indicate any breakdown in the standard models. On the main sequence we find evidence for a flat initial mass function and for mass segregation due to the dynamical evolution of the cluster. | astro-ph | astro-ph | The Luminosity Function of M31
R.T. Rood2, E. Carretta3, B. Paltrinieri4, F. R. Ferraro5,6
F. Fusi Pecci6,7, B. Dorman8,
A. Chieffi9, O. Straniero10, R. Buonanno11
ABSTRACT
We present a high precision, large sample luminosity function (LF) for the Galactic
globular cluster M3. With a combination of ground based and Hubble Space Telescope
data we cover the entire radial extent of the cluster. The observed LF is well fit by
canonical standard stellar models from the red giant branch (RGB) tip to below the main
sequence turnoff point. Specifically, neither the RGB LF-bump nor subgiant branch LF
indicate any breakdown in the standard models. On the main sequence we find evidence
for a flat initial mass function and for mass segregation due to the dynamical evolution
of the cluster.
Subject headings: globular clusters: individual (M3) -- stars: red giant -- -- stars: evo-
lution
9
9
9
1
y
a
M
5
1
1
v
3
9
1
5
0
9
9
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
1Based in part 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
2Astronomy Dept, University of Virginia, P.O.Box 3818, Charlottesville, VA 22903-0818; [email protected]
3Osservatorio Astronomico di Padova, vicolo dell'Osservatorio 5, 35122 Padova, ITALY; [email protected]
4Istituto di Astronomia-Universit´a "La Sapienza", via G.M. Lancisi 29, I-00161 Roma, Italy; [email protected]
5European Southern Observatory, Karl Schwarzschild Strasse 2, D-85748 Garching bei Munchen, Germany
6Osservatorio Astronomico di Bologna, via Ranzani 1, 40126 Bologna, ITALY; ferraro,[email protected]
7Stazione Astronomica di Cagliari, 09012 Capoterra, ITALY
8Laboratory for Astronomy & Solar Physics, Code 681, NASA/GSFC, Greenbelt MD 20771; [email protected]
9Istituto di Astrofisica Spaziale-CNR, 00044 Frascati, ITALY; [email protected]
10Osservatorio Astronomico di Collurania, 64100 Teramo, ITALY; [email protected]
11Osservatorio Astronomico di Roma, 00040 Monte Porzio, ITALY; [email protected]
1
1.
Introduction
One of the classical tests of stellar evolutionary cal-
culations is through the comparison of theoretical and
observed stellar luminosity functions (LFs) of globu-
lar clusters (e.g., Renzini & Fusi Pecci 1988). The
LF for the stars below the main sequence turn-off
(MSTO) can be related to the stellar initial mass func-
tion and be used as a probe of stellar dynamics within
a cluster. The LF of stars after the MSTO depends
primarily on the rate of evolution and provides a di-
rect and straight-forward test of evolutionary calcula-
tions. Basically the post-MSTO LF measures the de-
velopment of the hydrogen burning shell and its pro-
gression through the star. Features in the observed
LF can be related to interior structure. For example,
as first noted by Iben (1968), when the H-burning
shell advances through the composition discontinuity
left by the maximum penetration of convective enve-
lope, the luminosity briefly decreases. This leads to
the so-called LF bump. More subtle observables are
also present: the slope of the LF below and above the
LF bump are different because the H-burning shell is
in the first case passing through a region of varying
H abundance and later through a region of constant
H (Rood & Crocker 1985, Fusi Pecci et al. 1990). A
breakdown in canonical stellar evolution theory (Ren-
zini & Fusi Pecci 1988) can affect the LF in the region
of the subgiant branch (SGB).12 Examples include
extra energy transport by Weakly Interacting Mas-
sive Particles (WIMPs) (Faulkner & Swenson 1993)
or the gravitational settling of helium (Proffit & Van-
denBerg 1991, Castellani et al. 1997, Straniero et al.
1997), both of which affect the cluster age calibration.
To explore short evolutionary phases or to search
for the small changes in the LF produced by reason-
able modifications to the canonical models requires
large samples.
Indeed, to explore phenomena like
those cited above, such large samples are required
that photometry must extend to the crowded in-
ner parts of clusters. To be useful in constructing
12The canonical assumptions include the neglect of phenomena
like rotation, non-convective matter transport, etc.
in stellar
model calculations. Some of these are certainly present in real
stars, but they are very difficult to include in calculations. This
leads to two approaches to dealing with non-canonical phenom-
ena: one is to try to model the phenomenon ordinarily in a
simplistic way requiring the introduction of free parameters.
The other is to make the most stringent possible comparison
between observations and canonical models -- the approach we
adopt here.
a LF, photometric samples must be complete and
accurate -- a requirement which has come into conflict
with the necessity of working near cluster cores. Ad-
equate samples to identify the LF-bump were first
obtained for 47 Tuc by King, DaCosta & Demar-
que (1985) and for 11 other clusters by Fusi Pecci
et al. (1990) (see Cassisi & Salaris 1997 for a review).
At lower luminosities LF studies have produced some
puzzling results. Stetson (1991) in a LF formed from
combined data from the low metallicity clusters M68,
M92, and NGC 6397 found an excess of stars on
the SGB just above the MSTO. Bolte (1994) found
a similar result for M30 which has been confirmed
by Bergbusch (1996), Guhathakurta et al. (1998) and
Sandquist et al. (1999). This excess was especially
intriguing because that was the "signature" expected
if WIMP energy transport was important (Faulkner
& Swenson 1993). However Sandquist et al. (1996) in
the largest sample LF obtained to date did not find a
similar result for the somewhat more metal rich clus-
ter M5.
Clearly larger and better LFs are desirable, and
these are now possible with the Hubble Space Tele-
scope (HST ) which has the capability to obtain pho-
tometry near cluster centers. While very large HST
samples of fainter stars can be obtained, the small
field of view limits the number of evolved stars which
can be observed. However, the combination of ground
based photometry of the outer cluster and HST data
can yield a large sample LF which spans the range of
cluster stellar luminosities. In this paper we present
such an LF obtained for the cluster M3 (NGC 5272).
This is the largest, most precise, and most complete
LF yet obtained for a globular cluster.
2. HST-WFPC2 Data Sample
2.1. Observations and Data Analysis
The observations were obtained April 25, 1995 as
part of the HST program GO5496 (P.I. F. Fusi Pecci),
the first of our programs devoted to testing the the-
oretical models using luminosity functions of a set of
Galactic Globular Clusters with different metallicity.
A set of exposures was secured through UV (F255W
and F336W) and optical filters (F555W and F814W).
Preliminary results on UV data has been already
presented in Ferraro et al. (1997a) and Laget et al.
(1998); here we will report on results obtained using
only the V and I exposures. Total exposure times
were 400 sec in V (F555W) and 560 sec in I (F814W);
2
each exposure is listed in Table 1 by Ferraro et al.
(1997a).
The initial data analysis (including bias, dark and
flat-field corrections) was made through the standard
HST pipeline. A master unsharp mask image for
each filter was made using MIDAS to register all the
long-exposure images, yielding a median frame with
cosmic rays statistically eliminated. The search of the
individual star components was made on the V me-
dian frame, following the usual procedure with RO-
MAFOT (Buonanno et al. 1983). Then we performed
the PSF fitting on each individual frame separately
using a Moffat function plus a numerical map of the
residuals to better account for the contribution of the
stellar PSF wings. The average instrumental magni-
tudes for each star were transformed to the standard
Johnson V and I system using the Holtzmann et al.
(1995) recipes. The magnitudes thus obtained were
eventually shifted to match the independent zero-
point calibration adopted in Ferraro et al.
(1997b,
hereafter M3CCD). The V , V − I CMD for the HST
global sample is shown in Figure 1.
Special care has been devoted to ensure that the
effects of blends (and crowding) do not significantly
alter the derived LF. The effects of blending can be
quite problematic in CMDs of dense regions of glob-
ular clusters. On the main sequence, blended stars
can potentially be mistaken for binary stars, while on
the subgiant branch they be shifted erroneously to
brighter bins. Blends can mimic incompleteness by
shifting the two stars off the main loci such that they
are dropped from the LF. To minimize these effects
suspected blended objects with multiple stellar com-
ponents lying off the main branches of the CMD were
individually examined by eye, exploiting the fully in-
teractive capability of the package ROMAFOT. In
this way we were able to check and, if necessary, elim-
inate spurious objects. The degree to which blends
contaminate the final sample can be estimated from
the artificial stars tests described below. Some of the
artificial stars which were not recovered were lost as
blends. Thus the degree of incompleteness at a given
level serves as an upper limit to the blend-fraction.
2.2. Completeness
A crucial component in the calculation of an ac-
curate LF is the determination of corrections for in-
completeness.
In our data crowding is the primary
source of incompleteness. Moreover, because of the
large gradient in the spatial density of stars present
in our frame, the completeness at any level of magni-
tude must be determined as a function of the position
with respect to the cluster center.
In order to quantify the efficiency in detecting stars
as function of the magnitude and the degree of crowd-
ing we carried out extensive artificial star tests in the
V filter frames, following the procedure described be-
low.
1. Each chip was split into two annular regions cen-
tered on the cluster center:
INNER: r < 50′′ including the entire PC1 field
and the innermost part of the three WFC
fields;
OUTER: r > 50′′ covering the outer part of
the three WFCs.
M3 is a cluster of moderate central density.
With a core radius rc = 30′′ (Djorgovski 1993)
one would expect the stellar density to be ap-
proximately constant across the PC as is appar-
ent from Figure 2. However, the stellar density
has dropped appreciably at the outer parts of
the WFs. We have adopted 50′′ as a bound-
ary between regions of higher and lower pro-
jected stellar density. A map of the WFPC2
field of view is shown Figure 2 with the circle
(at r ∼ 50′′) delineating the two regions.
2. A set of artificial stars was generated using the
Tiny Tim package (Krist 1994), and these were
added to the original frames in random posi-
tions. In order to avoid a spurious crowding en-
hancement, the size of the added sample (about
300 total artificial images or ∼ 80 for each chip)
was ≤ 10% of the total number of stars origi-
nally found in each chip. Then the star search-
ing and fitting procedures were repeated on the
artifically enriched frames.
3. An artificial star was considered to be "recov-
ered" on the basis of the following criteria:
• the differences in position, ∆x, ∆y, be-
tween the input artificial star and the re-
covered one are ∆x, ∆y < 1 pixel;
• the difference in magnitude, ∆mag, is ∆mag <
0.3.
The fraction of artificial stars missed during any
step of the procedure is a good estimate of the
3
actual number of stars missed at each magni-
tude level during the reduction procedure.
and simply counting all the stars in each bin within
±5σV −I of the MRL.
4. The completeness factor (Φ) was finally com-
puted as the numbers of stars recovered (Nrec)
with respect to the number of stars simulated
(Nsim) during each trial. The final value of Φ
has been obtained as an average of 3 -- 6 inde-
pendent trials for each magnitude bin (0.2 mag
width) in each of the two regions (INNER and
OUTER).
The result of these tests is summarized in Fig-
ure 3a,b. The completeness at the MSTO level (V ∼
19) is > 90% in both the INNER and OUTER re-
gions, thus the star counts up to that level have been
corrected by less than 10%. The corrections for com-
pleteness were applied separately in each part (inner
or outer) of each field (PC, WFCs) so that the total
corrected sample properly accounts for the differing
degrees of crowding.
2.3. The Mean Ridge Line of the RGB, SGB,
and Upper MS
If there were no photometric errors all of the RGB,
SGB, and Upper MS stars would lie along a line in
the CMD. We approximate this line with a mean ridge
line (MRL) which has been determined by an iterative
procedure:
1. A first rough selection of stars belonging to
the RGB, SGB, MS was performed by eye, ex-
cluding the horizontal branch (HB), asymptotic
giant branch (AGB), and blue straggler stars
(BSS);
2. This preliminary sample was divided into 0.5
mag width bins;
3. For each bin the median V − I color was com-
puted;
4. All stars lying more than 5σV −I from the median
value was rejected.
Points (3) and (4) of the procedure were iterated
until the solution was stable. The resulting MRL
points are listed in Table 1. Only about 2% of the
sample was rejected in determining the MRL.
Luminosity functions were constructed for each
chip by dividing each individual CMD in 0.2 mag bins,
Figure 4 shows a zoom of luminosity function in
the Turn-Off region for the total HST sample, before
and after correction for completeness. Errors have
been computed accordingly to the following relation
(see also Bolte 1989):
σbin = [
N 1/2
Φ
+
N · σΦ
Φ2
]
(1)
where N is the number of stars observed in each bin,
Φ is the completeness factor, and σΦ is the associ-
ated error which was determined from the rms of the
repeated completeness trials.
3. Matching the Ground-Based Observations:
the Global LF
HST data presented here complement the large
data set we have obtained for M3 during the last
decades. The photographic sample was originally
presented by Buonanno et al. (1994) and the CCD
ground based data have been discussed by Ferraro
et al. (1993, 1997b). The entire photographic sample
extends over an area between 2′ and 7′ from the clus-
ter center. Here we consider a subsample (hereafter
referred to as PHOTO) of stars lying in the annulus
between 3.′5 < r < 6′ from the cluster center (region
AI in Buonanno et al. 1994). In this annulus all stars
down V ∼ 22 have been measured. The CCD data
cover a field of view 7′ × 7′ roughly centered on the
cluster center; stars in the innermost central regions
(r < 20′′) have been not measured. In this sample the
deeper exposures (V and I) only reached to V ∼ 20.
In order to ensure an high level (formally 100%) of
completeness we include only stars with V < 17 lying
in the annulus 100′′ -- 210′′ from the cluster center. We
will refer to this subsample as CCD. The area cov-
ered by the CCD subsample and the the others fields
in shown Figure 1 of Ferraro et al. (1997b).
In the PHOTO and CCD subsets (in analogy to the
HST sample) we applied the procedure described in
Section 2.2, selecting bona fide RGB-SGB-MS mem-
bers taking all stars within ±5σ in color from the
MRL in the V, B−V and V, V −I CMDs respectively.
The CMDs of all stars selected in the three samples
(PHOTO, CCD, HST respectively) are shown in Fig-
ure 5. For each of these samples we independently
derive the LF. Star counts have been corrected using
the respective completeness curve: the PHOTO sam-
4
V
HB
15.57
15.62
15.63
15.64
15.64
15.62
15.63
15.64
15.66
15.71
15.76
15.83
15.91
15.99
16.10
16.21
16.37
16.54
16.65
V − I
0.775
0.731
0.684
0.631
0.585
0.272
0.219
0.184
0.146
0.105
0.079
0.043
0.008
-0.021
-0.050
-0.069
-0.080
-0.088
-0.091
V
RGB
12.44
12.68
12.90
13.10
13.30
13.55
13.79
14.05
14.28
14.56
14.82
15.09
15.36
15.64
15.89
16.21
16.52
16.70
16.94
17.20
17.36
17.63
18.05
Table 1
Mean ridge line for the HST sample.
V − I
V
V − I
1.589
1.473
1.403
1.347
1.294
1.241
1.193
1.150
1.115
1.077
1.045
1.017
0.989
0.964
0.939
0.911
0.890
0.879
0.867
0.856
0.849
0.836
0.815
18.16
18.26
18.33
18.37
18.42
18.47
18.52
18.57
18.62
18.70
18.78
18.86
18.98
19.10
19.18
19.26
19.38
19.78
20.10
20.30
20.50
20.98
21.34
0.803
0.787
0.764
0.746
0.724
0.696
0.668
0.647
0.629
0.616
0.608
0.600
0.593
0.592
0.593
0.595
0.597
0.624
0.654
0.672
0.693
0.756
0.805
5
ple has been corrected using Table 7 by Buonanno et
al. (1994), the CCD sample (for V < 17) is virtually
complete, thus no correction has been applied, and
the HST sample has been corrected using the com-
pleteness curves plotted in Figure 3a,b. In construct-
ing the global LF we considered only the magnitude
range where the completeness was greater than 50%.
Thus the global LF is restricted to V ∼< 21 by the
most incomplete region in the HST sample -- the inner
region of the WF4 and WF2 cameras (Figure 3a,b).
Since the CCD sample spanned a different magni-
tude range than the others, we determined a scaling
factor to produce a proper match. We followed the
procedure explained in Bolte (1994):
let ni,k be the
number of stars, after the proper completeness cor-
rection, in bin i of the sub-sample k, where k = 1, 2, 3
indicates the PHOTO, the global HST , and the CCD
sub-samples, respectively. Thus, for k = 1, 2 the re-
sulting LF is simply given by the sum
N
′
i =
2
Xk=1
ni,k
in the range 12 < V < 21. However, in order to
include the CCD sub-sample (k = 3) we derived a
scaling factor s from the range in magnitude (12 <
V < 17.0) which is common to all sub-samples:
s = P2
k=1 ni,kPi ni,k
P3
k=1 ni,kPi ni,k
so that the number of stars for each bin in the com-
bined LF is
Ni = s × (N
′
i + ni,k=3).
Table 2 lists the final LF. Column 1 lists the mag-
nitude bins (which are 0.2 mag. in width); column 2
gives Ni, the total scaled number of objects per 0.2
mag; column 3 contains the errors in the Ni. Even
though it extends only 2 mag below the turnoff, this
LF, including more than 50000 stars, is the most pop-
ulated LF ever published for a galactic globular clus-
ter.
It is useful to compare our LF to that for M5 ob-
tained by Sandquist et al. (1996, hereafter SBSH), the
best previous globular cluster LF. Doing this, it is ap-
propriate to compare sample size in specific parts of
the CMD. Our "bright" sample which corresponds to
the upper 4.5 mag of the RGB (or V ≤ 17) contains
6
944 stars. For such bright stars SBSH were able to
work near the cluster center, and their sample in the
comparable RGB interval was about 1000 stars. For
fainter stars ground based photometry becomes in-
creasingly affected by problems of photometric blends
(see SBSH). Blends can be especially insidious when
they "move" a star from one part of the LF to an-
other, because this is just the sort of thing one would
expect in a breakdown of canonical stellar modeling.
To avoid blends only the sparsely populated outer
cluster can be used, and it is especially difficult to
get adequate samples of the SGB region. Here our
blend-free HST sample plays a crucial role. We have
6085 stars brighter than MSTO; SBSH have 3300. In
a 0.1 mag bin at turnoff we have 820 stars compared
to SBSH's 280.
4. Results
4.1. Preliminaries
There are certain issues which must be addressed
in comparing LFs either with other observational LFs
or with theoretical LFs. The first is how the samples
are normalized. Beyond this there is some interplay
between assumed distance modulus, age, and chemi-
cal composition.
4.1.1. Normalization
One must determine a "normalization" factor to
adjust for sample size.
In recent LF studies (e.g.
Stetson 1991, Bolte 1994, Sandquist et al. 1996) both
the lower RGB and the main sequence region have
been used for normalization. We adopt a different
approach normalizing to the total number of stars
brighter than the MSTO at V = 19.10. Operationally
this has the advantage that the turnoff is well defined
and because of the large number of stars involved the
normalization is not much affected by Poisson statis-
tics. The small imprecision in determining the loca-
tion of the turnoff (∼< 0.1 mag) produces only a very
small error (∼< 1%) in the normalization.
4.1.2. Chemical composition
For M3 Kraft et al. (1992), found [Fe/H] = −1.47,
based on detailed fine abundance analysis of high res-
olution spectra of red giant stars. However, their
study is not fully self-consistent since while stel-
lar analysis was performed using model atmospheres
from the grid of Bell et al. (1976, hereafter BEGN),
Luminosity function of the global sample (PHOTO+CCD+HST).
Table 2
V
12.70
12.90
13.10
13.30
13.50
13.70
13.90
14.10
14.30
14.50
14.70
14.90
15.10
15.30
15.50
15.70
15.90
16.10
16.30
16.50
16.70
16.90
17.10
17.30
Log(Ni)
σNi
0.70
0.30
0.78
0.85
0.95
0.85
0.90
0.90
1.11
1.08
1.32
1.15
1.34
1.38
1.76
1.60
1.51
1.65
1.70
1.79
1.85
1.95
2.04
2.06
0.45
0.26
0.45
0.41
0.52
0.50
0.50
0.47
0.61
0.55
0.67
0.58
0.77
0.72
0.92
0.87
0.75
0.83
0.93
0.96
1.00
1.06
1.09
1.11
V
17.50
17.70
17.90
18.10
18.30
18.50
18.70
18.90
19.10
19.30
19.50
19.70
19.90
20.10
20.30
20.50
20.70
20.90
21.10
21.30
21.50
21.70
21.90
Log(Ni)
σNi
2.05
2.22
2.26
2.46
2.72
2.97
3.06
3.18
3.25
3.32
3.37
3.42
3.45
3.49
3.53
3.56
3.56
3.59
3.58
3.58
3.55
3.57
3.66
1.10
1.20
1.23
1.35
1.52
1.68
1.75
1.83
1.88
1.94
1.98
2.01
2.04
2.08
2.12
2.14
2.16
2.19
2.21
2.24
2.27
2.35
2.48
7
the reference solar iron abundance was extracted from
the empirical Holweger and Muller (1974, hereafter
HM) solar model. This inconsistency was resolved by
Carretta & Gratton (1997) who use the Kurucz (1993)
grid of model atmospheres for both solar and stellar
analysis, giving [Fe/H] = −1.34. The offset with the
study of Kraft et al. (from which the observational
material was taken) is simply due to the fact that the
HM model is ∼ 150 K warmer than the BEGN models
in the line formation region, explaining the 0.13 dex
difference in the derived [Fe/H].
Oxygen in red giants can by affected by interior nu-
clear processing. Indeed, M3 is one of the several clus-
ters with an observed anti-correlation between O and
Na along the RGB (see for instance Kraft et al. 1992).
This anti-correlation arises because of non-standard
mixing on the upper RGB. Therefore, the best esti-
mate for the [O/Fe] ratio is that obtained considering
only unmixed stars, i.e. giants with [Na/Fe] ≈ 0.0
dex. Using Kraft et al. (1992) we estimate that
[O/Fe] = 0.19 ± 0.04 dex based on 5 stars.
One typically assumes that the other α-elements
(e.g. Kraft et al. 1993) have the same enhancement
as O and uses two metallicity parameters, [Fe/H] and
[α/Fe], in computing models. For many problems one
can get by with just one metallicity parameter using
the short cut of Salaris et al. (1993). This makes use of
the global metallicity ([M /H]) based on the mass frac-
tion of all elements heavier than helium as compared
to solar. They found that scaled solar models with
[Fe/H] = [M/H] mimicked models based on observed
[Fe/H] with α-enhancement. For M3 [M/H] = −1.2.
So while ideally we would compare M3 data to models
with [Fe/H] = −1.34 and [O/Fe] = 0.2, a comparison
to scaled solar models with [Fe/H] = [M/H] = −1.2
should be an adequate surrogate. Given typical er-
rors in metallicity determinations it is appropriate to
explore a region ±0.3 dex from these values.
4.1.3. Distance modulus and age
In principle, fitting cluster main sequences to sub-
dwarfs with precise Hipparcos parallaxes should lead
to relatively high precision distance moduli. However,
M3 was not included in the first round of clusters with
distances determined in this way (Reid 1997, Gratton
et al. 1997). The main reason was that the fiducial
cluster main sequence ridge line (from Ferraro et al.
1997b; CCD97) in the V, B − V plane did not reach
deep enough with the photometric accuracy required
to apply sub-dwarf fitting to determine distance mod-
8
uli. However, using an expanded sample of local sub-
dwarfs from the newly available Hipparcos catalogue
(Carretta et al. 1999a) it is possible to give an esti-
mate for the distance modulus of M3 in the framework
of the longer distance scale defined by Gratton et al.
(1997) and Reid (1997). With E(B −V ) = 0.02±0.01
and [Fe/H] = −1.30 (to be consistent with reddening
and metallicity scales used in Gratton et al. 1997), the
true distance modulus for M3 is (m − M )0 = 15.13
(Gratton 1998, private communication, adopting the
Ferraro et al 1997b photometry). A cautionary flag
to the accuracy of sub-dwarf fitting is set by the fact
that different groups using this method do not get
the same result (Pont et al. 1997; Reid 1997; Gratton
et al. 1997). Small details are important (Carretta
et al. 1999a), and systematic effects could leave resid-
ual errors of 0.2 -- 0.3 mag in (m − M ).
There are other independent estimates of the dis-
tance to M3. Cudworth (1979) gives a kinematic
distance for M3 of (m − M )0 = 14.91. From the
RR Lyrae Sandage and Cacciari (1990) give values
of (m − M )0 ranging from 14.81 to 15.00 depending
on which metallicity/absolute magnitude relation is
adopted for the RR Lyrae. Using the latest metal-
licity/absolute magnitude determined from Baade-
Wesselink distances to field RR Lyrae (Fernley et al.
1998) and our adopted metallicity ([M/H] = −1.2),
we get (m−M )0 = 14.86 (assuming E(B−V ) = 0.01).
In a recent analysis of the RR Lyrae distance scale,
De Santis (1996) gives (m − M )0 = 15.03. Marconi
et al. (1998) find that (m − M )0 = 14.94 gives a good
fit to the LF of the lower main sequence. Ferraro
et al. (1999) have recently obtained (m−M )0 = 15.05
within the framework of a homogeneous re-analysis
of the evolved sequences of the CMD in a sample of
about 60 GGCs. They derived the distance modu-
lus from the comparison of the observed level of the
ZAHB (VZAHB ) and the theoretical models computed
by SCL97. Their use of synthetic HBs to determine
the "observed" VZAHB is a significant improvement
over earlier applications of the same technique.
Given the present state of uncertainty we consider
it reasonable to consider (m − M )V = 14.8 -- 15.2. The
question of age is intertwined with that of distance.
The longer distance scales imply smaller ages. For
example, with the "sub-dwarf distance" given above
and using stellar models from Straniero et al. 1997
(private communication), the corresponding absolute
age for M3 is about 12 Gyr. An decrease in (m − M )
of 0.07 mag leads to an increase in age of 1 Gyr. Hence
ages of 11 -- 15 Gyr for M3 are not beyond question.
4.2. The Red Giant Branch
The differential LF for our global sample is shown
in Figure 6. The dominant determinant of RGB-LF
is simply energy conservation via the so-call energy
consumption theorem (Renzini & Buzzoni 1986). Ba-
sically the number of stars in some interval along the
RGB, Ni, depends inversely on the luminosity at that
point, Ni ∝ 1/L. So log Ni ∝ Mbol, and neglect-
ing bolometric corrections log Ni should drop linearly
with magnitude moving up the RGB just as seen.
Theoretical evolutionary models predict the exis-
tence of a "feature" in the RGB called the LF-bump.
This LF-bump marks the evolutionary stage at which
the H-burning shell passes through the composition
discontinuity left by the maximum penetration of
the convective envelope (Iben 1968, Renzini & Fusi
Pecci 1988). As emphasized first by Rood & Crocker
(1985), the practical detection of such a feature re-
quires very large RGB samples. Because of this, our
global sample gives us the possibility to obtain one of
the most precise determinations of the LF-bump.
Models also predict that in the integrated LF the
slope changes on either side of LF-bump. The slope
change results from the fact that the H-burning shell
is moving through a region of increasing hydrogen
abundance below the bump and constant hydrogen
abundance above the bump. This change in slope
can more reliably locate the bump than observations
of the bump itself (Rood & Crocker 1985; Fusi Pecci
et al. 1990).
In Figure 6 the LF-bump is very well defined in
both the differential and integrated LF, and is located
at V = 15.45 ± 0.05 (in good agreement with the pre-
vious determinations (Buonanno et al. 1994, Ferraro
et al. 1997b). Also note in Figure 6 the nomenclature
we use to refer to the different branches. Our nomen-
clature is that of the theorist in which the branches
are defined in terms of the supposed interior struc-
ture. There is often confusion, in particular, over the
term sub-giant branch. To the theorist, and in this
paper, SGB refers to the transition region between the
MSTO and the lower RGB where the ridge line in the
CMD is more horizontal V = 18.2 -- 18.6. Structurally
this is the branch of thick H-shell burning where the
degenerate core develops. For historical reasons ob-
servers often refer to stars along the lower-RGB as
sub-giants.
We compare our results to two new sets of theo-
retical evolutionary tracks/isochrones/LFs which use
the new OPAL opacities and equation of state (e.g.,
Rogers & Iglesias 1992). This should offer signif-
icant improvement over earlier models particularly
for opacity dependent LF features like the LF-bump.
The first set (the no-diffusion models described in
Straniero et al. 1997, hereafter SCL97) is available
for scaled-solar abundances and may be used for any
given value of α enhancement with the Salaris et al.
(1993) prescription. The second (Bergbush & Van-
denBerg 1999, herafter V97) has [α/Fe] = 0.3. The
V97 models also include improved color-temperature
relations. On the RGB the V97 models utilize a
non-Langrangian numerical technique which is not
optimal for investigations of the RGB bump. This
technique might be expected to smear and perhaps
slightly lower the LF-bump.
In Figure 7 the global differential LF is compared
to SCL97 and V97 models for (m − M )V = 15.03,
age = 12 Gyr. For SCL97 models we took [Fe/H] =
[M/H] = −1.2. The V97 models have [Fe/H] = −1.41
and [α/Fe] = 0.3 which is equivalent to [M/H] =
−1.2. The integrated LF compared to SCL97 is shown
in Figure 8. The fits are really quite good with the
region of poorest fit at V ∼ 18 just above the SGB.
That the theoretical and observed LFs agree so well
for stars brighter than V ∼ 18 shows that the hydro-
gen profile in the region M (r) ∼ 0.2 -- 0.4 M⊙ is given
fairly accurately by canonical models.
The values of (m − M ), age, and metallicity which
lead to the best LF fit have the virtues of also agree-
ing well with the longer Hipparcos distance scale with
their correspondingly shorter ages and with the abun-
dances obtained by Carretta & Gratton (1998). Fur-
thermore, these parameters fit in well with those ob-
tained from a new re-analysis of the location of the
LF-bump in a sample of more than 40 GGCs by Fer-
raro et al. (1999). They find that the previous discrep-
ancy in the bump location between the observations
and models is completely removed using new models
and considering the newer global metallicity scale.
The good fit to the LF-bump comes at the expense
of some mismatch at the SGB/lower-RGB transition.
To explore how varying the parameters affect the fit
we must quantify the quality of the fits. Formally we
can perform a Chi-squared test computing
χ2 =
N
Xi=1
(cid:18) log(Nobserved)i − log(Nmodel)i
log(Nobserved)i
(cid:19)2
9
For the best fit to the entire data set (13 < V < 20.5),
shown in Figure 7, there is a slightly better agreement
with the SCL97 models (χ2
V97 =
0.13).
SCL97 = 0.11 and χ2
One could conceivably prefer parameters which do
not give the best overall fit. The factors which de-
termine the luminosity of the "theoretical" LF-bump
have been discussed in Fusi Pecci et al. (1990), Chi-
effi and Gratton (1986) and more recently and exten-
sively in Cassisi & Salaris (1997). The most impor-
tant factors are the opacity in the 1 -- 2 × 106 K range
and the degree that convective elements undershoot
the neutral buoyancy point of the convective enve-
lope. Elsewhere the post-TO LF depends mostly on
the rate of fuel consumption (but see §4.3 below), and
one might expect it to be a considerably more robust
prediction of the models than the location of the LF-
bump. Given this one might tolerate some residual
error in the LF-bump location to achieve a better fit
elsewhere and, thus, exclude the bump region when
measuring the quality of the fit. If we do so for the
fits shown in Figure 7 we find χ2
SCL97 xbump = 0.01
and χ2
V97 xbump = 0.02.
In Figure 9 we explore the effect of varying the dis-
tance modulus, age and chemical composition. Since
the quality of the fit is essentially the same for SCL97
and V97, we show results only for the SCL97 mod-
els. We have found it impossible to fix parameters so
that the model fits the SGB region and the LF-bump
equally well. The best fit in the SGB region results
from a higher age of 14 Gyr. However the good fit
in the SGB region brings with it a worse fit for the
LF-bump. The fit to the LF-bump can be restored
at an age of 14 Gyr by adopting a higher metallic-
ity and (m − M ) smaller than we consider plausible
(Figure 9b). The upper end of the allowable (m − M )
range gives good fit to the LF-bump (Figure 9c) at the
expense of a slightly worse fit to the SGB region and
a small age (10 Gyr). The Chi-square values both in-
cluding and excluding the LF-bump region are given
in Table 3.
From the χ-values listed in Table 3 we conclude
that considering both the two main observables of
the LF (the LF-bump and the SGB-rise) we can find
a good agreement (with both SCL97 and V97 mod-
els) when we take age = 12, (m − M )V = 15.03 and
[M/H] ∼ −1.2 (see Figure 7). On the other hand,
if we consider that some residual error still might be
present in the theoretical prediction of the LF-bump
location, we get the best agreement between theory
10
and observations if age = 14, (m − M )V = 14.80 and
[M/H] ∼ −1.2 (see Figure 9a).
A full discussion of the age should also bring in
the isochrone fits. These are complicated both by the
color calibration and the color-temperature relations.
Probably because of this our isochrone fits using the
ages and distances determined from the LF give no
further indication of whether the age of M3 is closer
to 12 or 14 Gyr.
4.3. The Lower Red Giant Branch and Sub-
Giant Branch
The lower-RGB & SGB LF can be especially in-
teresting. This is the region in which there has been
some indication of a breakdown in the canonical mod-
els for very low metallicity clusters (Stetson 1991,
Bolte 1994) and recently confirmed for M30 (Berg-
bush 1996, Guhathakurta et al. 1998, Sandquist et al.
1999).
It is the region affected by non-canonical
assumptions like WIMP energy transport (Faulkner
& Swenson 1993, Degl'Innocenti, Weiss, & Leone
(1997)), and helium settling (Proffit & VandenBerg
1991, Straniero et al. (1997)), both of which could
reduce cluster age estimates. Rapid stellar rotation
(VandenBerg, Larsen, & de Propris 1998) and [α/Fe]
could also affect the SGB region of the LF. We illus-
trate all of these in Figure 11 (see also Figs. 2 & 3 of
Vandenberg et al. 1998). The predicted changes are
all small, and age differences can mimic these changes.
As shown in Figs. 7 & 8, the the theoretical and
observed LFs agree well in the SGB region and lower-
RGB. Indeed the fits are pretty good under a wide
range of assumptions (Figure 9). We should note that
for the excess of lower-RGB stars in M30 cited above,
the observed LF lies above the lower-RGB at essen-
tial every bin. For M3 our "bad" fits correspond to
only one or two mismatched bins. The fits can be es-
sentially perfect (Figure 9a) if we accept some error
in the bump location. Basically, there is no indica-
tion for a breakdown in the canonical models. SBSH
found a similar result for M5.
On the other hand the observational consequences
of plausible non-canonical effects are small so even
our good fit may not place very tight constraints on
their magnitude. We do not attempt to place such
constraints now because we feel that a traditional LF
analysis may not be the appropriate way to analyze
the SGB. By projecting the star count information
onto the "magnitude" axis one loses the information
Chi-square results between observations and SCL models
Table 3
13 < V < 17
17 < V < 20.5
13 < V < 20.5
[M/H] = −1.2
(m − M )V = 15.03
age = 12 Gyr
[M/H] = −1.2
(m − M )V = 14.80
age = 14 Gyr
[M/H] = −0.82
(m − M )V = 14.65
age = 14 Gyr
[M/H] = −1.3
(m − M )V = 15.20
age = 10 Gyr
0.101
0.194
0.103
0.112
0.014
0.008
0.026
0.018
0.114
0.201
0.130
0.131
contained in the color distribution of the stars. In a
future paper we will report on our efforts to study
the distribution "along" the SGB (see for example,
Bergbush & VandenBerg 1997), and provide quanti-
tative estimates as to the degree that non-canonical
assumptions can be ruled out.
4.4. Below the Main Sequence Turnoff
Below the MSTO the LF primarily probes the stel-
lar initial mass function (IMF) and cluster dynamics.
Having been acquired primarily to study post-MSTO
evolution, the data we describe in this paper reaches
only ∼ 2 mag below the turnoff which corresponds to
a mass significantly above the lower mass cutoff for
main sequence stars. Deeper LFs (based on almost
25,000 stars) exploring this region are presented in
Marconi et al. (1998) and Carretta et al. (1999c).
It is widely recognized that stars in dense systems
(like GGCs) are subject to energy exchange through
stellar encounters and rapidly evolve toward a state
of equipartition of energy (so that stars of lower mass
will have higher velocities). As a direct consequence
low mass stars should have larger average distance
from the center and their distribution will be less cen-
trally concentrated than that of higher mass stars.
This mass segregation effect is expected to produce
an observable effect on the LF if a sufficiently large
radial region of the cluster is sampled.
Such mass segregation has already been found in
GGCs surveyed by HST: NGC 6397 (King et al.
1995), 47 Tuc (Paresce et al. 1995 ), M15 (De Marchi
& Paresce 1996), NGC 6752 (Ferraro et al. 1997c). Pi-
otto et al. (1997) give an extended discussion of this
topic. Mass segregation has been found in M30 using
ground based data (Bolte 1989).
Even with our limited MS coverage, we can see ev-
idence for mass segregation in M3. In Figure 12 we
show a comparison of the LFs in three separate radial
regions: r < 20′′; 50′′ < r < 100′′; 210′′ < r < 360′′.
As above the LFs have been normalized to the total
number of stars brighter than the MSTO. The inner
regions show a significant drop of star counts over
the magnitude range V ∼ 20 − 21.5. We interpret
the observed depletion as due to the mass segrega-
tion effect. The depletion is obvious even in the small
interval of masses sampled. The effect shown in Fig-
ure 12 cannot due to residual incompleteness of the
samples. In order to completely remove the effect we
11
would have to assume a level of completeness which
is far smaller than the estimated value. For exam-
ple in the innermost sample r < 20′′ (which is almost
completely contained in the PC) the measured com-
pleteness must be almost a factor 2 in error (dropping
from 0.8 down to 0.45) in order to remove the effect.
Is the observed decrease in the star counts con-
sistent with the theoretical expectation? In order to
derive the relative distribution of stars of different
mass at different distances from the cluster center we
have used the formula given by King et al. (1995) fol-
lowing the procedure as explained in DeMarchi and
Paresce (1996). The faint end of the LF should be
depleted in the core region with respect to the outer
regions by the factor f = (ρ0/ρext)m2/m1−1 where ρ0
and ρext are the stellar densities in the core and in the
external region respectively, and m2 and m1 are the
stellar masses at the faint extreme and TO level of
the LF, respectively. We use the tables of SCL97 and
assume (m − M )V = 15.0 to relate observed magni-
tudes to mass -- the MSTO at V = 19.1 corresponds
to MV ∼ 4 and a mass of ∼ 0.84 M⊙, V = 20.5 cor-
responds to MV = 5.5 and a mass of ∼ 0.73 M⊙,
and V = 21.9 corresponds MV = 6.95 and a mass of
∼ 0.61 M⊙, respectively.
The spatial densities can be approximated by the
In the case of M3
Eqn. 5 in King et al. (1995).
the external region lies at rext ∼ 290′′, and taking
rc ∼ 30′′ (Djorgovski 1993), we estimate that the ra-
tio (ρ0/ρext) is ∼ 900. Adopting the mass values ob-
tained above, the density of low-mass stars should be
a factor 9000.13 ∼ 2.4 less than that in the central.
This value is compatible with the observed depletion
which is 975/548 ∼ 1.8.
We can also explore the stellar initial mass function
(IMF) using our data. The IMF is normally approx-
imated by a power-law mass spectrum of the form
φ(M )dM = M −(1+x)dM where φ(M )dM is the num-
ber of stars in the mass range M to M + dM . The
observed slope of the IMF measured in terms of x
should vary with radius in a way consistent with a
underlying global slope. To verify this we use the dy-
namical models of Pryor, Smith, & McClure (1986).
We compute x using Eq. 3 of Bolte (1989). The two
magnitude intervals used are 18.55 < V < 20.55
and 20.55 < V < 21.95. Figure 13 shows the ob-
served x plotted as a function of radius. We see
that x(observed) varies r in a way consistent with the
x(global) ∼ 0. This is close to the value we inferred
from our total sample.
5. Summary
We have obtained a V -band luminousity function
for the globular cluster M3 using a combination of
ground based photographic and CCD observations
and HST observations of the cluster center. The sam-
ple is the largest, most complete LF every obtained
for a globular cluster. It is free from the problems of
photometric blends which plague ground-based LFs.
In the crucial MSTO/SGB region our sample is more
than three times larger than previous LFs.
Following a traditional LF function analysis we find
no surprises -- canonical stellar models seem to fit the
data well. In particular,
• The location of the RGB LF-bump agrees well
new theoretical models incorporating the OPAL
opacities and enhancements of the α-elements.
There is no indication of significant "under-
shooting" by the convective envelope. The best
overall fit is achieved with age = 12, (m −
M )V = 15.03 and global metallicity [M/H] ∼
−1.2 (equivalent to [Fe/H] = −1.34 and [O/Fe] =
0.2). On the other hand, if we consider that the
physics determining the bump location might
have more uncertainty than that determing SGB
evolution, we might tolerate a worse fit for the
bump to achieve a better fit for the SGB. In
that case we find age = 14, (m − M )V = 14.80
and [M/H] ∼ −1.2
• The slope of the integrated LF changes across
the LF bump as predicted by the models. This
implies that the H-profile (and thus fuel con-
sumed in the earlier stages) is given reasonably
accurately by standard models.
• The LF is well fit in the SGB region. There is
no indication of a problem with the canonical
models as had been earlier found for low metal-
licity clusters.
• Below the TO we find evidence for a "flat" IMF
(x ≈ 0) and evidence for mass segregation which
is well fit by King models.
The two best observed GGC LFs, that of M3 pre-
sented here and M5 by SBSH, seem to be a vindi-
cation of the quality of theoretical models.
In con-
trast, the anomalous LF of M30 first found by Bolte
(1994) has inconveniently persisted in later studies
(Bergbush 1996, Guhathakurta et al. 1998, Sandquist
12
et al. 1999). Why should the theory work so well in
some cases and fail in others? One can explore possi-
ble deficiencies in the models as in VandenBerg et al.
(1998). However, we feel that the ultimate resolution
of the problem lies in the aquisition of more high qual-
ity LFs to determine how common anomalous LFs are
and what cluster parameters they correlate with.
We thank Don VandenBerg for providing informa-
tion prior to publication and for very helpful email
discussions. RTR & BD are supported in part by
NASA Long Term Space Astrophysics Grant NAG 5-
6403 and STScI/NASA Grants GO-5969, GO-6804.
This research was partially supported by the Agenzia
Spaziale Italiana (ASI) and by the MURST as part of
the project Stellar Evolution. FRF acknowledges the
ESO Visiting Program for its hospitality.
REFERENCES
Bell, R. A., Gustafsson, B., Nordh, H. L., & Olofsson,
S. G. 1976, A&A, 46, 391
Bergbush, P. A. 1996, AJ, 112, 1061
Bergbush, P. A., & VandenBerg, D. A. 1992, ApJS,
81, 163
Bergbush, P. A., & VandenBerg, D. A. 1997, AJ, 114,
2604
Bergbush, P. A., & VandenBerg, D. A. 1999, in prepa-
ration
Bolte, M. 1989, ApJ, 341, 168
Bolte, M. 1994, ApJ, 431, 223
Carretta, E., & Gratton, R. G. 1997, A&AS, 121, 95
Cassisi, S., & Salaris, M. 1997, MNRAS, 285, 993
Castellani, V., Ciacio, F., Degl'Innocenti, S., &
Fiorentini, G. 1997, A&A, 322, 801
Chieffi, A. & Gratton, R. G. 1986, Mem. Soc. Astron.
It., 57, 395
Cudworth, K. M. 1979, AJ, 84, 1312
Degl'Innocenti, S., Weiss, A., & Leone, L. 1997, A&A,
319, 487
De Marchi G., & Paresce, F. 1996, ApJ, 467, 658
De Santis, R. 1996, A&A, 306, 755
Djorgovski, S. G. 1993 in ASP Conf. Ser. 50, Struc-
ture and Dynamics of Globular Clusters, ed. S. G.
Djorgovski & G. Meylan (Sanfrancisco: ASP), 373
Faulkner, J., & Swenson, F. J. 1993, ApJ, 411, 200
Fernley, J., Carney, B. W., Skillen, I., Cacciari, C., &
Janes, K. 1998, MNRAS, 293, L61
Ferraro, F. R., Fusi Pecci, F., Cacciari, C., Corsi,
C.E., Buonanno, R., Fahlman, G. G., & Richer,
H.B., 1993, AJ, 106, 2324
Ferraro, F. R., Paltrinieri, B., Fusi Pecci, F., Cacciari,
C., Dorman, B., Rood, R.T., Buonanno, R., Corsi,
C., Burgarella, D., & Laget, M. 1997a, A&A, 324,
915
Ferraro, F. R., Carretta, E., Corsi, C., Fusi Pecci,
F., Cacciari, C., Buonanno, R., Paltrinieri, B., &
Hamilton, D., 1997b, A&A, 320, 757
Buonanno, R., Buscema, G., Corsi, C. E., Ferraro, I.,
& Iannicola, G. 1983, A&A, 126, 278
Ferraro, F. R., Carretta, E., Bragaglia, A., Renzini,
A., & Ortolani, S. 1997c, MNRAS, 286, 1012
Buonanno, R., Corsi, C. E., Buzzoni, A., Cacciari, C.,
Ferraro, F. R., & Fusi Pecci, F. 1994, A&A, 290,
69
Ferraro, F. R., Messineo, M., Fusi Pecci, F., De Palo,
A., Straniero, O., Chieffi, A., & Limongi, M., 1999,
AJ, submitted
Carretta, E., Gratton, R. G., Clementini, G., & Fusi
Pecci, F. 1999a, astro-ph 9902086
Fusi Pecci, F., Ferraro, F. R., Crocker, D. A., Rood,
R. T., & Buonanno, R. 1990 A&A, 238, 95
Carretta, E., Gratton, R. G., Sneden, C., & Bra-
gaglia, A., 1999b, in Galaxy Evolution: Connect-
ing the Distant Universe with the Local Fossil
Record, Observatoire de Paris-Meudon, September
1998, in press (astro-ph/9812095)
Carretta, E., et al, 1999c, in preparation
Gratton, R. G., Fusi Pecci, F., Carretta, E., Clemen-
tini, G., Corsi, C., & Lattanzi, M. 1997, ApJ, 491,
749
Guhathakurta, P., Webster, Z. T., Yanney, B.,
Schneider, D. P., & Bahcall, J. N. 1998, AJ, 116,
1757
13
Holweger, H., & Muller, E. A. 1974, Solar Physics,
39, 19
Holtzmann, J. A., Burrows, C. J., Casertano, S.,
Hester, J. J., Trauger, J. T., Watson, A. M., &
Worthey, G., 1995, PASP, 107, 1065
Iben, I. 1968, Nature, 220, 143
Laget, M., Ferraro, F. R., Paltrinieri, B., & Fusi Pecci,
F. 1998, A&A, 332, 93
King, I. R., Da Costa, G. S. & Demarque, P. 1985,
ApJ, 299, 674
King, I. R., Sosin, C., & Cool, A. 1995, ApJL, 452,
13
Kraft, R. P., Sneden, C., Langer, G. E., & Prosser,
C. F. 1992, AJ, 104, 645
Rood, R. T., & Crocker, D. A. 1985, in The Pro-
duction and Distribution of C, N, O Elements, ed.
J. Danziger, F. Matteucci, & K. Ajar (Garching:
ESO), 61
Sandquist, E. L., Bolte, M., Stetson, P. B., & Hesser,
J. E. 1996, ApJ, 470, 910
Sandquist, E. L., Bolte, M., Langer, G. E., Hesser, J.
E., & Mendes de Oliveira, C. 1999, ApJ, in press,
astro-ph/9810259
Salaris, M., Chieffi, A., & Straniero, O. 1993, ApJ,
414, 580
Stetson, P. B. 1991, in ASP Conf. Ser. 13, The Forma-
tion and Evolution of Star Clusters, ed. K. Janes
(Sanfrancisco: ASP), 88
Straniero, O., Chieffi, A., & Limongi, M. 1997, ApJ,
Kraft, R. P., Sneden, C., Langer, G. E., & Shetrone,
490, 425
M. D. 1993, AJ, 106, 1190
Krist, J. 1994, Tiny Tim User's Manual Version 4.0
Kurucz, R. L. 1993, CD-ROM 13 and CD-ROM 18
Marconi, G., Buonanno, R., Carretta, E., Ferraro, F.
R., Fusi Pecci, F., Montegriffo, P., De Marchi, G.,
Paresce, F., & Laget, M. 1998, MNRAS, 293, 479
Piotto, G., Cool, A., & King, I. R. 1997, AJ, 113,
1345
Paresce, F., De Marchi, G., & Jedrzejewski, R. 1995,
ApJL, 442, 57
Pont, F., Mayor, M., Turon, C., & VandenBerg, D.
A. 1998, A&A, 329, 87
Proffit, C. R., & VandenBerg, D. A. 1991, ApJS, 74,
473
Pryor, C., Smith, G. M., & McClure, R. D. 1986, AJ,
92, 1358
Reid, I. N. 1997, AJ, 114, 161
Reid, I. N. 1998, AJ, 115, 204
Renzini, A., & Buzzoni, A. 1986, in Spectral Evolu-
tion of Galaxies, ed. C. Chiosi, A. Renzini (Dor-
drect: Reidel), 135
Renzini, A., & Fusi Pecci, F., 1988, ARA&A, 26, 199
Rogers, F. J., & Iglesias, C. A. 1992, ApJ, 401, 361
Trager, S., Djorgovski, S., & King, I. 1993, in Struc-
ture and Dynamics of Globular Clusters: ASPCS
vol. 50, eds. S Djorgowski & G. Meylan (San Fran-
cisco: ASP), 347
Vandenberg, D. A., Larson, A. M., & De Propris, R.
1998, PASP, 110, 98
This 2-column preprint was prepared with the AAS LATEX
macros v4.0.
14
6. FIGURE CAPTIONS
360′′.
Fig. 1. -- The V , V − I CMD for the HST global
sample. The sample contains more than 37,000 stars.
Fig. 13. -- Comparison between the X value (com-
puted as in Bolte 1989) and the dynamical models
from Pryor et al. 1986.
Fig. 2. -- Map of the WFPC2 field of view The cir-
cle (at r ∼ 50′′) delineates the INNER and OUTER
regions.
Fig. 3. -- The smoothed interpolating curves of com-
pleteness as a function of V magnitude.
Fig. 4. -- The zoomed LF with and without com-
pleteness corrections. The Main-Sequence-Turn-Off
point is labeled.
Fig. 5. -- CMDs of stars selected for the LF in each
of the three adopted samples: PHOTO, CCD, HST,
respectively.
Fig. 6. -- Differential Luminosity Function of the
global sample (PHOTO+CCD+HST). The names we
use to refer to various parts of the LF are indicated.
7. -- Differential
Fig.
luminosity function of the
global sample compared with theoretical models from
SCL97 (panel a) and V97 (panel b). The chosen val-
ues for distance modulus, age and chemical composi-
tion are indicated.
Fig. 8. -- Integrated luminosity function of the global
sample compared with theoretical models by SCL97.
Fig. 9. -- The differential LF of the global sample
is compared with SCL97 theoretical models varying
metallicity, distance modulus and age.
Fig. 10. -- The LF-bump region of the integrated
(panel a) and differential (panel b) LF: The position
of the LF-bump is indicated. The solid line is the ob-
served LF. The dashed line is based on SCL97 models.
The short dashed lines in (panel a) are straight line
fits to the LF above and below the LF-bump and are
drawn to aid the eye in seeing the change in slope.
11. -- The dependence of the SGB LF on
Fig.
WIMPs, He diffusion, [α/F e].
Fig. 12. -- A comparison of the LFs in three separate
radial regions: r < 20′′; 50′′ < r < 100′′; 210′′ < r <
15
V
HB
15.57
15.62
15.63
15.64
15.64
15.62
15.63
15.64
15.66
15.71
15.76
15.83
15.91
15.99
16.10
16.21
16.37
16.54
16.65
V − I
0.775
0.731
0.684
0.631
0.585
0.272
0.219
0.184
0.146
0.105
0.079
0.043
0.008
-0.021
-0.050
-0.069
-0.080
-0.088
-0.091
V
RGB
12.44
12.68
12.90
13.10
13.30
13.55
13.79
14.05
14.28
14.56
14.82
15.09
15.36
15.64
15.89
16.21
16.52
16.70
16.94
17.20
17.36
17.63
18.05
Table 4
Mean ridge line for the HST sample.
V − I
V
V − I
1.589
1.473
1.403
1.347
1.294
1.241
1.193
1.150
1.115
1.077
1.045
1.017
0.989
0.964
0.939
0.911
0.890
0.879
0.867
0.856
0.849
0.836
0.815
18.16
18.26
18.33
18.37
18.42
18.47
18.52
18.57
18.62
18.70
18.78
18.86
18.98
19.10
19.18
19.26
19.38
19.78
20.10
20.30
20.50
20.98
21.34
0.803
0.787
0.764
0.746
0.724
0.696
0.668
0.647
0.629
0.616
0.608
0.600
0.593
0.592
0.593
0.595
0.597
0.624
0.654
0.672
0.693
0.756
0.805
16
Luminosity function of the global sample (PHOTO+CCD+HST).
Table 5
V
12.70
12.90
13.10
13.30
13.50
13.70
13.90
14.10
14.30
14.50
14.70
14.90
15.10
15.30
15.50
15.70
15.90
16.10
16.30
16.50
16.70
16.90
17.10
17.30
Log(Ni)
σNi
0.70
0.30
0.78
0.85
0.95
0.85
0.90
0.90
1.11
1.08
1.32
1.15
1.34
1.38
1.76
1.60
1.51
1.65
1.70
1.79
1.85
1.95
2.04
2.06
0.45
0.26
0.45
0.41
0.52
0.50
0.50
0.47
0.61
0.55
0.67
0.58
0.77
0.72
0.92
0.87
0.75
0.83
0.93
0.96
1.00
1.06
1.09
1.11
V
17.50
17.70
17.90
18.10
18.30
18.50
18.70
18.90
19.10
19.30
19.50
19.70
19.90
20.10
20.30
20.50
20.70
20.90
21.10
21.30
21.50
21.70
21.90
Log(Ni)
σNi
2.05
2.22
2.26
2.46
2.72
2.97
3.06
3.18
3.25
3.32
3.37
3.42
3.45
3.49
3.53
3.56
3.56
3.59
3.58
3.58
3.55
3.57
3.66
1.10
1.20
1.23
1.35
1.52
1.68
1.75
1.83
1.88
1.94
1.98
2.01
2.04
2.08
2.12
2.14
2.16
2.19
2.21
2.24
2.27
2.35
2.48
17
Chi-square results between observations and SCL models
Table 6
13 < V < 17
17 < V < 20.5
13 < V < 20.5
[M/H] = −1.2
(m − M )V = 15.03
age = 12 Gyr
[M/H] = −1.2
(m − M )V = 14.80
age = 14 Gyr
[M/H] = −0.82
(m − M )V = 14.65
age = 14 Gyr
[M/H] = −1.3
(m − M )V = 15.20
age = 10 Gyr
0.101
0.194
0.103
0.112
0.014
0.008
0.026
0.018
0.114
0.201
0.130
0.131
18
This figure "f5.gif" is available in "gif"(cid:10) format from:
http://arxiv.org/ps/astro-ph/9905193v1
X=2.0
X=1.35
X=0.67
X=0.0
|
astro-ph/9906229 | 1 | 9906 | 1999-06-14T09:48:56 | The Galactic contribution to high latitude diffuse gamma-ray emission | [
"astro-ph"
] | Recent evidence for a large Galactic halo, based on cosmic-ray radioactive nuclei, implies a significant contribution from inverse Compton emission at high Galactic latitudes. We present predictions for the expected intensity distribution, and show that the EGRET gamma-ray latitude distribution is well reproduced from the plane to the poles. We show that the Galactic component at high latitudes may be comparable to the extragalactic emission in some energy ranges. | astro-ph | astro-ph |
Proc. 26th ICRC (Salt Lake City, 1999), OG2.4.03
The Galactic contribution to high latitude diffuse γ-ray emission
Andrew W. Strong1, Igor V. Moskalenko1,2, and Olaf Reimer1
1MPI fur extraterrestrische Physik, D -- 85740 Garching, Germany
2Institute for Nuclear Physics, Moscow State University, 119 899 Moscow, Russia
Abstract
Recent evidence for a large Galactic halo, based on cosmic-ray radioactive nuclei, implies a significant con-
tribution from inverse Compton emission at high Galactic latitudes. We present predictions for the expected
intensity distribution, and show that the EGRET gamma-ray latitude distribution is well reproduced from the
plane to the poles. We show that the Galactic component at high latitudes may be comparable to the extra-
galactic emission in some energy ranges.
1 Introduction:
The origin of the truly extragalactic γ-ray background is still unknown. The models discussed range from
the primordial black hole evaporation (Page & Hawking 1976) and annihilation of exotic particles in the
early Universe (Cline & Gao 1992) to the contribution of unresolved discrete sources such as active galaxies
(Sreekumar et al. 1998), while the spectrum of the extragalactic emission itself is uncertain. The latter can
be addressed only by the accurate study of the Galactic diffuse emission at high Galactic latitudes. Moreover
there is growing evidence for a large halo contribution to the γ-ray background. An indication for a γ-ray halo
was also found by Dixon et al. (1998) from analysis of EGRET data. Studies of 10Be (Strong & Moskalenko
1998) gave the range zh = 4 − 12 kpc for nucleons, Webber & Soutoul (1998) find zh = 2 − 4 kpc from 10Be
and 26Al data, Ptuskin & Soutoul (1998) find zh = 4.9+4
−2 kpc.
Gamma rays provide a tracer of the electron halo via inverse Compton (IC) emission. A study of the
Galactic emission requires a systematic approach: computation of a realistic interstellar radiation field and
self-consistent calculation of the electron spectrum in 3D. We use our GALPROP model1, which has been
shown to be consistent with many kinds of data related to cosmic ray origin and propagation, to calculate the
Galactic contribution to the high latitude diffuse γ-ray emission (Strong, Moskalenko, & Reimer 1999). The
models have been described in full detail in Strong & Moskalenko (1998), for a recent review of our results
see Strong & Moskalenko (1999).
2 Description of the models:
The models are three dimensional with cylindrical symmetry in the Galaxy. For a given halo size the
diffusion coefficient (as a function of momentum) and the reacceleration parameters are determined by the
Boron-to-Carbon ratio; the momentum-space diffusion coefficient is related to the spatial coefficient (Berezin-
skii et al. 1990). The injection spectrum of particles is assumed to be a power law in momentum, if necessary
with a break. The magnetic field is adjusted to match the 408 MHz synchrotron longitude and latitude distri-
butions. The interstellar hydrogen and He distribution uses HI and CO surveys and information on the ionized
component. Energy losses for particles by ionization, Coulomb interactions, bremsstrahlung, inverse Comp-
ton, and synchrotron are included. The distribution of cosmic-ray sources is adjusted (Strong & Moskalenko
1998) to match the cosmic-ray distribution obtained from EGRET γ-ray data (Strong & Mattox 1996). The
interstellar radiation field is based on stellar population models and COBE results, plus the cosmic microwave
background (Strong, Moskalenko, & Reimer 1999). Inverse Compton scattering is treated as in Moskalenko
& Strong (1999) including the effect of the anisotropy of the ISRF, and gas related γ-ray intensities (π0-decay
and bremsstrahlung) are computed using the column densities of HI and H2 based on 21-cm and CO surveys.
1 For
interested users our model and data sets are available in the public domain on the World Wide Web,
http://www.gamma.mpe-garching.mpg.de/∼aws/aws.html
0.5<l< 30.0 , 330.0<l<359.0
-5.0<b< 5.0
70-100MeV
1.0<l<180.0/181.0<l<359.0
Figure 1: Gamma-ray energy spectrum of the in-
ner Galaxy (300◦ ≤ l ≤ 30◦, b ≤ 5◦) compared
with our model calculations (electron injection in-
dex −1.8, and modified nucleon spectrum). Curves
show the contribution of IC, bremsstrahlung, and π0-
decay, and the total. Data: EGRET (Strong & Mat-
tox 1996), COMPTEL (Strong et al. 1999), OSSE
(l = 0, 25◦: Kinzer et al. 1999).
Figure 2: High latitude distribution (enlarged) of
70 -- 100 MeV γ-rays from the EGRET compared to
our model calculation. Separate components:
IC
(dashes), bremsstrahlung (thin histogram), π0-decay
(thick histogram), horizonal line:
isotropic back-
ground. EGRET data (point sources removed): dot-
ted line.
3 Inner Galaxy:
Recent results from both COMPTEL and EGRET indicate that IC scattering is a more important contributor
to the diffuse emission that previously believed. The puzzling excess in the EGRET data > 1 GeV relative
to that expected for π0-decay has been suggested to originate in IC scattering from a hard interstellar electron
spectrum (e.g., Pohl & Esposito 1998). Our combined approach allows us to test this hypothesis (Strong,
Moskalenko, & Reimer 1999).
Our first concern was to reproduce the γ-ray spectrum of the inner Galaxy. This is possible by invoking
a hard electron spectrum with injection spectral index -- 1.8, which after propagation (with reacceleration)
provides consistency with radio synchrotron data. Following Pohl & Esposito (1998), for this model we do
not require consistency with the locally measured electron spectrum above 10 GeV because the rapid energy
losses cause a clumpy distribution so that this is not necessarily representative of the interstellar average. For
this case, the interstellar electron spectrum deviates strongly from that locally measured.
Further improvement can be obtained by allowing some freedom in the nucleon spectrum at low energies.
Some freedom is allowed since solar modulation affects direct measurements of nucleons below 20 GeV,
and the locally measured nucleon spectrum may not necessarily be representative of the average on Galactic
scales either in spectrum or intensity due to details of Galactic structure (e.g. spiral arms). By introducing
some flattening of the nucleon spectrum below 20 GeV, a small steepening above 20 GeV, and a suitable
normalization, an improved match to the inner Galaxy EGRET spectrum is indeed possible (Fig. 1).
The modified nucleon spectrum must be checked against the stringent constraints on the interstellar spec-
trum provided by antiprotons and positrons. (Such tests sample the Galactic-scale properties of CR p and He
0.5<l<179.0 , 180.5<l<359.0
70.0<b< 89.0
0.5<l<179.0 , 180.5<l<359.0
70.0<b< 89.0
Figure 3: Energy spectrum of γ-rays from high Galactic latitudes (b ≥ 70◦, all longitudes) for zh = 4
kpc (left) and zh = 10 kpc (right). Shaded areas: EGRET total intensity from Cycle 1 -- 4 data. COMPTEL
data: high-latitude total intensity (open boxes: Bloemen et al. 1999, diamonds: Kappadath 1998, crosses:
Weidenspointner et al. 1999).
rather than just the local region, independent of fluctuations due to local primary CR sources.) As expected
the ¯p and e+ predictions are higher than for the conventional model (with nucleon spectrum matching the local
measurements) but still within the observational limits (Strong, Moskalenko, & Reimer 1999).
This is our best model so far. Further tests against the γ-ray longitude and latitude profiles at all the EGRET
energies show a good overall agreement (an example is shown in Fig. 2).
In order to reproduce the low-energy (< 30 MeV) γ-ray emission via diffuse processes it is necessary
to invoke an upturn of the electron spectrum below about 200 MeV to compensate the increasing ionization
losses. (A steep slope continuing to higher energies would violate the synchrotron constraints on the spectral
index.) However, the adoption of such a steep low-energy electron spectrum has problems associated with
the very large power input to the interstellar medium (Skibo et al. 1997), and is ad hoc with no independent
supporting evidence. Moreover the OSSE-GINGA γ-ray spectrum is steeper than E−2 below 500 keV (Kinzer
et al. 1999) which would require an even steeper electron injection spectrum than adopted here. It is more
natural to consider that the COMPTEL excess is just a continuation of the same component producing the
OSSE-GINGA spectrum. Most probably therefore the excess emission at low energies is produced by a
population of sources such as supernova remnants, as has been proposed for the diffuse hard X-ray emission
from the plane observed by RXTE (Valinia et al. 1998), or X-ray transients in their low state as suggested for
the OSSE diffuse hard X-rays (Lebrun et al. 1999).
4 High latitude γ-rays and the size of the electron halo:
We use our model for calculation of the Galactic contribution to the high latitude diffuse γ-ray emission.
The high latitude γ-ray intensity increases with halo size due to IC emission (though much less than linearly
due to electron energy losses), which allows us to put an upper limit on the halo size. Fig. 3 shows the γ-ray
spectrum towards the Galactic poles for zh = 4 kpc, and 10 kpc. zh = 10 kpc is possible although the latitude
profiles around 100 MeV are then very broad and at the limit of consistency with EGRET data. Further the
isotropic component would have to approach zero above 300 MeV, so that this halo size can be considered an
upper limit.
If the halo size is 4 -- 10 kpc as we argue, the contribution of Galactic emission to the total at high latitudes
is larger than previously considered likely and has consequences for the derivation of the diffuse extragalac-
tic emission (e.g., Sreekumar et al. 1998). An evaluation of the impact of our models on estimates of the
extragalactic spectrum is beyond the scope of the present work.
5 Conclusions:
The large electron/IC halo suggested here reproduces well the latitude variation of γ-ray emission from
the plane to the poles, which can be taken as support for the halo size deduced from independent studies of
cosmic-ray composition. Halo sizes in the range zh = 4 − 10 kpc are favoured by both analyses.
References
Berezinskii, V.S., et al. 1990, Astrophysics of Cosmic Rays (Amsterdam: North Holland)
Bloemen, H., et al. 1999, Astroph. Lett. Comm. (3rd INTEGRAL Workshop), in press
Cline, D.B., & Gao, Y.-T. 1992, A&A 256, 351
Dixon, D.D., et al. 1998, New Astronomy 3, 539
Kappadath, S.C. 1998, PhD Thesis, University of New Hampshire, USA
Kinzer, R.L., Purcell, W.R., & Kurfess, J.D. 1999, ApJ 515, 215
Lebrun, F., et al. 1999, Astroph. Lett. Comm. (3rd INTEGRAL Workshop), in press
Moskalenko, I.V., & Strong, A.W. 1999, submitted (astro -- ph/9811296)
Page, D.N., & Hawking, S.W. 1976, ApJ 206, 1
Pohl, M., & Esposito, J.A. 1998, ApJ 507, 327
Ptuskin, V.S., & Soutoul, A. 1998, A&A 337, 859
Skibo, J.G., et al. 1997, ApJ 483, L95
Sreekumar, P., et al. 1998, ApJ 494, 523
Strong, A.W., & Mattox, J.R. 1996, A&A 308, L21
Strong, A.W., & Moskalenko, I.V. 1998, ApJ 509, 212
Strong, A.W., & Moskalenko, I.V. 1999, in ASP Conf. Ser. 171, 154
Strong, A.W., Moskalenko, I.V., & Reimer, O. 1999, submitted (astro -- ph/9811284)
Strong, A.W., et al. 1998, Astroph. Lett. Comm. (3rd INTEGRAL Workshop), in press
Valinia, A., & Marshall, F.E. 1998, ApJ 505, 134
Webber, W.R., & Soutoul, A. 1998, ApJ 506, 335
Weidenspointner, G., et al. 1999, Astroph. Lett. Comm. (3rd INTEGRAL Workshop), in press
|
astro-ph/0312452 | 1 | 0312 | 2003-12-17T15:54:02 | Strangeness in Neutron Stars | [
"astro-ph",
"nucl-th"
] | We discuss the role of strangeness on the internal constitution and structural properties of neutron stars. In particular, we report on recent calculations of hyperon star properties derived from microscopic equations of state for hyperonic matter. Next, we discuss the possibility of having a strange quark matter core in a neutron star, or the possible existence of strange quark matter stars, the so-called strange stars. | astro-ph | astro-ph |
Strangeness in Neutron Stars
Ignazio Bombacia
aDipartimento di Fisica "Enrico Fermi", Universit`a degli Studi di Pisa,
and INFN sezione di Pisa, via Buonarroti, 2, I-56127, Pisa, Italy
We discuss the role of strangeness on the internal constitution and structural properties
of neutron stars. In particular, we report on recent calculations of hyperon star properties
derived from microscopic equations of state for hyperonic matter. Next, we discuss the
possibility of having a strange quark matter core in a neutron star, or the possible existence
of strange quark matter stars, the so-called strange stars.
1. INTRODUCTION
The true nature and the internal constitutions of the ultra-dense compact stars known
as neutron stars is one of the most fascinating enigma in modern astrophysics [ 1, 2, 3].
Different models for the equation of state (EOS) of dense matter predict the neutron star
maximum mass Mmax to be in the range 1.4 -- 2.2 M⊙ (M⊙ ≃ 2 × 1033g is the mass of
the Sun), and a corresponding central density in the range of 4 -- 8 times the normal
saturation density (ρ0 ∼ 2.8 × 1014g/cm3) of nuclear matter. Thus neutron stars are the
most likely sites in the universe in which strangeness-bearing matter with a strangeness to
baryon ratio fS = −S/B ∼ 1 may exist. Strangeness in neutron stars is expected both in
a confined form (hyperons and or kaons), or in a deconfined form (strange quark matter).
Accordingly different types of "neutron stars" are expected theoretically, as schematically
summariezed in Fig. 1.
2. NEUTRON STARS OR HYPERON STARS?
In a conservative and oversimplified picture the core of a neutron star is modeled as
a uniform fluid of neutron rich nuclear matter in equilibrium with respect to the weak
interactions (β -- stable nuclear matter). The presence of hyperons in neutron stars was
first proposed in 1960 by Ambartsumyan and Saakyan [ 4], and since then it has been
investigated by many authors. The reason why hyperons are expected in the high dense
core of a neutron star is very simple, and it is mainly a consequence of the fermionic
nature of nucleons, which makes the nucleon chemical potentials a very rapidly increasing
function of density. As soon as the chemical potential of neutrons becomes sufficiently
large (see Fig. 2), the most energetic neutrons (i.e. those on the Fermi surface) can decay
via the weak interactions into Λ hyperons and form a Fermi sea of this new hadronic
species with µΛ = µn. The Σ− can be produced via the weak process e− + n → Σ− + νe
2
quark−hybrid
star
hyperon
star
o
c
r
e
p
su
N+e
N+e+n
n,p,e, µ
n d c t
u
ni
n,p,e,µ
g
−π
−
K
u,d,s
quarks
H
Σ,Λ,Ξ,∆
absolutely stable
strange quark
matter
u d s
µ
m
s
traditional neutron star
n s
u
p
e
rfluid
p
r
o
t
o
n
s
crust
neutron star with
pion condensate
Fe
6
3
g/cm
11
14
3
3
g/cm
g/cm
10
10
10
strange star
nucleon star
R ~ 10 km
M ~ 1.4 M
Figure 1. Schematic cross section of a neutron star according to different possibilities for
the stellar constituents (adapted from ref. [ 2]).
when the Σ− chemical potential fulfill the condition1 µΣ− = µn + µe. As we can see
from the results depicted in Fig. 2, hyperons appear at a relatively moderate density of
about 2 times the normal saturation density (n0 = 0.16 fm−3) of nuclear matter. Notice
that the Σ− hyperon appears at a lower density than the Λ, even though the Σ− is more
massive than the Λ. This is due to the contribution of the electron chemical potential
µe to the threshold condition for the Σ− (i.e. MΣ− = µn + µe, for free hyperons) and
to the fact that µe in dense matter is large and can compensate for the mass difference
MΣ− − MΛ = 81.76 MeV.
In Fig. 3, we show the profile of a such an hyperon star [ 6]. As we see the hyperonic
matter inner core of the star extend for about 8 km. This radius has to be compared with
the total stellar radius R ∼ 11 km, and with the thickness of the nuclear matter layer
(outer core) which is about 2 km. Thus neutron stars are "giant hypernuclei" [ 7] under
the influence of gravity and strong interactions.
The influence of hyperons on neutron stars properties has been investigated using dif-
ferent approaches to determine the EOS of hyperonic matter. One of the most popular
approaches, to solve this problem, is the relativistic mean field model [ 7, 8]. Some of the
1except from the very initial stage soon after neutron star birth, neutrinos freely escape the star and thus
the neutrino chemical potentials have not to be considered in the chemical equilibrium equations.
3
1300
Interacting hyperons
1250
MΣ−
MΛ
µ
n+ µ
e
µ
n
Free hyperons
µΣ−
µΛ
µ
n+ µ
e
µ
n
1200
1150
1100
1050
1000
950
0.4
0.2
1.2
Baryon number density [fm-3]
0.6
0.8
1
0
0.2
0.4
0.6
0.8
Baryon number density [fm-3]
1
1.2
1300
1250
1200
1150
1100
1050
1000
]
V
e
M
[
s
l
a
i
t
n
e
t
o
p
l
a
c
i
m
e
h
C
950
0
Figure 2. Chemical potentials in β-stable hyperonic matter. Left panel: free hyperons.
Right panel: interacting hyperons (NSC97e interaction). Adaped from ref. [ 5].
parametrizations of the lagrangian of the theory have tried to reconcile measurd values of
neutron star masses with the binding energy of the Λ particle in hypernuclei [ 9, 10]. A dif-
ferent approach is based on the use of local effective potentials to describe the in-medium
baryon-baryon (BB) interaction [ 11]. This method mimic and generalize to the case of
hyperonic matter the one based on the Skyrme nuclear interaction in the case of nuclear
matter. Here we will report on some recent results based on a third approach, which starts
from the basic BB interaction and solve the many-body problen to get the EOS for hyper-
onic matter. This method is based on an extension of the Brueckener-Bethe-Goldstone
(BBG) theory to include hyperonic degrees of freedom [ 12, 13, 14, 15, 6]. In particular,
the study of ref. [ 6] focus on the properties of a newborn neutron star, and explore the
consequences of neutrino trapping in dense matter on the structural properties and on
the early evolution of neutron stars [ 16].
In Fig. 4, we show the EOS for β-stable dense matter (i.e. in equilibrium with respect
[ 15] using the Nijmegen soft-core
to the weak interactions) obtained by Vidana et al.
[ 17] to describe the hyperon-nucleon (YN) and hyperon-
potential (NSC97e) of ref.
hyperon (YY) interaction within the Brueckner-Hartree-Fock (BHF) approximation of
the extended BBG theory. In ref. [ 15], the pure nucleonic contribution to the EOS has
been included using a parametrization of the Akmal-Pandharipande EOS [ 18], where a
semi-phenomenological three-nucleon (NNN) interaction of the Urbana type is added to
the nuclear hamiltonian to reproduce the empirical saturation point of nuclear matter.
As expected, the presence of hyperons makes the EOS much softer with respect to
the pure nucleonic case. The softening of the EOS caused by the presence of hyperons
has important consequences on many macroscopic properties of the star: the maximum
stellar mass is reduced by ∆Mmax ∼ 0.5 -- 0.8 M⊙, and the corresponding central density
is increased. Also, hyperon stars are more compact (i.e. they have a smaller radius) with
4
100
10-1
s
n
o
i
t
c
a
r
f
e
l
c
i
t
r
a
P
10-2
10-3
RY
RH R
n
p
Σ−
e-
µ−
Λ
10-4
0
1
2
4
3
10
Radial coordinate r [km]
8
9
5
6
7
11 12
13
Figure 3. The internal composition of a neutron star with hyperonic matter core. The
stellar baryonic mass is MB = 1.34 M⊙. RY is the radius of the hyperonic core. The
nuclear matter layer extend between RY and RH and has a thickness of about 2 km. The
stellar crust extend between RH and R and is about 1 km thick. (Adapted from Vidana
et al. [ 6]).
respect to pure nucleonic neutron stars. This is illustrated in Fig. 5, where we show the
mass radius-relation for traditional neutron stars and for hyperon stars obtained with the
microscopic EOS of ref. [ 15] (left panel) and with the relativistic mean field EOS (GM3
model) given in ref. [ 9]. The results depicted in Fig. 5 clearly demonstrate that to neglect
hyperons leads to an overstimate of the maximum mass of neutron stars.
It is important to notice the "low" value of the stellar maximum mass, predicted within
the approach of ref. [ 15], which is in contrast with some precise determination of neutron
star masses [ 19]. For example in the case of the neutron star associated to the pulsar
PSR1913+16, the measured stellar mass is [ 20]
MP SR1913+16 = (1.4411 ± 0.0007) M⊙ .
(1)
The prediction of a value for Mmax below the measured neutron star masses is a common
feature of all the present microscopic EOS of hyperonic matter based on G-matrix BHF
calculations [ 13, 15, 6]. For example, the authors of ref.
[ 13], in case of the Argonne
v18 NN interaction, found Mmax = 2.00 M⊙, a corresponding radius of R = 10.54 km
and a central density ρc = 1.11 fm−3 for neutron stars with a pure nucleonic core. When
hyperons are considered as possible stellar constituents, they found [ 13] Mmax = 1.22 M⊙,
a corresponding radius of R = 10.46 km and a central density ρc = 1.25 fm−3. Therefore
the current equations of state for hyperonic matter, deduced from microscopic G-matrix
BHF calculations, are "too soft" to explain observed neutron star masses.
Clearly, one should try to trace the origin of this problem back to the underlying YN
and YY two body interactions or to the possible repulsive three-body baryonic forces
2.5
2
1.5
1
0.5
pn-matter
NN and free hyperons
NN and YN
NN, YN and YY
]
3
m
c
/
g
5
1
0
1
x
[
ρ
y
t
i
s
n
e
d
s
s
a
m
0.2 0.4 0.6 0.8
0
0
1.2
Baryon number density [fm-3]
1
0
0.5
1
1.5
2
ρ [ x 1015 g/cm3]
5
10
8
6
4
2
]
2
m
c
/
e
n
y
d
5
3
0
1
x
[
P
e
r
u
s
s
e
r
P
0
2.5
Figure 4. Equation of state of dense hadronic matter with and without hyperons [ 15].
involving one or more hyperons, not included in the work of ref. [ 12, 13, 14, 15, 6].
Presently this is a subject of very active research by people working in this field. Therefore,
the use of microscopic EOSs of hyperonic matter in the contest of neutron star physics
is of fundamental importance for our understanding of the strong interactions involving
hyperons, and to learn how these interactions behave in dense many-body systems.
3. KAON CONDENSATION IN NEUTRON STARS
The inner core of neutron stars could also contain a Bose-Einstein condensate of negative
kaons [ 21, 22, 23, 24, 25]. As the density of stellar matter is increased, the K − energy
is lowered by the attractive vector mean field originating from dense nucleonic matter.
When the K − energy becomes smaller than the electron chemical potential µe (which is
an increasing function of density) the strangeness changing process e− → K − +ν becomes
possible. The critical density for this process has been calculated to be in the range 2.5
-- 5.0 n0 [ 23, 24].
Due to the lack of space, we do no have the possibilty to discuss the many relevant
implications that kaon condensation has for the structure and the evolution of neutron
stars. We refer the reader to the original literature on the subject (see e.g. [ 21, 22, 23,
24, 25] and references therein quoted).
4. HYBRID STARS
The core of the more massive neutron stars is one of the best candidates in the Universe
where a phase transition from hadronic matter to a deconfined quark phase should occur.
The quark-deconfinement phase transition proceeds through a mixed phase over a finite
range of pressures and densities [ 26, 1]. At the onset of the mixed phase, quark matter
6
2.5
2
1.5
1
]
s
t
i
n
u
s
s
a
m
r
a
l
o
s
[
G
M
0.5
pn-matter
NN and YN
NN, YN and YY
With hyperons
PSR 1913+16
2.5
pn-matter
2
1.5
1
0.5
0
8
9
11 12 13 14
10
Radius [km]
8
9
0
11 12 13 14
10
Radius [km]
Figure 5. Mass-Radius relation for "traditional" neutron stars and hyperon stars calcu-
lated [ 15] within the BHF approach with the NSC9e interaction (left panel) and with
the relativistic mean field EOS GM3 of ref. [ 9] (rigth panel). The dotted horizontal line
indicates the measured mass for the "neutron star" in the radio pulsar PSR1913+16 .
droplets form a Coulomb lattice embedded in a sea of hadrons and in a roughly uniform
sea of electrons and muons. As the pressure increases various geometrical shapes (rods,
plates) of the less abundant phase immersed in the dominant one are expected. Finally the
system turns into uniform quark matter at the highest pressure of the mixed phase [ 27].
Compact stars which possess a quark matter core, either as a mixed phase of deconfined
quarks and hadrons, or as a pure quark matter phase, are called Hybrid Stars [ 1, 2, 3].
Many possible astrophysical signals for the appearence of a quark core in neutron stars
have been proposed in the last few years (see [ 1, 2, 3] and references therein quoted).
Particularly, pulse timing properties of pulsars have attracted much attention since they
are a manifestation of the rotational properties of the associated neutron star. The onset
of quark-deconfinement in the core of the star, will cause a change in the stellar moment
of inertia [ 28]. This change will produce a peculiar evolution of the stellar rotational
period (P = 2π/Ω) which will cause large deviations of the so called pulsar braking
index n(Ω) = (Ω Ω/ Ω2) from the canonical value n = 3, derived within the magnetic
dipole model for pulsars and assuming a constant moment of inertia for the star. The
possible measurement of a value of the braking index very different from the canonical
value (i.e.
n >> 3) has been proposed [ 28] as a signature for the occurrence of the
quark-deconfinement phase transition in a neutron star. However, it must be stressed
that a large value of the braking index could also results from the pulsar magnetic field
decay and/or alignment of the magnetic axis with the rotation axis [ 29].
7
5. STRANGE STARS
The possible existence of a new class of compact stars completely made of deconfined
u,d,s quark matter (strange quark matter (SQM)) is one of the consequences of an hypoth-
esis [ 30] formulated by A.R. Bodmer in 1971 and revived by E. Witten in 1984. These
stars are usually called Strange Stars. According to the Bodmer-Witten hypothesis SQM
could be the true ground state of matter. In other words, at zero temperature and pres-
sure, the energy per baryon of SQM could be less than the energy per baryon of 56Fe,
which is the most tightly bound nucleus in nature. The strange matter hypothesis does
not conflict with the existence of atomic nuclei as conglomerates of nucleons, or with the
stability of "ordinary" matter [ 31, 32, 33]. Thus strange stars may exist in the universe.
One of the most likely strange star candidate is the compact object in the transient
X-ray burst source SAX J1808.4-3658 (ref. [ 34]). This X-ray source was discovered in
1996 by the BeppoSAX satellite. Two bright type-I X-ray bursts were detected, each
lasting less than 30 seconds. Analysis of the bursts in SAX J1808.4-3658 indicates that it
has a peak X-ray luminosity of 6 × 1036 erg/s in its bright state, and a X-ray luminosity
lower than 1035 erg/s in quiescence. SAX J1808.4-3658 is a X-ray millisecond pulsar with
a pulsation period of 2.49 ms, also it is a member of a binary stellar system with orbital
period of two hours. Using the observational data collected by the Rossi X-ray Timing
Explorer during the the 1998 April-May outburst, Li et al. [ 34] have obtained an upper
limit for the compact star radiud as a function of the unknown stellar mass. Comparing
this observational mass-radius (M-R) relation of SAX J1808.4-3658 with the theoretical
M-R realtions for traditional neutron stars, hyperon stars, stars with kaon condensation,
and strange stars Li et al. [ 34] (see their Fig. 1) argue that a strange star model is more
consistent with SAX J1808.4-3658, and suggest that it could be a strange star.
SAX J1808.4-3658 is not the only LMXBs which could harbour a strange star. Recent
studies have shown that the compact stars associated with the X-ray burster 4U 1820-30
(ref.[ 35]), the bursting X-ray pulsar GRO J1744-28 (ref.[ 36]), the X-ray pulsar Her X-1
(ref.[ 37]), the kHz QPOs source 4U 1728-34 (ref.[ 38]), are likely strange star candidates.
Recently, it has been suggested that the isolated compact star RX J1856.5-3754 (ref. [
39]) could be a strange star.
6. QUARK-DECONFINEMENT PHASE TRANSITION IN NEUTRON
STARS AND GAMMA-RAY BURSTS
Gamma Ray Bursts (GRBs) are one of the most violent and mysterious phenomena
in the universe (see e.g.
ref.[ 40] for a general introduction on this subject). During
the last ten years two satellites, the Compton Gamma Ray Observatory (CGRO) and
BeppoSAX, have revolutionized our understanding of GRBs. The Burst And Transient
Source Experiment (BATSE) on board of the CGRO has demonstrated that GRBs origi-
nate at cosmological distances. The BeppoSAX discovered the X-ray afterglow. This has
permitted to determine the position of some GRBs, to identify the host galaxy and, in
a number of cases, to measure the red-shift. If the energy is emitted isotropically, the
measured fluence of the bursts implies an energy of the order of 1053 erg. However, there
is now compelling evidence that the γ-ray emission is not isotropic, but displays a jet-like
geometry. In this case the GRB energy is of the order of 1051 erg [ 41].
8
Many cosmological models for the energy source of GRBs have been proposed. Presently
one of the most popular is the so-called "collapsar", or "hypernova" model. Alternative
models are the merging of two neutron stars (or a neutron star and a black hole) in a
binary system, or the accretion of matter into a black hole. The present report is not the
appropriate place to discuss the various merits and drawbacks of the many theoretical
models for GRBs. In the following, we will mention some recent research which try to
make a connection between GRBs and quark-deconfinement phase transition.
A possible central engine for GRBs is the conversion of a pure hadronic compact star
to a strange star. The stellar conversion is triggered by the formation of a SQM drop in
the center of the hadronic star. This idea was proposed long time ago by Alcook et al. [
42]. Recently detailed calculations, based on different realistic models for the equation of
state of neutron star matter and SQM, have been performed by the authors of ref.[ 43].
They showed that the total amount of energy liberated in the conversion is in the range
(1 -- 4)×1053erg. This energy will be mainly taken away by the neutrinos produced during
the quark-deconfinement phase transition. If the efficiency of the conversion of neutrinos
to γ is of the order of a few percent [ 44], then the birth of a strange star from a neutron
star could be the energy source for GRBs.
A mounting number of observational data suggest a clear connection between supernova
(SN) explosions and GRBs [ 45, 46, 47, 48, 49]. Particularly, in the case of the gamma ray
burst of July 5, 1999 (GRB990705) and in the case of GRB011211, it has been possible to
estimate the time delay between the two events. For GRB990705 the supernova explosion
is evaluated to have occurred a few years before the GRB [ 45, 50], while for GRB011211
about four days before the burst [ 46].
The scenario which emerges from these findings is the following two-stage scenario: (i)
the first event is the supernova explosion which forms a compact stellar remnant, i.e. a
neutron star (NS); (ii) the second catastrophic event is associated with the NS and it is
the energy source for the observed GRB. These new observational data, and the scenario
outlined above, poses severe problems for most of the current theoretical models for the
central energy source of GRBs. The main difficulty of all these models is to give an answer
to the following questions: what is the origin of the second "explosion"? How to explain
the long time delay between the two events?
In the so-called supranova model [ 51] for GRBs the second catastrophic event is the
collapse to a black hole of a supramassive neutron star, i.e. a fast rotating NS with a
baryonic mass MB above the maximum baryonic mass MB,max for non-rotating configu-
rations. In this model, the time delay between the SN explosion and the GRB is equal
to the time needed by the fast rotating newly formed neutron star to get rid of angular
momentum and to reach the limit for instability against quasi-radial modes where the
collapse to a black hole occurs [ 52]. The supranova model needs a fine tuning in the
initial spin period Pin and baryonic stellar mass MB,in to produce a supramassive neutron
star that can be stabilized by rotation up to a few years. For example, if Pin ≥ 1.5 ms,
then the newborn supramassive neutron star must be formed within ∼ 0.03M⊙ above
MB,max [ 52].
In a very recent paper, Berezhiani et al. [ 53] (see also ref. [ 54]) have proposed a new
model to explain the SN -- GRB association and in particular the long time delay inferred
for GRB990705 and GRB011211. In the model of ref.[ 53], the second explosion is related
9
to the conversion from a metastable purely Hadronic Star (neutron star or hyperon star)
into a more compact star in which deconfined quark matter is present (i.e. a hybrid
star or a strange star). The new and crucial idea in the work of ref.
[ 53] with respect
to previous work [ 43], is the metastability of the hadronic star due to the existence of
a non-vanishing surface tension at the interface separating hadronic matter from quark
matter. The mean-life time of the metastable hadronic star can then be connected to the
delay between the SN explosion and the GRB. The nucleation time (i.e. the time to form
a critical-size drop of quark matter) can be extremely long if the mass of the star is small.
Via mass accretion the nucleation time can be dramatically reduced and the star is finally
converted from the metastable into the stable configuration [ 53, 54, 55]. A huge amount
of energy, of the order of 1052 -- 1053 erg, is released during the conversion process and can
produce a powerful gamma ray burst. Within the model proposed by Berezhiani et al. [
53] is is possible to have different time delays between the two events since the mean-life
time of the metastable hadronic star depends on the value of the stellar central pressure.
Thus the model of ref. [ 53] is able to interpret a time delay of a few years (as observed
in GRB990705 [ 45, 50]), of a few days (as in the case of GRB011211 [ 46]), or the nearly
simultaneity of the two events (as in the case of SN2003dh and GRB030329 [ 56]).
Acknowledgments
I am greatly indebted to I. Vidana for providing me some unpublished results from his
Ph.D. thesis and for the great help in preparing most of the figures of the present work.
REFERENCES
1. N.K. Glendenning, Compact Stars: Nuclear Physics, Particle Physics, and General
Relativity, Springer Verlag, 1996.
2. F. Weber, Pulsars as Astrophysical Laboratories for Nuclear and Particle Physics, IoP
Publishing, 1999.
3. D. Blaschke, N. K. Glendenning and A. Sedrakian (eds.), Physics of Neutron Star
Interiors, Lecture Notes in Physics, vol. 57, Springer Verlag, 2001.
I. Vidana, Ph.D. Thesis, University of Barcelona, 2001.
I. Vidana, I. Bombaci, A. Polls, and A. Ramos, Astron. & Astrophys. 399 (2003) 687.
4. V.A. Ambartsumyan and G.S. Saakyan, Sov. Astron. AJ. 4 (1960) 187.
5.
6.
7. N.K. Glendenning, Astrophys. Jour. 293 (1985) 470.
8. J. Ellis, J.I. Kapusta and K.A Olive, Nucl. Phys. B348 (1991) 345.
9. N. K. Glendenning and S.A. Moszkowski, Phys. Rev. Lett. 67 (1991) 2414.
10. H. Huber, M.K. Weigel and F. Weber, Z. Naturforsch. 54A (1999) 77.
11. S. Balberg and A. Gal, Nucl. Phys. A625, (1997) 435.
12. M. Baldo, G.F. Burgio, and H.J. Schulze, Phys. Rev. C 58 (1998) 3688.
13. M. Baldo, G.F. Burgio, and H.J. Schulze, Phys. Rev. C 61 (2000) 055801.
14. I. Vidana, A. Polls, A. Ramos, M. Hjorth-Jensen, and V.G.J. Stoks, Phys. Rev. C 61,
(2000) 025802.
15. I. Vidana, A. Polls, A. Ramos, L. Engvik, and M. Hjorth-Jensen, Phys. Rev. C 62
(2000) 035801.
10
16. M. Prakash, I. Bombaci, M. Prakash, P.J. Ellis, R. Knorren and J.M. Lattimer, Phys.
Rep. 280 (1997) 1.
17. V. G. J. Stock and Th. A. Rijken, Phys. Rev. C 59 (1999) 3009.
18. A. Akmal, V. R. Pandharipande, Phys. Rev. C56 (1997) 2261.
19. M. H. van Kerkwijk, J. van Paradijs, and E.J. Zuiderwijk, Astron. & Astrophys. 303
(1995) 497.
20. J. H. Taylor, and J.M. Weisberg, Astrophys. Jour. 345, 434 (1989).
21. D.B. Kaplan and A. E. Nelson, Phys. Lett. B 175 (1986) 57; 179 (1986) 409(E).
22. G.E. Brown, C.-H. Lee, M. Rho and V. Thorsson, Nucl. Phys. 567 (1994) 937.
23. V. Thorsson, M. Prakash and J.M. Lattimer, Nucl. Phys. A572 (1994) 693.
24. C.-H. Lee, Phys. Rep. 275 (1996) 255.
25. N.K. Gelndenning and J. Schaffner-Bielich, Phys. Rev. Lett. 81 (1998) 4564.
26. N.K. Glendenning, Phys. Rev. D 46 (1992) 1274
27. H. Heiselberg, C.J. Pethick and E.F. Staubo, Phys. Rev. Lett., 70 (1993) 1355
28. N.K. Glendenning, S. Pei and F. Weber, Phys. Rev. Lett. 79 (1997) 1603.
29. T.M. Tauris and S. Konar, Astron. & Astrophys. 376 (2001) 543.
30. A. R. Bodmer, Phys. Rev. D 4 (1971) 1601; E. Witten, Phys. Rev. D 30 (1984) 272.
31. E. Farhi and R. L. Jaffe, Phys. Rev. D 30 (1984) 2379.
32. J. Madsen, Lecture Notes in Physics, Springer Verlag, 516 (1999).
33. I. Bombaci, in ref. [ 3], pag. 253.
34. X.-D. Li, I. Bombaci, M. Dey, J. Dey, and E.P.J. van den Heuvel, Phys. Rev. Lett. 83
(1999) 3776.
35. I. Bombaci, Phys. Rev. C 55 (1997) 1587.
36. K.S. Cheng, Z.G. Dai, D.M. Wai, and T. Lu, Science 280 (1998) 407.
37. M. Dey, I. Bombaci, J. Dey, S. Ray, and B.C. Samanta, Phys. Lett. B 438, 123 (1998);
erratum, Phys. Lett. B 467, 303 (1999).
38. X.-D. Li, S. Ray, J. Dey, M. Dey, and I. Bombaci, Astrophys. Jour. 527 (1999) L51.
39. J. Drake et al., Astrophys. Jour. 572 (2002) 996.
40. T. Piran, Phys. Rep. 314 (1999) 575; Phys. Rep. 333 (2000) 529.
41. D. A. Frayl et al., Astrophys. Jour. 562 (2001) L55 (2001.
42. C. Alcock, E. Farhi, and A. Olinto, Astrophys. Jour. 310 (1986) 261.
43. I. Bombaci, and B. Datta, Astrophys. Jour. 530 (2000) L72.
44. J.D. Salmonson, and J.R. Wilson, Astrophys. Jour. 517 (1999) 859.
45. L. Amati et al., Science 290 (2000) 953.
46. J.N. Reeves et al., Nature 414 (2002) 512.
47. J.S. Bloom et al., Nature 401 (1999) 453.
48. L. Piro et al., Science 290 (2000) 955.
49. L.A. Antonelli et al., Astrophys. Jour. 545 (2000) L39.
50. Lazzati et al., Astrophys. Jour. 556 (2001) 471.
51. M. Vietri and L. Stella, Astrophys. Jour. 507 (1998) L45.
52. B. Datta, A.V. Thampan and I. Bombaci, Astron.& Astrophys. 334 (1998) 943.
53. Z. Berezhiani, I. Bombaci, A. Drago, F. Frontera, and A. Lavagno, Astrophys. Jour.
586 (2003) 1250.
54. I. Bombaci, I. Parenti and I. Vidana, Astrophys. Jour. (2003) submited.
55. I. Vidana, I. Bombaci, I. Parenti, these proceedings.
56. J. Hjorth et al., Nature 423 (2003) 847.
11
|
astro-ph/0505416 | 1 | 0505 | 2005-05-19T19:15:30 | FUSE Measurements of Far Ultraviolet Extinction. II. Magellanic Cloud Sight Lines | [
"astro-ph"
] | We present an extinction analysis of 9 paths through the LMC and SMC based on FUSE observations. To date, just two LMC sight lines have probed dust grain composition and size distributions in the Clouds using spectra including wavelengths as short as 950 A. We supplement these with results from 4 regions distinguished by their IR through UV extinction curves and grouped as LMCAvg, LMC2, SMC bar and SMC wing. Despite the distinct characters of Milky Way and Magellanic Cloud extinction, our results are generally analogous to those found for Galactic curves in that the FUSE portions of each extinction curve are described reasonably well by FM curves fitted only to longer wavelength data and lack any dramatic new extinction features, and any deviations from the CCM formalism continue into FUV wavelengths. An MEM analysis of these curves suggests that LMCAvg and SMC wing sight lines require more silicon and/or carbon in dust than current abundance measurements would allow, while the requirements for LMC2 and SMC bar sight lines do not fully tax the available reservoirs. An intermediate product of this analysis is the measurement of new H_2 abundances in the Magellanic Clouds. Collectively considering Cloud sight lines that possess significant H_2 column densities, E(B-V)/N(HI) ratios are reduced by significant factors relative to the Galactic mean, whereas the corresponding E(B-V)/N(H_2) values more closely resemble their Galactic counterpart. These trends reflect the fact that among these sight lines f(H_2)-values are lower than those common in the Milky Way for paths with similar degrees of reddening. | astro-ph | astro-ph |
FUSE Measurements of Far Ultraviolet Extinction. II. Magellanic
Cloud Sight Lines1
Stefan I. B. Cartledge1, Geoffrey C. Clayton1, Karl D. Gordon2, Brian L. Rachford3, B. T.
Draine4, P. G. Martin5, John S. Mathis6, K. A. Misselt2, Ulysses J. Sofia7, D. C. B.
Whittet8, and Michael J. Wolff9
ABSTRACT
We present an extinction analysis of nine reddened/comparison star pairs in
the Large and Small Magellanic Clouds based on Far-Ultraviolet Spectroscopic
Explorer (FUSE) FUV observations. To date, just two LMC sight lines have
probed dust grain composition and size distributions in the Magellanic Clouds
using spectral data for wavelengths as short as 950 A. We supplement these two
with data from 4 regions distinguished by their IR through UV extinction curves
and grouped as LMCAvg, LMC2, SMC bar and SMC wing. Despite the distinct
characters of extinction in the Clouds and Milky Way, our results are generally
analogous to those found for Galactic curves -- namely, that the FUSE portions
of each extinction curve are described reasonably well by Fitzpatrick & Massa
curves fitted only to longer wavelength data and lack any dramatic new extinc-
tion features, and any deviations from the Cardelli, Clayton, & Mathis (CCM)
1Department of Physics and Astronomy, Louisiana State University, Baton Rouge, LA 70803; scar-
[email protected], [email protected]
2Steward Observatory, University of Arizona, Tucson, AZ 85721; [email protected], mis-
[email protected]
3Center for Astrophysics and Space Astronomy, Department of Astrophysical and Planetary Sciences,
University of Colorado at Boulder, Campus Box 389, Boulder, CO 80309-0389; [email protected]
4Princeton University Observatory, Peyton Hall, Princeton, NJ 08544; [email protected]
5Canadian Institute for Theoretical Astrophysics, University of Toronto, Toronto, Ontario M5S 3H8
Canada; [email protected]
6University of Wisconsin-Madison, 475 North Charter Street, Madison, WI 53706; [email protected]
7Department of Astronomy, Whitman College, Walla Walla, WA 99362; [email protected]
8Department of Physics and Astronomy, Rensselaer Polytechnic Institute, Troy, NY 12180-3590;
[email protected]
9Space Science Institute, 3100 Marine Street, Suite A353, Boulder, CO 80303-1058; [email protected]
-- 2 --
formalism continue into FUV wavelengths. A Maximum Entropy Method analy-
sis of all of these curves suggests that LMCAvg and SMC wing sight lines, whose
extinction parameters more closely resemble those for Galactic paths, require
more silicon and/or carbon in dust than current abundance measurements would
indicate are available. The requirements for LMC2 and SMC bar sight lines do
not fully tax the available reservoirs, in part because large grains contribute less
to the extinction in these directions. An intermediate product of this extinction
analysis is the measurement of new H2 abundances in the Magellanic Clouds.
Collectively considering Cloud sight lines that possess significant H2 column den-
sities, E(B −V )/N(H I) ratios are reduced by significant factors relative to the
Galactic mean, whereas the corresponding E(B −V )/N(H2) values more closely
resemble their Galactic counterpart. These trends reflect the fact that among
these sight lines f (H2)-values are lower than those common in the Milky Way for
paths with similar degrees of reddening.
Subject headings: ISM: abundances -- ultraviolet: ISM
1.
Introduction
An important step toward a complete understanding of the formation, structure, and
composition of interstellar dust is the study and discrimination of factors that produce
changes in the observed wavelength-dependent extinction of stars due to grains. The Large
and Small Magellanic Clouds (LMC and SMC, respectively) are metal-poor relative to the
Galaxy; recent estimates imply that the LMC and SMC have overall metallicities at levels
0.5 and 0.2 -- 0.25, respectively, of the Galactic ISM (Welty et al. 2001). Dust-to-gas ratios
determined for these galaxies, as represented by E(B −V )/N(H I), are also reduced (by
factors of about 4 and 10; Bouchet et al. 1985; Fitzpatrick 1986), indicating that LMC and
SMC dust components and their formation mechanisms are not inconsistent with Galactic
components and mechanisms. Nevertheless, there are differences in extinction among these
three galaxies.
Early investigations of the extinction properties of Magellanic Cloud dust identified
several sight lines that exhibited marked differences from curves produced by grains in the
Milky Way (e.g., Nandy et al. 1980; Koornneef & Code 1981; Rocca-Volmerange et al. 1981).
1Based on observations with the NASA-CNES-CSA Far-Ultraviolet Spectroscopic Explorer, which is op-
erated for NASA by the Johns Hopkins University under NASA contract NAS-32985.
-- 3 --
It was noted in comparing these curves that the 2175 A bump strength was reduced in the
LMC relative to the Galaxy, and somewhat further diminished for curves characteristic of
the SMC. Similarly, the FUV rise evident in LMC and SMC extinction curves is stronger
than that common in the Galaxy. Subsequent studies have identified regional variation in
Magellanic Cloud curves, associating a distinctive LMC wavelength dependence with the
supershell LMC2 (Misselt, Clayton, & Gordon 1999), a steeper UV curve with the star-
forming bar region of the SMC, and extinction more closely resembling Galactic curves with
other portions of the Clouds. Recently, Gordon et al. (2003) (hereafter G03) completed a
comprehensive comparison of Galactic, LMC, and SMC extinction curves from near-infrared
to ultraviolet (UV) wavelengths (λ ∼> 1150 A or 1/λ ∼< 8.7µm−1); G03 confirmed that
curves characteristic of the LMC2 grouping and the SMC bar do not conform to the Cardelli,
Clayton, & Mathis (1989) (CCM) parameterization of Galactic extinction based on RV [≡
A(V )/E(B −V )].
The strength of their UV extinction is a key feature distinguishing LMC, SMC, and
Galactic curves; in particular, the steeper short-wavelength slopes associated with Magel-
lanic Cloud extinction curves can be diagnostic of differences in dust compositions and grain
size distributions, given appropriate measures of elemental abundances in their respective
interstellar media (Draine 2003). However, the extension of Magellanic Cloud extinction
curves into the FUV (> 8.7 µm−1) is complicated by the requirement that interstellar H2
absorption be removed from the spectra before an extinction curve can be produced us-
ing the pair method. Previous efforts in this area, even for Milky Way sight lines, have
been hampered by a lack of high quality data. In particular, instrumental issues of scattered
light, time-variable sensitivity, and a limited sample (both target and comparison objects) re-
stricted the utility of the Copernicus dataset (cf. Jenkins et al. 1986; Snow, Allen, & Polidan
1990), the Voyager UVS data suffer from low resolution which prevents explicit identification
and removal of the molecular hydrogen contribution, and the remaining available data [e.g.,
rocket - Green et al. 1992; Lewis, Cook, & Chakrabarti 2005; Hopkins Ultraviolet Telescope
(HUT) - Buss et al. 1994; Orbiting Retrievable Far and Extreme Ultraviolet Spectrometer
(ORFEUS) - Sasseen et al. 2002] include only a few Galactic sight lines. Far Ultraviolet
Spectroscopic Explorer (FUSE), however, has revolutionized our ability to study interstellar
molecular hydrogen. Its combination of slightly better spectral resolution and much greater
detector sensitivity than Copernicus allows H2 to be detected throughout much larger and
more diverse portions of the Galaxy and even in the Magellanic Clouds (Shull et al. 2000;
Rachford et al. 2002; Tumlinson et al. 2002). Thus, FUSE data are uniquely suited to inves-
tigating the FUV extinction properties of Milky Way, LMC, and SMC sight lines and what
any differences reveal about the dust populations in each galaxy.
Recently, Clayton et al. (2003) used a modified maximum entropy method (MEM) to
-- 4 --
fit extinction components to IR through UV curves for stars both in the Galaxy and in the
Clouds; they demonstrated that the average SMC bar extinction curve was best fit using a
grain distribution in which small silicate and amorphous carbon grains play a much larger
role than in the mean Galactic extinction. However, this result is at odds with Welty et
al. (2001), whose data suggested negligible silicon depletion in several gas cloud components
along a sight line probing the SMC ISM. In order to address issues such as dust composition,
particularly for small grains enhancing FUV extinction in the Clouds, it would be helpful
to explore LMC and SMC extinction curves shortward of the UV range that Clayton et al.
(2003) were confined to by their International Ultraviolet Explorer (IUE) data; unfortunately,
FUV extinction curves have been published to date for only very few reddened Magellanic
Cloud stars (Clayton et al. 1996; Hutchings & Giasson 2001).
In this paper, we construct new extinction curves for nine pairs of reddened and unred-
dened LMC and SMC stars by supplementing the IR through UV curves published by G03
with recent FUSE observations of 16 Magellanic Cloud targets. The FUSE data allow us to
extend the Clayton et al. (2003) analysis of dust composition and grain size distribution to
FUV wavelengths, where extinction is dominated by small grains and the differences between
Galactic and Magellanic Cloud curves are particularly prominent. This is the second in a
series of three papers exploring extinction in the FUV. Paper I (Sofia et al. 2005) dealt with
a small set of Galactic extinction curves characterized by a broad range of RV -values; Paper
III (in preparation) will examine a much larger sample of Galactic sight lines in an effort to
search rigorously for any trends among observed extinction parameters when FUV data are
fully considered.
2. Observations and Data Extraction
The present sample of 16 LMC and SMC stars includes all stars in the FUSE archive for
which the spectral match between the reddened and unreddened stars has been rigorously
evaluated and there exist spectral data from IR to UV wavelengths. Each reddened star
has previously been studied by G03; we use the same photometry sources and procedures to
construct the IR to UV portions of the extinction curves for the pairs adopted in this study,
with the exception that early 2MASS photometry has been superseded in our analysis by
data from the recent All Sky Release. In particular, it should be noted that the RV values
listed in Table 1, which summarizes the properties of each extinction pair, differ slightly from
those published by G03 due to this update of the IR data. Consequently, the values of AV
and N(H I)/AV in the table are also affected. The FUV portions of the extinction curves
(shortward of the H I Lyα line) were derived from the new and archival FUSE observations
-- 5 --
listed in Table 2. All of the raw data were processed using the latest release of CALFUSE
(v3.0). The calibrated spectra for each channel were cross-correlated, shifted to a common
wavelength scale, and combined based on exposure time in cases where several observations
of a star existed. Also, in cases where the different observations or different FUSE channels
produced conflicting flux levels for a given wavelength (e.g., due to channel misalignment),
the data for these observations or channels were linearly adjusted to match the maximum
observed flux across the regions of overlap. General details on FUSE observations have
been published by Sahnow et al. (2000a). We point out here, however, that each observation
produces eight spectra nearly 100 A in length with individual dispersion characteristics. The
LiF channel spectra (LiF1A, LiF1B, LiF2A, LiF2B) cover most wavelengths from 990 -- 1190
A twice over; the spectral region from 910 -- 1110 A are similarly sampled by the SiC channels
(SiC1A, SiC1B, SiC2A, SiC2B). Sample spectra of both poor and fine quality appear in
Figure 1, using the data from the reddened star AzV 456 and its comparison star AzV 70,
respectively.
2.1. Molecular Hydrogen Modelling and Removal
Strong FUV H2 absorption bands can significantly alter the appearance of stellar spectra,
depending upon the column density of molecular gas along the line of sight. Consequently,
a molecular hydrogen model was constructed for each sight line using procedures patterned
generally after those used by Rachford et al. (2001, 2002). The underlying assumption we
made in measuring the interstellar H2 for each sight line was that the model includes only
two velocity components, one corresponding to Milky Way gas and one associated with H2 in
the appropriate Magellanic Cloud. The equivalent width measurements for distinct profiles
of each component were entered into separate tables, so that molecular hydrogen measure-
ments could be made individually for each galaxy using a curve-of-growth (CoG) analysis.
Unfortunately, the velocity separation of Galactic and Magellanic Cloud components often
resulted in blended absorption profiles at several points in a given spectrum. In general, the
construction of a reasonable CoG for most of the sight lines required more equivalent width
measurements than were available from the unblended and isolated H2 absorption lines. To
remedy this problem, the profile-fitting code FITS6P2 and IRAF3 plotting routines were
2The FITS6P code models absorption profiles based on varying the column densities, b-values, and veloc-
ities of input interstellar components. More details on the algorithm can be found in Welty, Hobbs, & York
(1991).
3IRAF is distributed by the National Optical Astronomy Observatories, which are operated by the As-
sociation of Universities for Research in Astronomy, Inc., under cooperative agreement with the National
-- 6 --
used in conjunction with our IDL4 code (Rachford et al. 2001, 2002) to deconvolve blended
profiles and fill out each CoG with as many measurements as could reliably be made. The
CoG analysis, however, was generally limited to J ≥ 3 because the J ≤ 2 lines were sat-
urated to the extent that they possessed broad damping wings that the equivalent width
measurement algorithm could not reliably distinguish from the stellar continuum. Thus, an
iterative curve-fitting procedure was an additional requirement so that simultaneous deter-
minations of Galactic and Magellanic Cloud column densities for these lower H2 rotational
excitation levels could be made. The complete molecular hydrogen models, including col-
umn densities for the available J levels, the associated b-value, and the derived kinetic and
excitation temperatures (these last properties are discussed in section 3.1), are summarized
in Tables 3 and 4 for all of our LMC and SMC sight lines; Table 3 characterizes the Galactic
component for each sight line where one was detected, while Table 4 details the properties
of each Magellanic Cloud component.
Once molecular hydrogen absorption models had been constructed, each spectrum was
corrected for this absorption. All of the FUSE data for each star were combined into a single
spectrum by shifting the calibrated channel spectra to a common wavelength reference and
averaging the fluxes, weighted by the flux uncertainty at each point in the overlapping
regions. Although there are significant differences in the data dispersion properties for each
channel, the merged spectrum for each star is of sufficient quality for the construction of a
reliable extinction curve. The "worm" problem in FUSE data (Sahnow et al. 2000b) was
strong only in LiF1B. Its appearance in the data for this detector segment was identified
by a comparison with the LiF2A flux as the data for all detector segments were merged
into a single spectrum for each sight line; the portions of the LiF1B spectrum contaminated
by the worm were eliminated from the merging process. A running cross-correlation of
the normalized H2 absorption profile corrected for any wavelength mismatches, and the
full final model was divided into the merged FUSE spectrum to complete the "removal" of
the molecular hydrogen features. Atomic hydrogen was not measured for each sight line
individually, but its signature, specifically the Lyα and Lyβ lines, was removed after the
ratio of the reddened and comparison spectra was calculated as a function of wavelength.
Although FUSE spectra cover wavelengths from 905 through 1187 A, data shortward of 1044
A (our Lyβ cutoff) were not included in this analysis due to limitations in data quality.
In the interest of consistency in instrument response, FUSE flux levels were then man-
Science Foundation.
4IDL is an acronym for Interactive Data Language, a common programming tool developed by Research
Systems, Inc.
-- 7 --
ually rescaled to match IUE fluxes over the wavelength range common to both instruments
(1150 − 1185 A) when the full IR to FUV spectra were constructed; given recent improve-
ments to the calibration of IUE data (Massa & Fitzpatrick 2000), flux errors associated with
this instrument have not yet been surpassed by the latest FUSE calibration (Sahnow et al.
2000a). Generally, mean fluxes for the data from each instrument agreed to within 20%,
although there were a few targets for which the discrepancies exceeded this range. The scal-
ings applied to each FUSE spectrum (the ratio of IUE to FUSE flux in the overlap region)
are given in Table 5.
2.2. The Extinction Curves
We have constructed extinction curves using the pair method (e.g., Massa, Savage, &
Fitzpatrick 1983), normalized to A(V ) through RV ; the pairs of reddened and unreddened
stars in the LMC and SMC were selected (G03) and are listed in Table 1.
In an effort
to minimize uncertainties in each extinction curve, particularly from spectral mismatch, a
detailed comparison of the FUSE spectrum for each star with similarly-typed unreddened
candidates was performed, and the extinction pairs matched by G03 using IUE were also
found to be well-matched in the FUV. Assembly of the extinction curve corresponding to
each pair was accomplished using the same procedures and near-infrared through UV data
outlined by G03. We have also constructed average extinction curves for the subsets of the
current sample associated with the LMCAvg and LMC2 regions considered previously in the
UV (Misselt, Clayton, & Gordon 1999; G03). Each of the individual LMC curves and the
average curves have been fit to the Fitzpatrick & Massa (1990) parameterization; the results
are presented in Table 6. The two sight lines in the SMC, one representing each of the bar
and wing regions, are considered individually.
A quick examination of the individual Magellanic Cloud extinction curves presented
in Figure 2 reveals that the FUSE portion of each curve is generally consistent with an
extension of the data from longer wavelengths. When the calibrated and H2-adjusted fluxes
were initially compared, however, slight offsets were apparent between the portions of the
curve on either side of the Lyα line. Comparing the size of the gaps in the Magellanic Cloud
data with the Galactic curves studied in Paper I led to the conclusion that the quality of the
data for our reddened stars might have compromised the accuracy of the extinction curve
construction. In particular, the S/N of the FUSE data seemed to be inversely proportional
to the size of the offset and there were noticeable flux level differences for some spectra
between the IUE and the merged FUSE data in the spectral overlap region. The Galactic
sight lines had much better S/N values, negligible offsets, and good agreement between IUE
-- 8 --
and FUSE fluxes, whereas the Magellanic Cloud data suffered generally poor S/N, noticeable
offsets, and sometimes poor flux agreement in the overlap. In order to reduce the offsets,
the FUSE data for the Cloud sight lines were linearly adjusted (a wavelength-independent
multiplicative factor was applied to the flux) over the entire FUV bandpass so that their
fluxes matched IUE in the instruments' overlap region. An example of one of the more
egregious cases is plotted in Figure 3, including a comparison of the FUSE flux levels both
before and after adjustment to the IUE flux level. The magnitude of each shift was estimated
in increments of 0.05 × f luxIUE; the systematic uncertainty associated with this adjustment
has been propagated through all subsequent error calculations. Typically, the shifts improved
the appearance of the curves, although the offsets were not eliminated in all cases. Given the
general smoothness of the transition between UV and FUV portions of the extinction curve
in most cases, it might seem reasonable to introduce further shifts to align these segments;
nevertheless, because the observed offsets fall generally within the error bounds, because we
are interested in the detailed shape of the curve at UV and FUV wavelengths, and since we
have no independent objective basis for such shifts, no further processing of the individual
curves has been performed. It should also be noted that any residual offsets appear to be
biased in the direction of decreased A(λ)/A(V ); thus, abundance requirements derived from
the FUV data may be slight underestimates.
3. Discussion
3.1. H2 in the Magellanic Clouds
As already mentioned, a notable difficulty in deriving extinction curves that extend into
the FUV is the existence of strong molecular hydrogen absorption bands longward of the
Lyα line. Nevertheless, stellar observations of sufficient quality to allow the H2 absorption
to be modelled and removed not only permit the construction of an extinction curve but
also provide more direct information on the intervening ISM. Recently, Tumlinson et al.
(2002) completed a survey of molecular hydrogen in the LMC and SMC using several FUSE
observations of each Cloud; our sample introduces 15 new sight lines into the mix.
In considering these new molecular hydrogen measurements, however, it should be rec-
ognized that the current FUSE data generally have much lower signal-to-noise ratios than
Galactic observations. In fact, they are also somewhat lower than the values for previously-
observed Magellanic Cloud targets. Tumlinson et al. (2002) reported typical 4σ equivalent
width limits of 30 -- 40 mA, a level similar to values of 20 -- 50 mA for our comparison stars
but much lower than the 50 -- 120 mA characteristic of the reddened targets. Because of these
data limitations, and because the goal of generating a reliable FUV continuum was given
-- 9 --
a somewhat higher priority than making a rigorous assessment of molecular hydrogen, any
column densities smaller than about 1017 cm−2, and a few that are somewhat larger, are
subject to sizable uncertainties. Nevertheless, we consider the models listed in Tables 3 and
4 to be reasonable tallies of the H2 along each sight line.
In order to more fully understand how the new sight lines mesh with the Tumlinson et al.
(2002) Magellanic Cloud H2 analysis, we have adopted the atomic hydrogen measurements
previously published by Fitzpatrick (1985a,b) for the individual sight lines (see Table 7);
for the pair listings in the table, we use the values derived directly from the extinction
curves by G03. It should be noted here that the process of determining atomic hydrogen
column densities for the Magellanic Clouds is complicated by the large quantities detected
along each sight line. The breadth of the Lyα absorption profile produced by typical inter-
stellar hydrogen distributions with line-of-sight lengths exceeding even a few kiloparsecs is
sufficient to mask distinctions between separate velocity components; in particular, the pro-
files for atomic hydrogen gas in the LMC or SMC are often inextricably blended with their
Galactic counterparts. Tumlinson et al. (2002) dealt with this problem by calibrating 21 cm
emission measurements to Lyα from the sight lines in their sample for which the Galactic
and Magellanic Cloud 1215 A profiles were not severely blended. Atomic hydrogen column
densities for the remaining paths in the data set were then estimated using this calibration
and the observed 21 cm emission. G03 assessed Magellanic Cloud H I column densities by
measuring the Lyα profile in the reddened-to-comparison ratio spectrum for each extinction
pair, under the assumption that the Milky Way components would cancel each other. Each
of these methods is subject to a significant level of uncertainty; if vertical error bars were
to be plotted in Figure 4, among the Magellanic Cloud sight lines they might typically be
about 0.3 dex and occasionally larger than 1.0 dex. Nevertheless, differences between the
Fitzpatrick (1985a,b) H I values for the individual sight lines in each extinction pair generally
are well matched by the corresponding column densities derived by G03.
Considered individually, our LMC and SMC sight lines and those studied by Tumlinson
et al. (2002) exhibit similar ranges in characteristics such as the molecular hydrogen fraction
f (H2) (≡ 2N(H2)/[N(H I) + 2N(H2)]) and the Magellanic Cloud portion of the color excess
E ′(B −V );5 values for the two LMC sight lines studied by Gunderson, Clayton, & Green
(1998) also match these samples. Since the current dataset was selected for its extinction
properties, we cannot speak to the overall frequency of molecular hydrogen detection in
the Clouds. However, as shown in Figure 4, the f (H2)-values for paths along which we
detect a Cloud-based H2 component agree with the levels set by previously observed LMC
5Following the procedure of Tumlinson et al. (2002), we adopt the definition E ′(B −V ) = E(B −V ) − x,
where x is 0.075 and 0.037 for LMC and SMC paths, respectively.
-- 10 --
and Galactic sight lines of large E(B −V ). Notably, the sight lines toward AzV 456 and
Sk −67◦2 possess larger molecular hydrogen fractions than any examined by Tumlinson et
al. (2002); meanwhile, among the paths directed toward our comparison stars we detect no
Magellanic H2 components.
Based on the previously-studied sight lines toward Sk -66◦19 and Sk -69◦270, it was
found that while the gas-to-dust ratio N(H I)/E ′(B −V ) is much larger in the Clouds than
typically seen in the Milky Way, the quantity N(H2)/E ′(B −V ) has a similar value in each
galaxy (Clayton et al. 1996; Gunderson et al. 1998). This suggests a close relationship
between dust and H2. Similar behavior is evident for our new sight lines, since although
the ratios N(H I)/E ′(B −V ) for these paths are roughly 4 -- 10 times larger than the Galactic
mean, comparing H2 column densities with the same color excesses leads to a range of values
around the Galactic average. In this way our new sight lines complement the paths studied
by Gunderson et al. (1998) and the more heavily reddened paths of Tumlinson et al. (2002),
implying generally lower values for f (H2) in the Clouds; nevertheless, the current dataset
includes sight lines with unique properties. Plotting the LMC N(H I)/E ′(B −V ) ratio as a
function of E(B − V ) (see Figure 5), there is a clear distinction between Sk -67◦2 and other
reddened sight lines: this one LMC path is characterized by an "atomic" ratio appropriate
to Galactic curves, but because of its large molecular hydrogen column density the total
gas-to-dust ratio approaches the values for other LMC sight lines. Coupled with its Milky
Way-like N(H I) ratio, the agreement exhibited between the gas-to-dust ratios for Sk -67◦2,
the sight line with the largest f (H2)-value in our sample, and sight lines dominated by atomic
gas reinforces the concept that the mechanisms governing dust formation in the Galaxy
and LMC are similar and that any differences in extinction arise out of how interstellar
environmental conditions are manifested in dust grain population characteristics. The SMC
sight line AzV 456 also possesses a "Galactic" value for N(H I)/E ′(B −V ) that gives a
N(HTotal)/E ′(B −V ) ratio more closely resembling the values determined by Tumlinson et
al. (2002) for other SMC paths.
Further details of the interstellar conditions, specifically the temperatures T01 and Tex,
in the Clouds can be derived from the relative rotational level populations of the individual
sight lines (Table 4). When the cloud density is sufficiently high and H2 column densities
are large enough for the gas to become substantially molecular, ortho-para conversion will
bring T01 (≡ E1−E0
ln(9N (0)/N (1)) ) close to the kinetic temperature.
Except for the AzV 462 sight line, its value is fairly uniform among the LMC and SMC
components of our H2 models and we derive a weighted mean temperature of 50±3 K. The
Magellanic Cloud component of H2 toward AzV 462 has the lowest column density among
our extragalactic values, and the assumption that this value is consistent with the kinetic
temperature breaks down. The mean temperature we have derived for the Magellanic Cloud
log(N (J=0)/g0)−log(N (1)/g1) =
k
log(e)
170K
-- 11 --
components neglects AzV 462 and is somewhat lower than the overall Galactic mean of
77±17 K (Savage et al. 1977), but is consistent with both the value 55±8 K derived from
the Galactic subsample for which N(H2) > 20.4 (Rachford et al. 2002) and temperatures
derived by Tumlinson et al. (2002) for Magellanic Cloud gas along sight lines with some of
the largest molecular column densities. It should be noted that the errors ascribed to each
temperature listed in Tables 3 and 4 were derived from column density uncertainties, which
were often large. Thus, kinetic temperatures with uncertainties larger than about 60 K should
be regarded as less reliable. In particular, the formal uncertainty is listed in each case, even
where it exceeds the calculated temperature, in order to indicate the relative reliability of
each determination. We have also determined Tex for each sight line from the slope in a plot
of N(J = 2, 3, 4, 5) versus the excitation potential.6 This excitation temperature reflects the
influence of fluorescent UV pumping and H2 formation in excited states on the molecular
hydrogen rotational level population (Jura 1975; Black & Dalgarno 1976). Like T01, it is
fairly uniform among the LMC and SMC sight lines and we calculate a weighted mean of
244±33 K, similar to previously-determined values for Magellanic Cloud sight lines (Shull et
al. 2000). Comparing the column density ratios N(4)/N(2) and N(5)/N(3) for our sample
with Figure 11 from Tumlinson et al. (2002), we find that these sight lines are reasonably
typical of their high N(H2) paths; consequently, our measurements support their conclusions
that the H2 formation rate in the Clouds is low (10 -- 40% of Galactic rates) and that at least
some regions are illuminated by relatively intense radiation fields (10 -- 100 times the Galactic
mean). We note, however, that the proportion of outliers in our sample that do not fit the
range of Galactic models calculated by Tumlinson et al. (2002) is much smaller than theirs.
Values of T01 and Tex were also derived for the Galactic components of our H2 models. A
mean of T01 = 67±4 K was derived for the sight lines with larger column densities and more
robust measurements, although higher, and likely to be in part non-thermal, temperatures
are implied for several of the more diffuse paths. Excitation temperatures for the Galactic
components (239±38 K) are similar to those we have determined for the Magellanic Clouds.
3.2. FUV Extinction in the Magellanic Clouds
Before FUSE, only two reddened sight lines outside the Galaxy had been studied in the
FUV (Clayton et al. 1996). Both regional extinction groups in the LMC are represented
6More generally, Tex is associated with values of TJ −2,J ≡ EJ −EJ−2
log[N (J −2)/gJ−2]−log[N (J)/gJ ] =
(for J ≥ 2), the temperature derived from comparing the populations of successive even or
170K×(2J −1)
N (J−2)
ln[ 2J+1
2J−3
odd J levels.
N (J)
]
k
log(e)
-- 12 --
by these two stars; specifically, Sk -66◦19 is associated with LMCAvg and Sk -69◦270 with
LMC2 (Misselt et al. 1999). The UV extinction curves for these stars are typically non-
CCM, but they share the trait of most Galactic curves in that their UV extinction slope
blends smoothly into the FUV. These two sight lines have been included for comparison in
Table 6, which defines the LMC2 and LMCAvg subsamples, although they were left out
of the average extinction curves for those regions. Hutchings & Giasson (2001) published
FUV extinction curves for 3 sight lines in each of the LMC and the SMC using FUSE data.
Unfortunately, the values of ∆(B-V) for the reddened/unreddened star pairs chosen for that
study are too small (∼0.02-0.07 mag) for detailed extinction studies. More specifically,
these reddening values are comparable in size to the variation in the foreground E(B −V )
and similar to the magnitude of the photometric uncertainties (Schwering & Israel 1991;
Oestreicher, Gochermann, & Schmidt-Kaler 1995; Massey 2002). So, within the uncertainties
there is no significant reddening difference for any of the star pairs in Hutchings & Giasson
(2001), and those sight lines are not included in this study.
The new LMC and SMC FUV extinction curves presented here seem to follow the trend
of Sk -69◦270 and Sk -66◦19. Despite any small remaining offsets between the IUE and FUSE
extinction curves (see § 2.1), the new FUV curves closely follow an extrapolation of their
FM fits from the UV. Out of our sample, only two curves exhibit significant deviations from
this overall trend, AzV 456 and Sk -69◦228. AzV 456 is our lone reddened star situated in
the SMC wing, and the FUSE portion of its extinction curve more closely resembles its FUV
CCM curve than an extrapolation of its FM fit in the UV. Coincidentally, the disagreement
with the FM fit begins near the onset of molecular hydrogen absorption features; however,
attempting to reconcile the two portions of the FUSE curve by adjusting the H2 column
densities indicates that a single component for the Clouds cannot reproduce the observed
continuum level shift, nor does it appear that a simple combination of overlapping compo-
nents would accomplish this result. The effect does not appear to be related to a difference in
the sensitivity of the various FUSE channels either, since the continuum across this spectral
region is contained in both the LiF1B and LiF2A detector bands. Similar kinks are apparent
to a much smaller degree in the Galactic curves for HD 62542, HD 73882, and HD 210121
(Paper I); nevertheless, those curves match the FM fits within the formal uncertainties [ap-
proximately 14% in A(λ)/A(V) across the FUSE band] and we feel that it would not be
appropriate to speculate further about the origin of these features until a larger data sample
is compiled.
The FUV extinction curve for Sk -69◦228 is very noisy, which would usually indicate a
poor match with its unreddened comparison star, but this same star is a good match through
the UV. The problem seems to arise with a complex of photospheric Fe III lines evident
in the comparison star spectrum, Sk -65◦15. An examination of intermediate continuum
-- 13 --
segments demonstrates that the reddened and comparison stellar spectra have very similar
shapes aside from these lines, and that connecting the lower wavenumber portion of this
star's FUV extinction curve with the two highest wavenumber points would be a reasonable
approximation. Notably, this procedure results in an FUV curve that is consistent with an
extension of the UV FM fit, despite the impression that a small offset is still present after
the FUSE flux adjustment. The curve is still included in further analysis for the sake of
completeness and because our comparison sample is limited, but we note that the mean
curve derived for the LMC2 group, which includes this path, is consistent with the other
curves in the group unaffected by any taint of mismatch. On the whole, the ease with which
the UV and FUV extinction curves line up for our Magellanic Cloud sight lines complements
the results for Milky Way sight lines; namely, that most Galactic sight lines also seem to
show FUV extinction that is a good extrapolation of the FM fits made to the extinction
curves in the UV. Agreement between the CCM relation and FUV extinction, however, has
proved less ubiquitous. For instance, Buss et al. (1994) found that the FUV extinction for
two Galactic sight lines with large (ρ Oph) and small (HD 25443) values of RV follow CCM
closely, but that there are exceptions; one notable example is the bright-nebula sight line
toward HD 37903 which shows a steeper extinction in the FUV, based on Copernicus data,
than would be predicted by CCM. Paper I, which included paths such as HD 210121 and
HD 62542 that do not follow CCM in the UV, came to similar conclusions. Although the
UV extinction curves for each sight line they studied could be smoothly extrapolated into
the FUV using an FM fit, only three paths that followed CCM through UV wavelengths
were also in accord across the FUV. The CCM relations for four others from Paper I either
over or underestimated their FUV extinction, as small discrepancies in the UV became more
pronounced at shorter wavelengths, and the two non-CCM curves diverged more strongly
from their RV -based curves with increasing wavenumber. Because the LMC and SMC curves
do not have a CCM-like wavelength dependence and because the S/N ratios of the individual
extinction curves are relatively low, subtle variations in the FUV slope such as those seen
by Buss et al. (1994) cannot be ruled out.
3.3. FUV MEM Modelling: Grain Properties
To increase the S/N for MEM modelling purposes, we have constructed average extinc-
tion curves for the LMC average and LMC2 regions outlined by Misselt et al. 1999 (see
Table 6); the curves are plotted in Figure 6. The two sight lines in the SMC, which rep-
resent the bar and wing regions, are considered individually. Using the two average LMC
curves and the two SMC curves, we have extended the dust-grain population analysis of
Clayton et al. (2003) into the FUV. We employ a (slightly) modified version of the MEM
-- 14 --
extinction-fitting algorithm (Kim et al. 1994; Kim & Martin 1996).
Instead of using the
number of grains of a given size to describe the dust population, the algorithm employs the
mass distribution in which m(a)da is the mass of dust grains per H atom in the size interval
from a to a+da. Thus, the traditional MRN-type model (Mathis, Rumpl, & Nordsieck 1977)
becomes m(a) ∝ a−0.5. We use a power law with exponential decay (PED) as the template
function for each component. The data are examined at 34 wavelengths, and the grain cross
sections are computed over the range 0.0025 -- 2.7 µm with 50 logarithmically spaced bins.
The shape of the mass distribution is strongly constrained only for data over the interval
0.02 -- 1 µm. Below 0.02 µm, the Rayleigh scattering behavior constrains only total mass;
above 1 µm, the "grey" nature of the dust opacity also forces the MEM algorithm to simply
adjust the total mass, using the shape of the template function to specify the size dependence
of the distribution. The total mass of dust is constrained using both the gas-to-dust ratio
and "cosmic" abundances (i.e., we try not to use more carbon or silicon than is available).
The elemental abundance standards used in this analysis are those adopted in Paper I for
the Galaxy (358 and 35 atoms per million H for C and Si, respectively), and by Clayton
et al. (2003) for the LMC and SMC (110 and 65 for the LMC and 54 and 11 atoms per
million H for the SMC); Table 7 reports the values of the gas-to-dust ratios used here for
the LMC and SMC. We consider only three-component models of homogeneous, spherical
grains: modified "astronomical silicate" (Weingartner & Draine 2001), amorphous carbon
(Zubko et al. 1996), and graphite (Laor & Draine 1993). While it must be acknowledged
that the three grain component system we have used to model these extinction curves is
simpler than might be expected of actual interstellar dust, the results can be quite useful for
identifying grain population properties that distinguish sight lines from one another. The
same component-specific PED constraints (e.g., the onset of the exponential cutoff) adopted
in Paper I were utilized for the current modelling.
The MEM fits to FM parameterizations of the average Magellanic Cloud extinction
curves listed in Table 6 are presented in Figures 7 and 8. The first plot shows the amount of
extinction provided by the three distinct grain components, as well as the total extinction
of the model, compared to the FM fit associated with each extinction curve. The error bars
plotted on these fits are indicative of the mean gap between the FM fit and the underlying
extinction curve in each wavelength bin. The fractions of the adopted elemental abundances
available for each grain component that are required by the best MEM fit are listed in
Table 8. Figure 8 depicts the corresponding mass distributions for different sizes of grains
belonging to each model component relative to the mass of hydrogen.
The proportions of the available silicon and carbon used in the MEM fits to the various
FM parameterizations cover a very wide range, as was seen in fits to IR through UV data
alone by Clayton et al. (2003). Three general factors determine the fraction of silicon and
-- 15 --
carbon that any individual sight line will use. First, the higher the gas-to-dust ratio is, the
more metals are available in the gas phase. Second, the higher the abundances of metals
are, the more material is available. Finally, high values of RV imply a greater than average
mass fraction in larger grains which are not as efficient per unit mass as smaller grains. For
instance, it can be seen from the MEM fits that the SMC wing (AzV 456), which has a
low gas-to-dust ratio and low elemental abundances relative to the Galaxy, uses more than
100% of the available silicon. However, the LMC2 and the SMC bar (AzV 18) regions, which
also have both low elemental abundances but are characterized by higher gas-to-dust ratios,
use less than half the amount of available silicon that the LMCAvg and SMC wing regions
require and also significantly less carbon.
A recent observation has suggested that silicon is relatively undepleted in the SMC bar
(Welty et al. 2001). However, model fits to the average SMC bar extinction indicate that its
curve cannot be fitted with carbon grains alone (Weingartner & Draine 2001; Clayton et al.
2003). Figures 7 and 8 clearly show that this conclusion is born out in the current fits. SMC
extinction for both bar and wing is very steep in the FUV, necessitating the presence of a
large population of small grains. Yet because the 2175 A bump is absent in the bar region,
small (a < 0.02µm) carbon grains do not meet the SMC FUV extinction requirements; thus,
silicates play a very important role in the models. Even our SMC wing sight line, whose
extinction curve follows CCM reasonably well, places a much higher demand on silicon
reserves than carbon: a typical Galactic curve conforming to CCM allows graphite to fulfill
a significant part in FUV extinction due to the strength of its UV bump, but the SMC wing
bump is somewhat weaker than the RV =2.19 CCM curve would imply. The next steepest
sight lines, in the LMC, do not show much difference in demand from a typical Galactic
curve (HD 14250 was chosen from Paper I for illustration purposes), once the gas-to-dust
ratio differences are taken into account. Typically, carbon grains are responsible for most of
the visible extinction and silicon grains for most of the UV extinction; therefore, in general,
both species of grains are needed along any sight line to get a good fit to the extinction
curve. Apart from deficits of large silicate and carbon grains in the SMC, the MEM analysis
of the Magellanic Cloud FUV extinction produces results similar to those derived from an
analysis of several Galactic sight lines (Paper I). The Galactic sight lines, like the SMC wing
or LMCAvg, are more likely to use more than 100% of the available silicon and/or carbon.
Among these sight lines, the strength of the 2175 A bump in relation to the FUV rise is
the factor that distinguishes whether the stiffer abundance demand is placed on silicon or
carbon. Likewise, the element most demanded by the SMC bar sight line and the mean
LMC2 curve is determined by this factor, although the minimal presence of the bump in
these curves allows the modelling to make more efficient use (especially in small grains) of
the available elements.
-- 16 --
The general structure in the size vs m(a) curves for extinction out to 8 µm−1 for the
Magellanic Cloud sight lines does not change much when the FUSE data are included.
Figures 7 and 8 showing the MEM fits and size vs m(a) curves cannot be directly compared
to Figure 3 of Clayton et al. (2003) because there are differences in the input parameters
such as the gas-to-dust ratios. But the MEM models run with the same input parameters
for extinction curve that are cutoff at 8 µm−1 and those shown here, which extend to 10
µm−1, are quite similar; there is at most a few percent difference in the amount of silicon
and carbon required by the two sets of models for the four SMC and LMC curves. The
structure evident in the mass distributions of Figure 8 particularly identify grain sizes for
which each component in our simple model can account for the detailed shape of each
extinction curve. For instance, the two peaks in the amorphous carbon distribution for HD
14250 correspond roughly to maxima appearing in the silicate distribution of Kim et al.
(1994) in their silicate-graphite model. These authors ascribe the peaks to the requirements
of fitting the optical and UV portions of the extinction curve using the spectral wavelength
dependence appropriate to each of the two grain components in their model. In the case
of AzV 456 (the SMC wing), however, efforts to reproduce the extinction curve using this
three-component model appear to require dramatic peaks and dips in the mass distribution
as a function of grain size. Aside from this curve, however, the scale of structure in the
distribution is consistent with levels noted by previous studies (Kim et al. 1994; Clayton
et al. 2003). For AzV 456, the solution requires that the graphite distribution be strongly
peaked in order to account for the 2175 A bump and provide as much extinction as possible
at FUV wavelengths. Since the FUV portion of this curve extends, or even amplifies, the
steep rise at the short-wavelength end of the previous IR through UV result (Clayton et al.
2003), there is no relaxation of the demand on silicon reserves. Specifically, when comparing
MEM solutions for the SMC wing curve with and without the FUSE data, nearly identical
proportions of the available silicon and carbon are utilized. The only notable distinction
between the two solutions is that a few percent of the amorphous carbon demand is shifted
to graphite when the FUV data are included. The extinction curves for our other groups
also smoothly extend from the UV into FUV wavelengths, and the models derived from IR-
to-UV and IR-to-FUV data require almost precisely the same amounts of silicon and carbon
be present in dust grains. The similarity in the MEM results for these two sets of models
may be due to the efficiency of the small grains in FUV extinction, in that large numbers
of such grains are not needed, or perhaps the observed amounts of FUV extinction can be
provided by the larger grains which are also important for extinction at longer wavelengths.
This issue will be investigated further as we study the distinction between global and sight-
line-specific FUV extinction characteristics of using a much larger number of sight lines, in
a future paper.
-- 17 --
4. Conclusion
This paper is the second in a series investigating the FUV characteristics of extinction
due to interstellar dust. The first paper dealt with a small sample of Galactic sight lines
with a variety of RV values, including paths whose extinction does not conform to the CCM
parameterization. This paper examined a similarly-sized set of sight lines probing the Small
and Large Magellanic Clouds, with the goal of determining how the addition of FUV data
to their corresponding extinction curves affected models of Magellanic Cloud dust grain
properties.
Using FUSE observations of four stars in the SMC and 12 stars in the LMC, we were able
to construct two extinction curves for the SMC and seven representing the LMC. Despite the
poor quality of the data for some of the reddened stars, molecular hydrogen was measured for
each sight line and their corresponding absorption features were removed from consideration
in the extinction analysis. Flux mismatches between IUE and FUSE data, likely due to
poor S/N values in the affected observations, were alleviated by rescaling the entire FUSE
spectrum. Comparison of the resulting FUV extinction with the curve constructed only
through wavelengths observed by IUE demonstrated that the FUV portions were generally
consistent with a smooth extrapolation of the IR-to-UV curve. Nevertheless, we note that
a "kink" appeared in the AzV 456 curve, similar to much smaller features present in curves
analysed in Paper I. We plan to address the existence and nature of this kink in a subsequent
paper examining the variation in properties of FUV extinction.
It was anticipated that including FUV data in the MEM analysis of the Magellanic Cloud
dust grain population might give rise to significantly different properties than Galactic dust
since the FUV rise in the Clouds is generally much steeper than it is in the Milky Way,
and the UV bump is smaller. However, aside from some complex structure likely driven by
the simplicity of our grain component model, the sight lines through the Clouds exhibitting
Galactic-like extinction are similar to Galactic paths in that they demand 100% or more
of the available carbon and/or silicon. The only real differences between the populations
for these galaxies are that the SMC wing sight line has a much smaller grain size cutoff
for astronomical silicates and amorphous carbon and a generally smaller dust mass. MEM
solutions for the sight lines through the Clouds that exhibited more distinctly Magellanic-
Cloud-type extinction were able to satisfy their curves using smaller proportions of the
elements available to them; the SMC bar solution, like that for the SMC wing, is characterized
by a smaller grain size cutoff and lower dust mass than Galactic or LMC solutions. The
addition of the FUSE data to the analysis does not dramatically alter the properties of each
dust population, except to shift small amounts of the carbon demand between amorphous
carbon and graphite.
-- 18 --
The molecular hydrogen abundances determined in the process of constructing the full
IR-to-FUV extinction curves supplement those already appearing in the literature. Although
the emphasis has been on eliminating H2 absorption features from the spectra rather than
deriving robust column density measurements, our results are consistent with the larger
recent survey of molecular hydrogen in the Magellanic Clouds by Tumlinson et al. (2002)
and we complement their paths by probing several sight lines with larger column densities.
Among these sight lines, E(B −V )/N(H I) ratios are generally reduced relative to Galactic
values while E(B −V )/N(H2) are roughly the same, implying somewhat lower f (H2)-values
in the Clouds than are typical in the Milky Way for similar degrees of reddening.
This study was supported by NASA grant NAG5-108185.
Black, J. H., & Dalgarno, A. 1976, ApJ, 203, 132
REFERENCES
Bouchet, P., Lequeux, J., Maurice, E., Pr´evot, L., & Pr´evot-Burnichon, M. L. 1985, A&A,
149, 330
Browning, M. K., Tumlinson, J., & Shull, J. M. 2003, ApJ, 582, 810
Buss, R. H., Jr., Allen, M., McCandliss, S., Kruk, J., Liu, J., & Brown, T. 1994, ApJ, 430,
630
Cardelli, J. A., Clayton, G. C., & Mathis, J. S. 1989, ApJ, 345, 245 (CCM)
Clayton, G. C., Green, J., Wolff, M. J., Zellner, N. E. B., Code, A. D., Davidsen, A. F.,
WUPPE Science Team, & HUT Science Team 1996, ApJ, 460, 313
Clayton, G. C., Wolff, M. J., Sofia, U. J., Gordon, K. D., & Misselt, K. A. 2003, ApJ, 588,
871
Diplas, A., & Savage, B. D. 1994, ApJ, 427, 274
Draine, B. T. 2003, ARA&A, 41, 241
Fitzpatrick, E. L. 1985a, ApJ, 299, 219
Fitzpatrick, E. L. 1985b, ApJS, 59, 77
Fitzpatrick, E. L. 1986, AJ, 92, 1068
-- 19 --
Fitzpatrick, E. L., & Massa, D. M. 1990, ApJS, 72, 163
Gordon, K. D., & Clayton, G. C. 1998, ApJ, 500, 816
Gordon, K. D., Clayton, G. C., Misselt, K. A., Landolt, A. U., & Wolff, M. J. 2003, ApJ,
594, 279 (G03)
Green, J. C., Snow, T. P., Jr., Cook, T. A., Cash, W. C., & Poplawski, O. 1992, ApJ, 395,
289
Gunderson, K.S., Clayton, G.C., & Green, J.C. 1998, PASP, 110, 60
Hutchings, J. B., & Giasson, J. 2001, PASP, 113, 1205
Jenkins, E. B., Savage, B. D., & Spitzer, L. 1986, ApJ, 301, 355
Jura, M. 1974, ApJ, 191, 375
Jura, M. 1975, ApJ, 197, 581
Kim, S.-H., Martin, P. G., & Hendry, P. D. 1994, ApJ, 422, 164
Kim, S.-H., & Martin, P. G. 1996, ApJ, 462, 296
Koornneef, J., & Code, A. D. 1981, ApJ, 247, 860
Koornneef, J. 1982, A&A, 107, 247
Laor, A., & Draine, B. T. 1993, ApJ, 402, 441
Lewis, N. K., Cook, T. A., & Chakrabarti, S. 2005, ApJ, 619, 367
Longo, R., Stalio, R., Polidan, R. S., & Rossi, L. 1989, ApJ, 566, 267
Massa, D., & Fitzpatrick, E. L. 2000, ApJS, 126, 517
Massa, D., Savage, B.D., & Fitzpatrick, E. L. 1983, ApJ, 266, 662
Massey, P. 2002, ApJS, 141, 81
Misselt, K. A., Clayton, G. C., & Gordon, K. D. 1999, ApJ, 515, 128
Mathis, J. S., Rumpl, W., & Nordsieck, K. H. 1977, ApJ, 217, 425
Nandy, K., Morgan, D. H., Willis, A. J., Wilson, R., Gondhalekar, P. M., & Houziaux, L.
1980, Nature, 283, 725
-- 20 --
Oestreicher, M. O., Gochermann, J., & Schmidt-Kaler, T. 1995, A&AS, 112, 495
Rachford, B. L., Snow, T. P., Tumlinson, J., Shull, J. M., Roueff, E., Andr´e, M., D´esert, J.
-M., Ferlet, R., Vidal-Madjar, A., & York, D. G. 2001, ApJ, 555, 839
Rachford, B. L., Snow, T. P., Tumlinson, J., Shull, J. M., Blair, W. P., Ferlet, R., Friedman,
S. D., Gry, C., Jenkins, E. B., Morton, D. C., Savage, B. D., Sonnentrucker, P.,
Vidal-Madjar, A., Welty, D. E., & York, D. G. 2002, ApJ, 577, 221
Rocca-Volmerange, B., Pr´evot, L., Pr´evot-Burnichon, M. L., Ferlet, R., Lequeux, J. 1981,
A&A, 99, L5
Sahnow, D. J., Moos, H. W., Ake, T. B., Andersen, J., Andersson, B.-G., Andr´e, M., Artis,
D., Berman, A. F., Blair, W. P., Brownsberger, K. R., Calvani, H. M., Chayer, P.,
Conard, S. J., Feldman, P. D., Friedman, S. D., Fullerton, A. W., Gaines, G. A.,
Gawne, W. C., Green, J. C., Gummin, M. A., Jennings, T. B., Joyce, J. B., Kaiser,
M. E., Kruk, J. W., Lindler, D. J., Massa, D., Murphy, E. M., Oegerle, W. R., Ohl,
R. G., Roberts, B. A., Romelfanger, M. L., Roth, K. C., Sankrit, R., Sembach, K.
R., Shelton, R. L., Siegmund, O. H. W., Silva, C. J., Sonneborn, G., Vaclavik, S. R.,
Weaver, H. A., Wilkinson, E. 2000a, ApJ, 538, L7
Sahnow, D. J., Gummin, M. A., Gaines, G. A., Fullerton, A. W., Kaiser, M. E., & Siegmund,
O. H. 2000b, Proc. SPIE, 4139, 149
Sasseen, T. P., Hurwitz, M., Dixon, W., V., & Airieau, S. 2002, ApJ, 566, 267
Savage, B. D., Bohlin, R. C., Drake, J. F., & Budich, W. 1977, ApJ, 216, 291
Savage, B. D., & Mathis, J. S. 1979, ARA&A, 17, 73
Schwering, P. B. W., & Israel, F. P. 1991, A&A, 246, 231
Shull, J. M., Tumlinson, J., Jenkins, E. B., Moos, H. W., Rachford, B. L., Savage, B. D.,
Sembach, K. R., Snow, T. P., Sonneborn, G., York, D. G., Blair, W. P., Green, J. C.,
Friedman, S. D., & Sahnow, D. J. 2000, ApJ, 538, L73
Snow, T. P., Jr., Allen, M. M., Polidan, R. S. 1990, ApJ, 359, L23
Snow, T. P., Jr., & York, D. G. 1975, Ap&SS, 34, 19
Sofia, U. J., Wolff, M. J., Rachford, B. L., Gordon, K. D., Clayton, G. C., Cartledge, S. I.
B., Martin, P. G., Draine, B. T., Mathis, J. S., Snow, T. P., & Whittet, D. C. B.
2005, ApJ, accepted (Paper I)
-- 21 --
Tumlinson, J., Shull, J. M., Rachford, B. L., Browning, M. K., Snow, T. P., Fullerton, A.
W., Jenkins, E. B., Savage, B. D., Crowther, P. A., Moos, H. W., Sembach, K. R.,
Sonneborn, G., & York, D. G. 2002, ApJ, 566, 857
Weingartner, J. C., & Draine, B. T. 2001, ApJ, 548, 296
Welty, D. E., Hobbs, L. M., & York, D. G. 1991, ApJS, 75, 425
Welty, D. E., Lauroesch, J. T., Blades, C., Hobbs, L. M., & York, D. G. 2001, ApJ, 554, L75
York, D. G., Drake, J. F., Jenkins, E. B., Morton, D. C., Rogerson, J. B., & Spitzer, L. 1973,
ApJ, 182, L1
Zubko, V. G., Menella, V., Colangeli, L., & Bussoletti, E. 1996, MNRAS, 282, 1321
This preprint was prepared with the AAS LATEX macros v5.2.
-- 22 --
1050
2e-13
1075
1100
1125
1150
1175
)
1
-
Å
1
-
s
2
-
m
c
g
r
e
(
x
u
l
F
1e-13
0
2e-12
1e-12
2e-13
1e-13
0
2e-12
1e-12
0
1050
1075
1100
1125
Wavelength (Å)
1150
1175
0
Fig. 1. -- Sample FUSE spectra.
Sample FUSE spectra are plotted above, including an example of poor quality data in the
upper panel (for AzV 456) and fine quality data in the lower panel (for AzV 70); these
spectra also comprise an extinction pair. Of further note, the AzV 456 continuum drops
dramatically between 1080 and 1120 A, giving rise to a peculiar kink in the extinction curve
discussed later in the text.
-- 23 --
0
2
4
6
8
10
0
2
4
6
8
10
0
2
4
6
8
10
)
V
A
(
/
)
λ
(
A
6
4
2
0
6
4
2
0
6
4
2
0
AzV 18
RV = 2.90
AzV 456
RV = 2.19
Sk -67° 2
RV = 3.75
Sk -68° 26
RV = 3.45
Sk -68° 155
RV = 2.81
Sk -68° 129
RV = 3.37
Sk -69° 228
RV = 3.54
Sk -68° 140
RV = 3.34
Sk -69° 279
RV = 3.54
6
4
2
0
6
4
2
0
6
4
2
0
2
4
6
8
10
0
2
4
6
1/λ [µm-1]
8
10
0
2
4
6
8
0
10
Fig. 2. -- Magellanic Cloud Extinction Curves.
The extinction curves for each of the reddened/unreddened star pairs in Table 1 are plotted
above; IR, optical, UV (from IUE), and FUV (from FUSE) data are all included. A CCM
curve based on the RV value derived from the IR and optical portions of each extinction
curve is also plotted using a dashed line, and the FM fit to the full curves are indicated in
each panel by solid grey lines. In each case but AzV 456 and Sk -69◦228, the FM fit for only
IR through UV data almost precisely overlaps the FM fit to the full curve. For these two
exceptions, the IR-to-UV fit is represented by a solid dark line. The vertical lines identify
central wavelengths for Lyα and Lyβ.
-- 24 --
7.6
7.8
8.0
8.2
8.4
8.6
8.8
9.0
3e-13
2e-13
1e-13
0
3e-13
2e-13
1e-13
0
7.6
7.8
8.0
8.4
8.2
1/λ [µm-1]
8.6
8.8
3e-13
2e-13
1e-13
0
3e-13
2e-13
1e-13
0
9.0
]
1
-
Å
1
-
s
2
-
m
c
g
r
e
[
x
u
l
F
Fig. 3. -- Sk -69◦228 Spectral Match.
IUE and FUSE spectra in a region that includes these instruments' overlapping wavelength
coverage have been plotted above for the star Sk -69◦228; IUE fluxes are represented by the
heavy solid line and the light grey line signifies FUSE data. The top and bottom panels
depict the flux levels before and after the linear corrections were applied. These corrections
generally reduced the size of extinction curve offsets between UV and FUV data. The vertical
dashed lines delimit portions of the spectrum contaminated by Lyα absorption that have
been eliminated from the curves plotted in Figure 2; the central portions have also been set
to zero in this plot.
-- 25 --
Current LMC/SMC Data
LMC/SMC (Tumlinson et al. 2002)
LMC (Gunderson et al. 1998)
MW (Savage et al. 1977)
0
-2
]
)
2
H
f(
[
0
1
g
o
l
-4
-6
-8
0
0.05
0.1
0.15
E(B-V)
0.2
0.25
0.3
Fig. 4. -- Galactic and Magellanic Cloud Molecular Hydrogen Fractions.
The variation of molecular hydrogen fraction f (H2) with color excess E(B −V ) is plotted
above for sight lines passing through material associated with the Magellanic Clouds and
the Milky Way. The current sample includes stars that are more heavily reddened that
the Tumlinson et al. (2002) sample and that emphasize the similarities between how f (H2)
relates to E(B −V ) in these galaxies. The label E(B −V ) used in this caption and for the
x-axis in the plot refer to unadjusted values in the Galaxy but E ′(B −V ) for Magellanic
Cloud sight lines.
-- 26 --
0
0.05
0.1
0.15
0.2
0.25
)]
V
-
B
(
'
E
/
)
I
H
N(
[
0
1
g
o
l
22.5
22.0
21.5
)]
V
-
B
(
'
E
/
)
l
a
t
o
T
H
N(
[
0
1
g
o
l
22.5
22.0
21.5
LMC
Milky Way
LMC
Milky Way
SK -67° 2
New data
Tumlinson et al (2002)
Gunderson et al (1998)
22.5
22.0
21.5
22.5
22.0
21.5
0.00
0.05
0.10
0.15
E'(B-V)
0.20
0.25
Fig. 5. -- LMC Gas-to-Dust Ratios.
Gas-to-dust ratios for LMC extinction pairs are plotted above as a function of E(B − V ).
Of particular note, Sk -67◦2 alone among the reddened LMC stars is characterized by a
log10[N(H I)/E ′(B −V )] ratio consistent with that of a Galactic star (dotted line; Diplas
& Savage 1994), whereas the corresponding value of log10[N(HTotal)/E ′(B −V )] agrees with
other LMC ratios. The LMC value for log10[N(H I)/E ′(B −V )] (dashed line; Koornneef
1982) is also plotted in both graphs for reference.
-- 27 --
2
4
6
8
10
SMC Bar
LMC2
SMC Wing
LMCAvg
10
5
0
10
5
0
10
5
0
10
5
)
V
(
λ)/A
(
A
0
0
2
4
6
1/λ [µm-1]
8
0
10
Fig. 6. -- Comparing the LMC, SMC, and Galactic Average Extinction Curves.
Average extinction curves for the Magellanic Cloud regions LMC2 and SMC bar, those more
distinct in character from Galactic extinction, are shown in the upper panel of the above
plot; the lower panel depicts the mean curves for the LMCAvg and SMC wing groupings.
Both panels include the CCM RV =3.1 curve in the role of a Galactic reference. The LMC
and SMC curves are offset 1 and 2 units in A(λ)/A(V ), respectively, from the Galactic curve
in each panel.
-- 28 --
Fig. 7. -- MEM Average Extinction Curve Fits.
The MEM fits of astrophysical silicate, amorphous carbon, and graphite dust components
to FM parameterizations of the average extinction curves in Figure 6 are plotted above.
It should be noted that the MEM models were adjusted to reproduce the FM fit to each
extinction curve; consequently, the SMC wing curve does not match the AzV 456/AzV 70
extinction plot of Figure 2 in full detail. The error bars are indicative of the deviation
between the FM fit and the extinction curve in nearby bins. For comparison, a typical
Galactic sight line is represented in this plot by HD 14250 (RV =2.98±0.14; Paper I).
-- 29 --
Fig. 8. -- MEM Dust Grain Mass Distributions.
The MEM grain size vs m(a)da functions for each of the average extinction curves of Figure 6
are plotted above. Of particular note, the SMC grain distributions do not include silicate
or amorphous carbon grains as large as are required to fit extinction curves in the LMC or
the Galaxy. Also, the graphite grain distribution for the SMC wing is strongly peaked by
the constraints implied in the bump strength and FUV rise of its extinction curve. As in
Figure 7, HD 14250 stands in for a typical Galactic sight line.
Table 1. Extinction Curve Pairs
Reddened Comparison Spectral ∆(B −V )
Star
(mag)
Star
Type
RV
AV
(mag)
N(H I)/AV
(1021 H I atoms/AV )
AzV 462
AzV 18
AzV 70
AzV 456
Sk -66◦35
Sk -67◦2
Sk -66◦35
Sk -68◦26
Sk -68◦129 Sk -68◦41
Sk -68◦140 Sk -68◦41
Sk -68◦155 Sk -67◦168
Sk -69◦228 Sk -65◦15
Sk -69◦279 Sk -65◦63
B2 Ia
O9.7 Ib
B2 Ia
B3 Ia
O9 Ia
B0 Ia
O8 Ia
B2 Ia
O9 Ia
0.17±0.03
0.26±0.03
0.15±0.05
0.19±0.03
0.17±0.05
0.20±0.05
0.20±0.05
0.15±0.05
0.21±0.05
2.90±0.42
2.19±0.23
3.75±0.36
3.45±0.27
3.37±0.29
3.34±0.25
2.81±0.22
3.54±0.34
3.54±0.25
0.49±0.11
0.57±0.08
0.56±0.23
0.64±0.16
0.57±0.22
0.67±0.22
0.56±0.18
0.53±0.23
0.75±0.23
17.27±4.30
7.01±1.22
1.78±0.67
5.46±1.08
6.98±2.36
5.99±1.75
8.88±9.21
6.60±2.50
5.38±1.50
--
3
0
--
Note. -- The Galactic foreground extinction is considered to be comparable for the stars in each
pair; consequently, the properties listed in this table are appropriate only to the Magellanic Cloud
dust component (Misselt, Clayton, & Gordon 1999). Spectral types refer to the UV classification
of each star; the sources are Gordon & Clayton (1998) and Misselt, Clayton, & Gordon (1999).
-- 31 --
Table 2. FUSE LMC and SMC Observations
Sight Line FUSE Data Set
Date
Exposure Time (s)
AzV 18
AzV 70
AzV 456
AzV 462
Sk -65◦15
Sk -65◦63
Sk -66◦35
Sk -67◦2
Sk -67◦168
Sk -68◦26
Sk -68◦129
Sk -68◦140
Sk -68◦155
Sk -69◦228
Sk -69◦279
A1180101000
B0890101000
A1180202000
A1180203000
B0900601000
Q1070101000
Q1070104000
Q1070106000
Q1070102000
Q1070103000
P2210201000
A1180301000
B0861001000
A0490701000
M1142001000
B0861101000
B1280101000
B0860301000
B0860101000
B0860901000
B0860201000
B0860501000
B0860601000
B0860701000
B0860401000
B0860801000
2000 May 29
2001 Jun 13
2000 Oct 03
2000 Oct 05
2001 Jun 15
2000 Oct 04
2000 Oct 06
2000 Oct 09
2000 Oct 10
2000 Oct 12
2001 Jun 14
2000 Jul 03
2001 Nov 17
1999 Dec 16
2000 Sep 26
2001 Sep 22
2001 Oct 25
2002 Sep 22
2001 Aug 14
2001 Sep 22
2001 Sep 17
2001 Sep 22
2001 Sep 17
2001 Sep 22
2001 Sep 23
2001 Sep 22
9293
45245
2831
1971
6152
3691
4903
5641
2687
4352
8278
5550
4293
5927
5581
4125
4030
4480
6935
4102
11391
6681
10712
8020
9422
5991
-- 32 --
Table 3. Molecular Hydrogen Models: Milky Way
log10 N (H2)MW log10 N (0)
(cm−2)
(cm−2)
log10 N (1)
(cm−2)
log10 N (2)
(cm−2)
log10 N (3)
(cm−2)
log10 N (4)
(cm−2)
log10 N (5)
(cm−2)
b
km s−1
a
T01
(K)
b
Tex
(K)
16.57(0.64)
18.55(0.05)
15.92(0.16)
18.15(0.06)
18.00(0.07)
17.78(0.46)
16.83(0.36)
16.38(0.35)
16.02(0.35)
15.31(0.27)
16.73(0.59)
19.01(0.12)
18.35(0.50)
18.00(0.20)
16.17(0.16)
16.33
18.42
15.55
17.79
17.75
17.24
16.56
15.60
15.61
14.31
15.76
18.54
18.26
17.29
15.70
16.17
17.95
15.48
17.86
17.54
17.60
16.19
16.25
15.32
15.14
16.39
18.82
17.64
17.80
15.93
SMC
LMC
14.82
15.90
14.88
16.75
16.66
16.34
15.98
15.24
15.58
14.39
16.15
17.37
15.77
17.22
14.88
14.27
15.46
14.98
16.04
16.56
15.74
15.81
14.74
14.67
14.31
15.87
16.28
15.36
16.14
14.71
14.66
14.08
15.13
14.37
14.39
8.1+2.9
−1.9
4.2+0.9
−1.6
5.9+2.3
−4.9
2.4+1.2
−1.4
1.4+1.1
−0.4
4.8+2.6
−2.5
6.3+5.9
−5.2
3.5+1.2
−2.2
5.4+6.0
−4.4
4.3+4.0
−3.3
2.9+6.0
−1.9
2.5+2.4
−1.5
4.6+2.2
−3.6
2.5+5.9
−1.5
8.3+5.9
−6.0
66±192
52±6
72±34
187±296
357±786
419±223
84±6
64±8
125±310
56±90
200±95
60±62
326±230
228±868
110±42
47±133
167±189
102±59
165±330
303±263
210±227
277±197
200±95
354±519
326±230
409±868
227±50
330±183
224±620
277±308
14.17
--
3
3
--
MC Star
AzV 18
AzV 70
AzV 462
Sk −65◦15
Sk −65◦63
Sk −66◦35
Sk −67◦2
Sk −67◦168
Sk −68◦26
Sk −68◦41
Sk −68◦129
Sk −68◦140
Sk −68◦155
Sk −69◦228
Sk −69◦279
aThe H2 kinetic temperature is assumed to be equivalent to T01, the temperature derived from N (J=0) and N (1) assuming a Boltzmann distribution
(T01 = E1−E0
k
log(N(0)/g0 )−log(N(1)/g1) ).
log(e)
bThe excitation temperature Tex reflects the sum of effects such as UV photon-pumping and excited state formation, and is identified in this table
with the slope of the log[N (J)/gJ] vs. E(J) plot for J ≥ 2.
Note. -- This table describes the Milky Way component of the molecular hydrogen models constructed for each target star from the FUSE data.
Table 4. Molecular Hydrogen Models: Magellanic Clouds
MC Star
log10 N (H2)M C
(cm−2)
log10 N (0)
(cm−2)
log10 N (1)
(cm−2)
log10 N (2)
(cm−2)
log10 N (3)
(cm−2)
log10 N (4)
(cm−2)
log10 N (5)
(cm−2)
b
km s−1
T01
(K)
Tex
(K)
AzV 18
AzV 456
AzV 462
Sk −66◦35
Sk −67◦2
Sk −68◦26
Sk −68◦129
Sk −68◦140
Sk −68◦155
Sk −69◦228
Sk −69◦279
20.36(0.07)
20.93(0.09)
17.65(0.13)
19.13(0.19)
20.95(0.08)
20.38(0.08)
20.05(0.10)
19.50(0.13)
19.99(0.24)
18.70(0.14)
20.31(0.07)
20.13
20.85
16.65
19.02
20.86
20.20
20.00
19.09
19.75
18.28
20.10
19.97
20.15
17.54
18.46
20.21
19.90
19.08
19.25
19.62
18.37
19.90
SMC
LMC
17.58
17.67
16.56
17.02
18.41
18.50
18.06
18.13
17.21
17.79
16.46
17.79
15.62
16.16
15.89
16.43
17.76
16.65
17.45
16.37
17.13
16.11
16.06
15.36
14.26
15.88
15.55
15.45
14.84
14.78
15.40
16.43
15.87
15.97
15.42
15.34
8.1+2.9
−1.9
6.0+5.9
−5.0
2.1+1.2
−1.1
4.4+3.0
−3.1
5.6+6.0
−4.6
2.3+1.0
−1.3
8.0+3.8
−4.2
5.4+2.6
−2.0
8.2+3.3
−4.9
3.1+6.0
−2.1
10.0+6.0
−6.0
66±9
45±4
· · ·
49±23
46±5
59±7
40±13
93±35
68±42
86±37
64±9
270±52
82±175
234±564
302±152
159±75
253±260
309±157
248±71
341±152
249±555
389±437
Note. -- This table details the Magellanic Cloud component of the molecular hydrogen models constructed for each target star from the FUSE
data. The temperatures T01 and Tex possess the same characteristics as those in Table 3.
--
3
4
--
-- 35 --
Table 5. Flux Shifts
Star
IUE/FUSE Flux Ratio Star
IUE/FUSE Flux Ratio
AzV 18
AzV 70
AzV 456
AzV 462
Sk −65◦15
Sk −65◦63
Sk −66◦35
Sk −67◦2
1.10
1.05
0.85
1.00
1.20
0.85
1.05
0.65
Sk −67◦168
Sk −68◦26
Sk −68◦41
Sk −68◦129
Sk −68◦140
Sk −68◦155
Sk −69◦228
Sk −69◦279
1.05
0.85
0.95
1.00
1.10
1.20
1.25
1.30
Note. -- The table ratios are derived from the calibrated fluxes recorded by
each instrument in the spectral overlap from 1150 -- 1185 A.
Table 6. FM Fit Parameters
Curve
c1
c2
c3
c4
x0
γ
AzV 18
−4.902±1.036
2.255±0.436
AzV 456
−0.419±0.130
0.908±0.094
−4.66
Sk −66◦19a
Sk −67◦2
−3.479±1.585
Sk −68◦26
0.003±0.020
Sk −68◦129 −2.174±0.784
−1.704±0.253
Average
2.02
1.723±0.648
0.937±0.150
1.393±0.443
1.268±0.041
SMC Bar
0.165±0.213
SMC Wing
5.026±1.625
LMC
1.21
3.108±1.296
2.581±0.513
1.184±0.351
2.902±0.187
LMC2
0.001±0.026
4.697±0.078
0.738±0.012
0.513±0.100
4.770±0.079
1.470±0.025
0.85
4.54
0.73
0.881±0.313
0.203±0.054
0.207±0.082
0.266±0.030
4.573±0.048
4.638±0.049
4.567±0.071
4.602±0.023
0.935±0.016
0.898±0.015
0.688±0.041
0.930±0.015
Sk −68◦140 −1.929±0.820
Sk −68◦155 −2.842±0.982
Sk −69◦228 −2.800±1.219
Sk −69◦270a
Sk −69◦279 −2.696±0.857
−2.487±0.303
Average
−3.51
1.323±0.387
1.615±0.458
1.449±0.529
0.161±0.055
1.093±0.525
0.897±0.336 −0.012±0.013
0.641±0.237 −0.418±0.164
4.440±0.074
4.611±0.062
4.714±0.079
0.811±0.014
0.684±0.028
0.609±0.054
1.52
0.97
0.24
4.62
0.78
1.364±0.364
1.443±0.054
0.853±0.232 −0.124±0.057
0.844±0.140 −0.055±0.016
4.603±0.077
4.586±0.058
0.664±0.047
0.697±0.012
aIn the interest of providing additional extinction curves for comparison, the data for sight lines previ-
ously observed by Hopkins Ultraviolet Telescope (Sk −66◦19; Sk −69◦270; Clayton et al. 1996; Gunderson,
Clayton, & Green 1998) have been included in this table.
--
3
6
--
Table 7. Properties of the Magellanic Cloud ISM
Sight Line or
Extinction Pair
N(H I)
(cm−2)
N(H2)
(cm−2)
E ′(B −V )a N(H I)/E ′(B −V ) N(H2)/E ′(B −V )
log10fH2
(mag)
(cm−2 mag−1)
(cm−2 mag−1)
AzV 18
AzV 456
AzV 462
AzV 18/AzV 462
AzV 456/AzV 070
Sk −66◦35
Sk −67◦2
Sk −68◦26
Sk −68◦129
Sk −68◦140
Sk −68◦155
Sk −69◦228
Sk −69◦279b
9.0×1021
1.5×1021
6.0×1020
8.5×1021
4.0×1021
4.0×1020
1.0×1021
3.5×1021
5.2×1021
5.5×1021
4.2×1021
4.0×1021
3.5×1021
Sk −66◦19/Sk −69◦83
Sk −67◦2/Sk −66◦35
Sk −68◦26/Sk −66◦35
Sk −68◦129/Sk −68◦41
7.0×1021
1.0×1021
3.5×1021
4.0×1021
2.3×1020 −1.35
8.5×1020 −0.28
4.5×1017 −2.83
2.3×1020 −1.29
8.5×1020 −0.53
SMC: individual targets
0.174
0.332
0.007
SMC: extinction pairs
0.17
0.26
LMC: individual targets
0.055
0.190
0.181
0.201
0.215
0.217
0.145
0.221
LMC: extinction pairs
0.25
0.18
0.19
0.17
1.3×1019 −1.20
8.9×1020 −0.19
2.4×1020 −0.92
1.1×1020 −1.38
3.0×1019 −1.97
9.8×1019 −1.35
4.9×1018 −2.61
2.0×1020 −0.98
1.6×1020 −1.64
8.9×1020 −0.19
2.3×1020 −0.94
1.1×1020 −1.27
--
3
7
--
5.2×1022
4.5×1021
8.6×1022
5.0×1022
1.5×1022
7.3×1021
5.3×1021
1.9×1022
2.6×1022
2.6×1022
1.9×1022
2.8×1022
1.6×1022
2.5×1022
5.6×1021
1.8×1022
2.3×1022
1.3×1021
2.6×1021
6.4×1019
1.4×1021
3.3×1021
2.4×1020
4.7×1021
1.3×1021
5.5×1020
1.4×1020
4.5×1020
3.4×1019
9.0×1020
6.4×1020
5.0×1021
1.2×1021
6.6×1020
Table 7 -- Continued
Sight Line or
Extinction Pair
N(H I)
(cm−2)
N(H2)
(cm−2)
E ′(B −V )a N(H I)/E ′(B −V ) N(H2)/E ′(B −V )
log10fH2
(mag)
(cm−2 mag−1)
(cm−2 mag−1)
LMCAvg curvec
Sk −68◦140/Sk −68◦41
Sk −68◦155/Sk −67◦168
Sk −69◦228/Sk −65◦15
Sk −69◦270/Sk −67◦78
Sk −69◦279/Sk −65◦63
LMC2 curvec
· · ·
4.0×1021
5.0×1021
3.5×1021
3.5×1021
4.0×1021
· · ·
· · ·
−1.00
3.0×1019 −1.83
9.8×1019 −1.42
4.9×1018 −2.56
0.7×1020 −1.69
2.0×1020 −1.04
−1.50
· · ·
· · ·
0.20
0.20
0.15
0.19
0.21
· · ·
1.1×1022
2.0×1022
2.5×1022
2.3×1022
1.8×1022
1.9×1022
1.9×1022
5.9×1020
1.5×1020
4.9×1020
3.3×1019
3.7×1020
9.7×1020
3.1×1020
aAs defined in the text, E ′(B −V ) refers to the portion of E(B −V ) for each star arising in the Magellanic
Clouds; the value listed for each extinction pair is equivalent to the quantity ∆(B −V ) in Table 1.
bThe atomic hydrogen column density for Sk −69◦279 is based on N(H I)LM C from nearby stars, since no
previously-published measurement could be found.
cGas-to-dust ratios for each of the two LMC group mean extinction curves are included. In order to derive
representative molecular gas-to-dust ratios, an f (H2)-value was assumed for each curve and compared with
the atomic gas-to-dust ratio determined from our averaging code; the results of these calculations are roughly
consistent with the sight line properties in each group. RV -values determined for the LMCAvg and LMC2 mean
--
3
8
--
curves are 3.49 ± 0.17 and 3.24 ± 0.13, respectively.
Note. -- Comparison of our H2 measurements with Tumlinson et al. (2002) required estimates of N(H I)LM C
and E′(B −V ), the Magellanic Cloud portion of the color excess; our sources are Fitzpatrick (1985a,b). The
properties for extinction pairs studied by Gunderson, Clayton, & Green (1998), Sk -66◦19 and Sk -69◦270, are
listed for comparison.
--
3
9
--
-- 40 --
Table 8. MEM Curve Abundance Requirements
Extinction
Grouping
Silicon Carbon
AS
Total AMC Graphite
SMC Wing
SMC Bar
LMCAvg
LMC2
HD 14250
170% 44% 15%
78%
19% 14%
45% 100% 77%
19%
78% 66%
114% 75% 51%
29%
5%
23%
12%
24%
Note. -- Our MEM modelling reproduced
curves corresponding to the FM parameters fit to
the average extinction curves for each sight line
grouping including wavenumbers from 0.455 to
9.575 µm−1; a typical Galactic sight line is repre-
sented here by HD 14250 (RV =2.98±0.14; Paper
I). The demands on each element are expressed as
a percentage of the total interstellar abundance
for each of the astronomical silicate (AS), amor-
phous carbon (AMC), and graphite grain compo-
nents.
|
astro-ph/9604085 | 1 | 9604 | 1996-04-16T13:27:41 | The Supernova Remnant G11.2-0.3 and its central Pulsar | [
"astro-ph"
] | The plerion inside the composite Supernova Remnant G11.2-0.3 appears to be dominated by the magnetic field to an extent unprecedented among well known cases. We discuss its evolution as determined by a central pulsar and the interaction with the surrounding thermal remnant, which in turn interacts with the ambient medium. We find that a plausible scenario exists, where all the observations can be reproduced with rather typical values for the parameters of the system; we also obtain the most likely period for the still undetected pulsar. | astro-ph | astro-ph | The Supernova Remnant G11.2 -- 0.3 and its central Pulsar
Osservatorio Astrofisico di Arcetri, Largo E. Fermi 5, 50125 Firenze, Italy
R. Bandiera
F. Pacini
Osservatorio Astrofisico di Arcetri, and
Dipartimento di Astronomia e Scienza dello Spazio, Universit`a degli Studi,
Largo E. Fermi 5, 50125 Firenze, Italy
and
M. Salvati
Osservatorio Astrofisico di Arcetri, Largo E. Fermi 5, 50125 Firenze, Italy
Received 3 February 1996;
accepted 11 April 1996
6
9
9
1
r
p
A
6
1
1
v
5
8
0
4
0
6
9
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
-- 2 --
ABSTRACT
The plerion inside the composite Supernova Remnant G11.2 -- 0.3 appears to
be dominated by the magnetic field to an extent unprecedented among well
known cases. We discuss its evolution as determined by a central pulsar and the
interaction with the surrounding thermal remnant, which in turn interacts with
the ambient medium. We find that a plausible scenario exists, where all the
observations can be reproduced with rather typical values for the parameters of
the system; we also obtain the most likely period for the still undetected pulsar.
Subject headings:
ISM: Supernova Remnants -- ISM: Individual: G11.2 -- 0.3 --
Stars: Pulsars: General
-- 3 --
1.
Introduction
Vasisht et al. (1996, hereafter VA) have recently discussed the nature of the remnant
G11.2 -- 0.3 (Historical Supernova 386 AD) and have shown that it is likely to be of a
composite nature, as already suggested on the basis of radio data by Morsi & Reich (1987).
The hard X-ray emission observed by ASCA, combined with previous Einstein data and
observations at other wavelengths, strongly suggests (despite the fact that a pulsar has not
yet been discovered) the presence of a plerion surrounded by a thermal shell. The data
across the electromagnetic spectrum can then be combined with the theory of the evolution
of pulsar-powered remnants (Pacini & Salvati 1973 (PS); Bandiera, Pacini & Salvati 1984;
Reynold & Chevalier 1984) and allow a determination of the physical parameters of the
system. The main outcome of the interpretation proposed by VA is that the plerion
should be characterized by a magnetic energy much higher than the energy channelled into
relativistic electrons, at variance with, e.g., the Crab Nebula.
In our paper we study the evolution of the plerion coupled with that of the surrounding
shell. We show that the magnetic field determination of VA is only marginally consistent,
from a dynamical point of view, if the thermal and non-thermal remnants are in pressure
equilibrium. We identify, on the other hand, a scenario that is in agreement with the
observations, and derive the initial and present parameters of the central pulsar. In
particular, we show that the apparent preponderance of the magnetic component in the
pulsar output may be a natural outcome of the evolution, and does not imply a strong
unbalance in the mechanism of production.
Following VA we assume that G11.2 -- 0.3 is the remnant of SN 386 AD and therefore has
an age t = 1610 yr = 5 × 1010 s. For its distance we use a reference value of 5 kpc and define
d5 = d/(5 kpc). The outer thermal shell has radius Rs = 3.3 d5 pc, electron temperature
Ts = 0.73 keV = 8.5 × 106 K, X-ray luminosity Ls ∼ 1036d2
5 erg s−1 (in the 0.6 -- 3.3 keV
-- 4 --
band), from which VA derive an ambient density no ∼ 1.5 d−1/2
5
cm−3, by assuming a Sedov
(1959) expansion in the interstellar medium. The central plerionic component has a radius
Rp ∼ 1 d5 pc and its spectrum is equal to
Lν ≃ 3 × 1016(cid:18)
ν
2.5 × 1018 Hz(cid:19)−0.8
d2
5 erg s−1Hz−1,
(1)
corresponding to an X-ray luminosity Lp ≃ 1 × 1035d2
5 erg s−1, in the 0.5 -- 4 keV band. eq. (1)
is consistent with the observations down to 32 GHz, but overestimates the plerion emission
at 1 GHz. Accordingly, the spectrum must have a break at a frequency νo between 1 GHz
and 32 GHz.
2. The model
The assumption of the PS model is that the central pulsar feeds both magnetic field
and relativistic particles into the expanding plerion. Taking into account the adiabatic
losses and the rate of production, the evolution of the field strength is given by
d
p
dt B2R4
6 ! = p ERp.
(2)
Here p is the fraction of the pulsar energy loss E that goes into magnetic energy; it is
assumed that p is constant. According to standard pulsar electrodynamics Ω ∝ −Ωn, with
n being the so-called braking index, and
E =
Eo
(1 + t/τo)α ,
(3)
with α = (n + 1)/(n − 1) = 2 in a pure dipole field. Measured values of n range roughly
between 1.4 for the Vela pulsar (Lyne, private communication) and 2.8 for PSR 1509-58
(Kaspi et al. 1994). The pulsar is assumed to inject also relativistic particles, whose
energies are distributed according to a power law with constant index γ, at a rate (1 − p) E.
-- 5 --
The evolution of the electron energies is determined by both adiabatic and synchrotron
losses.
An immediate implication of the theory is that the emitted spectrum of the plerion
should show a break at a frequency νb such that particles radiating at νb have a lifetime
equal to the age of the nebula. Since νb ∝ B−3t−2, if the observed break is identified with
νb one directly derives the field strength in G11.2 -- 0.3
B = 2.1 × 10−3(cid:18) νb
32 GHz(cid:19)−1/3
G,
which corresponds to a total field energy
WB = 2.1 × 1049(cid:18) νb
32 GHz(cid:19)−2/3
d3
5 erg.
The synchrotron spectrum above νb, up to a maximum frequency νmax, is given by
Lν =
2 − γ
γ − 1
(1 − p) E
2νmax (cid:18) ν
νmax(cid:19)−γ/2
.
(4)
(5)
(6)
From the observations γ = 1.6 and νmax > 2.5 × 1018 Hz; by combining eqs. (1) and (6)
we find the present-time energy loss from the pulsar:
E = 2.24 × 1035(cid:18)
νmax
2.5 × 1018 Hz(cid:19)0.2 d2
1 − p
5
erg s−1.
(7)
We describe the relative importance of the magnetic and particle components in terms
of the dimensionless quantity Q = B2R3
p/6(1 − p) Et. By substituting eq. (3) into eq. (2)
and integrating, we find
Q =
p
(1 + t)α
1 − p
t1+r
t
Z
0
xr
(1 + x)α dx;
(8)
here we put t = t/τo, and assumed a power law for the expansion, Rp ∝ tr. Note that Q is
a measure of the balance at production rather than a gauge of equipartition: in fact, when
p ≈ 0.5, Q ≈ 1 until t ≤ τo (and later on as well if α = 1 + r). The ratio of magnetic to
-- 6 --
particle energy, instead, can be arbitrarily high if the radiative lifetime of the energetically
important particles is sufficiently shorter than t. On the observational side, eqs. (5) and (7)
give
Q = 1860(cid:18) νb
32 GHz(cid:19)−2/3(cid:18)
νmax
2.5 × 1018 Hz(cid:19)−0.2
d5 :
(9)
the magnitude of Q in the present case, to be compared with Q ≈ 1 for the Crab plerion, is
a quantitative measure of the discrepancy which we want to address. For G11.2 -- 0.3, then,
either p is very close to unity (1 − p < 10−3), or t ≫ 1, (α − 1 − r) > 0. We regard the
former possibility as very unlikely, and in the following we shall limit ourselves only the
latter.
When t ≫ τo, it is appropriate to assume flux conservation for the evolution of the
magnetic field, B ∝ t−2r; furthermore, the plerion spectrum shows two breaks (PS), of
which the higher one, νb, corresponds to the radiative lifetime at the current time t, and
the lower one, νc, is the signature of the radiative lifetime at τo. If we identify the observed
break with νb, eq. (4) holds, and the pressure interior to the plerion is relatively very high.
As we show in the following, the magnetic pressure corresponding to eq. (4) exceeds the
shell pressure by a large factor; this is possible only if the plerion expansion is limited by
the inertia of some cold matter tied to the plerion itself, in which case we expect r ≥ 1.
Then r ≥ 1 and α ≤ 3, as observed in all measured cases, give p of order 0.5 only at the
expense of t ≥ 1000, and an initial magnetic energy of order t × WB > 1052d3
5 erg. We
consider very unlikely a value so much larger than the canonical supernova output, and
propose that: i) the plerion and the shell are in pressure equilibrium, so that r < 1; and
ii) the observed break is to be identified with νc, so that νb = νc × t10r−2 ≫ νo, and Q
is decreased with respect to eq. (9). However, the break at νb entails a change of slope
ǫ = (2.8r − 0.4α)/(5r − 1) (see PS); since G11.2 -- 0.3 shows a straight spectrum from the
microwaves to the X rays, we must require ǫ = 0, α = 7r.
-- 7 --
Within this framework more detailed results can be obtained by coupling the plerion
evolution to the hydrodynamics of the thermal shell. We can distinguish two different cases,
according to the distribution of the ambient density being of the interstellar type (ρo ∼
constant, henceforth IS), or circumstellar type (ρo ∼ r−2, henceforth CS). If the ambient
medium has a constant density, Rs ∝ t2/5, Rp ∝ t3/10, and, because of the condition on ǫ,
α = 2.1. If the shell expands inside the pre-supernova wind, one has Rs ∝ t2/3, Rp ∝ t4/7,
and α = 4.0.
We have computed in detail the two cases, and give the results in Table 1. In particular,
we have adjusted the unknown distance d5 so as to make the total energy of the shell
remnant equal to the canonical 1051 erg. We have chosen the value of νo within the allowed
interval so as to bring our results as close as possible to 'typical' values. And, finally, we
have neglected the work done by the plerion on the shell remnant, which is justified a
posteriori by the relatively small plerion energetics resulting from the computations. The
precise value of the ambient density in the IS and CS cases was computed so as to reproduce
the temperature and flux of the thermal X rays given in VA.
Note that the CS assumption leads to very 'palatable' estimates for all the unknowns,
with one exception: the slowing down exponent α implies a braking index n = 1.67. We
note that such value is still in the acceptable range between a purely dipolar (α = 3, n = 3)
and a purely monopolar geometry (α = ∞, n = 1), but that for most pulsars the braking
index is between 2 and 3. On the other hand, the IS hypothesis carries along some deviation
from perfect balance at production (p ∼ 0.9) and a pulsar magnetic field at the extreme of
the observed range (B∗ = 1.3 1013 G, like in PSR 1509-58). All in all, we regard the latter
interpretation as the most plausible one, also because it places G11.2 -- 0.3 exactly on the
empirical relation between Lp and E given by Seward & Wang (1988); we suggest that the
corresponding values in Table 1 be taken as a reasonable guess for further observing efforts.
-- 8 --
A major improvement in the modelling could ensue if future observations were to give
the precise frequency of the break(s) in the spectrum and the related change(s) of slope, or
the detection of the pulsar.
3. Conclusions
If G11.2 -- 03 is a composite remnant deriving from the historical explosion of SN 386
AD, the main parameters of the nebula and of the central pulsar can be estimated in
the framework of the PS model. They appear to be consistent with the properties of the
remnant and the standard views on pulsars.
The apparent preponderance of the magnetic channel in the output of the central pulsar
is the consequence of the dynamical evolution under the influence of the external shell.
This has slowed down the expansion and therefore reduced the importance of the adiabatic
losses with respect to the synchrotron losses. Since most of the energy was released by the
pulsar during a relatively short initial phase, it is natural that the subsequent evolution
has led to a large value of Q, eq. (8), even if the pulsar output was always well balanced,
p ≈ 0.5. We strongly stress this point as relevant to all sorts of non-thermal sources: only
if the generation of particles and fields keeps occurring for a time scale longer or roughly
equal to the age of the source does Q reflect the balance at production between fields and
particles. This is why Q ≈ 1 in the Crab plerion, for which t ≈ 3 × τo. In addition, and
for quite independent reasons, the Crab is not far from equipartition, since its νb is not far
from the frequency where most of the power is emitted.
This work was partly supported by a Grant of the Italian Space Agency.
-- 9 --
Table 1. Parameters of the Models
Parameter
IS Model
CS Model
νo
d
p
Bp
P
E
B∗
τo
Po
Eo
4 GHz
9.2 kpc
0.9
32 GHz
7.4 kpc
0.3
9.3 × 10−4 G
1.8 × 10−4 G
0.34 s
2.73 s
1.55 × 1037 erg s−1
1.25 × 1036 erg s−1
2.5 × 1013 G
1.6 × 1012 G
5.8 × 108 s
7.0 × 109 s
0.029 s
0.139 s
2.4 × 1049 erg
1.0 × 1048 erg
-- 10 --
REFERENCES
Bandiera, R., Pacini, F., & Salvati, M. 1984, ApJ, 285, 134
Kaspi, V. M., Manchester, R. N., Siegman, B., Johnston, S., & Lyne, A. G. 1994, ApJ, 422,
L83
Morsi, H. W., & Reich, W. 1987, A&AS, 71, 189
Pacini, F., & Salvati, M. 1973, ApJ, 186, 249 (PS)
Reynolds, S. P., & Chevalier, R.A. 1984, ApJ, 278, 630
Sedov, L. I. 1959, Similarity and Dimensional Methods in Mechanics (New York: Academic)
Seward, F. D., & Wang, Z. -- R. 1988, ApJ, 332, 199
Vasisht, G., Aoki, T., Dotani, T., Kulkarni, S. R., & Nagase, F. 1996, ApJ, 456, L59
This manuscript was prepared with the AAS LATEX macros v4.0.
|
0704.2179 | 1 | 0704 | 2007-04-17T19:26:29 | Spitzer spectral line mapping of supernova remnants: I. Basic data and principal component analysis | [
"astro-ph"
] | We report the results of spectroscopic mapping observations carried out toward small (1 x 1 arcmin) regions within the supernova remnants W44, W28, IC443, and 3C391 using the Infrared Spectrograph of the Spitzer Space Telescope. These observations, covering the 5.2 - 37 micron spectral region, have led to the detection of a total of 15 fine structure transitions of Ne+, Ne++, Si+, P+, S, S++, Cl+, Fe+, and Fe++; the S(0) - S(7) pure rotational lines of molecular hydrogen; and the R(3) and R(4) transitions of hydrogen deuteride. In addition to these 25 spectral lines, the 6.2, 7.7, 8.6, 11.3 and 12.6 micron PAH emission bands were also observed. Most of the detected line transitions have proven strong enough to map in several sources, providing a comprehensive picture of the relative distribution of the various line emissions observable in the Spitzer/IRS bandpass. A principal component analysis of the spectral line maps reveals that the observed emission lines fall into five distinct groups, each of which may exhibit a distinct spatial distribution: (1) lines of S and H2 (J > 2); (2) the H2 S(0) line; (3) lines of ions with appearance potentials less than 13.6 eV; (4) lines of ions with appearance potentials greater than 13.6 eV, not including S++; (5) lines of S++. Lines of group (1) likely originate in molecular material subject to a slow, nondissociative shock that is driven by the overpressure within the supernova remnant, and lines in groups (3) - (5) are associated primarily with dissociative shock fronts with a range of (larger) shock velocities. The H2 S(0) line shows a low-density diffuse emission component, and - in some sources - a shock-excited component. | astro-ph | astro-ph | Spitzer spectral line mapping of supernova remnants:
I. Basic data and principal component analysis
David A. Neufeld1, David J. Hollenbach2, Michael J. Kaufman3, Ronald L. Snell4,
Gary J. Melnick5, Edwin A. Bergin6, and Paule Sonnentrucker1
ABSTRACT
We report the results of spectroscopic mapping observations carried out to-
ward small (1′ × 1′) regions within the supernova remnants W44, W28, IC443,
and 3C391 using the Infrared Spectrograph of the Spitzer Space Telescope. These
observations, covering the 5.2 − 37 µm spectral region, have led to the detection
of a total of 15 fine structure transitions of Ne+, Ne++, Si+, P+, S, S++, Cl+,
Fe+, and Fe++; the S(0) – S(7) pure rotational lines of molecular hydrogen; and
the R(3) and R(4) transitions of hydrogen deuteride.
In addition to these 25
spectral lines, the 6.2, 7.7, 8.6, 11.3 and 12.6 µm PAH emission bands were also
observed. Most of the detected line transitions have proven strong enough to map
in several sources, providing a comprehensive picture of the relative distribution
of the various line emissions observable in the Spitzer/IRS bandpass. A principal
component analysis of the spectral line maps reveals that the observed emission
lines fall into five distinct groups, each of which may exhibit a distinct spatial
distribution: (1) lines of S and H2 (J > 2); (2) the H2 S(0) line; (3) lines of ions
with appearance potentials less than 13.6 eV; (4) lines of ions with appearance
potentials greater than 13.6 eV, not including S++; (5) lines of S++. Lines of
group (1) likely originate in molecular material subject to a slow, nondissociative
shock that is driven by the overpressure within the supernova remnant, and lines
in groups (3) – (5) are associated primarily with dissociative shock fronts with a
range of (larger) shock velocities. The H2 S(0) line shows a low-density diffuse
emission component, and – in some sources – a shock-excited component.
7
0
0
2
r
p
A
7
1
]
h
p
-
o
r
t
s
a
[
1
v
9
7
1
2
.
4
0
7
0
:
v
i
X
r
a
1Department of Physics and Astronomy, Johns Hopkins University, 3400 North Charles Street, Baltimore,
MD 21218
2NASA Ames Research Center, Moffett Field, CA 94035
3Department of Physics, San Jose State University, 1 Washington Square, San Jose, CA 95192
4Department of Astronomy, University of Massachusetts, 710 North Pleasant Street, Amherst, MA 01003
5Harvard-Smithsonian Center for Astrophysics, 60 Garden Street, Cambridge, MA 02138
6Department of Astronomy, University of Michigan, 825 Dennison Building, Ann Arbor, MI 48109
– 2 –
Subject headings: ISM: Molecules - ISM: Abundances - ISM: Clouds – molec-
ular processes – shock waves
1.
Introduction
Supernovae have long been recognized as a critical source of energy input to the ISM (e.g.
Cox & Smith 1974; McKee & Ostriker 1977). They send shock waves propagating through
the ISM, creating large cavities filled with hot ionized material. Eventually, supernova-driven
shock waves become radiative, emitting strong line emissions – primarily at optical and ul-
traviolet wavelengths – that have been widely observed from supernova remnants (e.g. Weiler
& Sramek 1988). Where supernova remnants encounter molecular clouds, they drive slower
shock waves that compress and heat the molecular material and result in strong infrared line
emission. Depending upon the shock velocity, such shock waves may be either dissociative –
resulting in the destruction of molecules by collisional processes – or nondissociative, in that
molecules are heated but nevertheless survive the passage of the shock front. The theory
of interstellar shock waves, both in atomic and molecular media, has been the subject of
extensive investigation (reviewed, for example, by Draine & McKee 1993).
Using the Infrared Spectrograph (IRS) of the Spitzer Space Telescope, we have carried
out infrared spectroscopic observations of four regions in which a supernova remnant interacts
with a molecular cloud: W44, W28, 3C391, and IC443. The IRS covers the 5.2 – 37 µm
region with unprecedented sensitivity, providing access to the lowest eight pure rotational
transitions of molecular hydrogen together with fine structure lines of [SI] and a variety of
In the observations reported here, IRS spectroscopy indicated the presence
atomic ions.
of interstellar gas over a huge temperature range, from <
∼ 100 K (H2 S(0) 28 µm line) to
∼ 105 K ([NeIII] 15.6 and 36 µm lines).
>
In this paper, we present a general overview of the data set we have obtained toward
these four supernova remnants. In §2, below, we give a brief overview of the four supernova
remnants selected for this study. In §3, we discuss the observations and methods used to
reduce the observational data, and we present spectral line maps for a representative sample
of the transitions that we detected. In §4, we describe how Principal Component Analysis can
be used to identify distinct groups of transitions that share a common origin in a particular
component of the interstellar gas. In §5, we discuss the similarities and differences in the
spatial distributions of the various spectral lines we observed. More detailed analyses of the
different shocked gas components will follow in future papers.
– 3 –
2. Source selection
The four supernova remnants we have observed – W44, W28, 3C391 and IC443 – repre-
sent classic examples of the interaction between supernova-driven shock waves and molecular
clouds. Brief descriptions of each source are given below, and finder charts appear in Figures
1 – 4. In each source, we have centered the mapped region at a position that is known from
previous observations to be a luminous source of molecular emissions; these are all positions
that have been observed previously by the Short- and/or Long Wavelength Spectrometer of
the Infrared Space Observatory (ISO). All four sources lie close to the Galactic plane, with the
mapped regions lying at Galactic latitudes b ∼ −0.43◦, −0.30◦, 0.03◦ and 3.02◦ respectively
for W44, W28, 3C391 and IC443 (see Table 1).
2.1. W44
W44 lies at a distance of 2.5 kpc and has a radius of 11 – 13 pc (Cox et al. 1999). The
remnant is associated with the pulsar B1853+01, which is believed to be the progenitor star
and provides an age of 20,000 years (Wolszczan et al. 1991; based upon the ratio of the pulsar
period to the period derivative). W44 is a prototype for a new class of supernova remnants
called "mixed-morphology" SNRs, because the radio continuum images exhibit a shell-like
morphology while the center is filled in with X-ray emission (Jones et al. 1993; Rho & Petre
1998). Such morphologies have been attributed to the interaction between the supernova
shock front and a molecular cloud (Chevalier 1999; Yusef-Zadeh et al. 2003), and W44 is
indeed associated with a Giant Molecular Cloud, recently mapped in CO emission (Reach,
Rho, & Jarrett 2005: hereafter RRJ05).
RRJ05 and Cox et al. (1999) provide nice summaries of the available observations. In
general, there is strong evidence for supernova interaction with a dense molecular cloud
through the detection of OH 1720 MHz maser emission, broad (FWHM ∼ 20 – 30 km/s)
CO molecular line emission, vibrational and rotational lines of molecular hydrogen, and
atomic fine structure lines (Claussen et al. 1997; Seta et al. 1998; Reach & Rho 2000;
RRJ05). Comparison to shock models are consistent with a mixture of dissociative and
non-dissociative shocks impacting on gas with densities of 100 − 104 cm−3 (Reach & Rho
2000; RRJ05), although Cox et al. (1999) present a model that matches most data but
requires no dense cloud interaction. The region we have observed with Spitzer (Figure 1)
lies at the northeast edge of the W44 supernova remnant, coincident with OH maser Region
E (Claussen et al. 1997). Observations of CO and X-rays imply total hydrogen column
densities of 2 × 1022 cm−2 along the sight-line (Rho et al. 1994; RRJ05), corresponding to a
visual extinction AV ∼ 10 mag.
– 4 –
2.2. W28
The supernova remnant W28 is located in a complex region of the Galactic disk where a
number of nearby large H II regions (M8 and M20) and young clusters (e.g., NGC 6530) are
present (Goudis 1976). Kinematic considerations based on molecular line broadening place
the remnant at a distance of 1.9 ± 0.3 kpc (Vel´azquez et al. 2002). The distribution of radio
molecular emission lines indicates that the remnant has an irregular shell-like structure with
prominent peak emissions in the northern and northeast sides of the shell where an interaction
between the remnant and a nearby molecular cloud was suggested (Wootten 1981). Further
evidence for such an interaction arises from the detection of numerous discrete OH 1720
MHz maser spots along the northeast ridge where the remnant is believed to be interacting
with a molecular cloud (Claussen et al. 1997). This maser activity has been shown to
constitute a tracer of supernova remnant-molecular cloud interaction (Yusef-Zadeh et al.
2003). Detection of the H2 S(3) and S(9) lines as well as a number of ionic species with
ISO suggests the presence of shocks in the northern part of the remnant (Reach & Rho
2000) as well. In the optical, short emission filaments are observed superposed over a more
diffuse and centrally-peaked emission seen throughout most of the interior of W28 (van den
Bergh 1973; Long et al. 1991). Diffuse thermal X-ray emission is also seen throughout
the region delineated by the radio shells and shows a rather patchy but centrally-peaked
emission distribution (Dubner et al. 2000; Rho & Borkowski 2002). Like W44, W28 shows a
"mixed-morphology" that combines a shell-like radio structure with a centrally-filled diffuse
X-ray emission structure. Observations of optical [SII] line ratios (Long et al. 1991) suggest
a reddening of E(B − V ) ∼ 1 − 1.3 mag, corresponding to a visual extinction of 3 − 4 mag
for an AV /E(B − V ) ratio of 3.1.
2.3.
3C391
3C391 (G31.9+0.0) is an X-ray- and radio-bright supernova remnant that lies in the
Galactic plane. The X-ray emission is diffuse (i.e., no limb brightening) and centered within
the radio shell (Rho & Petre 1996). 12CO J = 1−0 maps toward 3C391 reveal the presence
of moderately dense (n(H2) ∼ 300 – 1000 cm−3) molecular gas along the northwest portion
of the remnant (Wilner, Reynolds & Moffett 1998). The radio continuum emission from the
remnant (Reynolds & Moffett 1993), shown in Fig. 3 with the region mapped by Spitzer
superposed, suggests evolution along a strong density gradient, with compression of the
supernova blast wave to the northwest in the direction of the strongest CO emission and
lower continuum surface brightness emission to the southeast away from the molecular gas. It
is likely that the progenitor star exploded within a dense molecular cloud and the expanding
– 5 –
gas has now broken out of the cloud to the southeast.
Evidence exists for both dissociative and nondissociative shocks toward 3C391 (Reach
et al. 2002). Narrowband filter images in the [Fe II] 1.64 µm line show a pronounced intensity
peak along the bright radio bar (18h 49m 16s, –0o 55′ 00′′) at the interface between the su-
pernova remnant and the molecular cloud. Similarly, the ISOCAM CVF 5 – 18 µm spectrum
of the same region is dominated by emission lines from atomic ions – such as [Fe II] 5.5 µm,
[Ar II] 6.9 µm, [Ne II] 12.8 µm, and [Ne III] 15.5 µm – while the mid-infrared H2 rotational
lines in the same spectrum are either weak or absent (Reach et al. 2002). Approximately
2.8′ to the south and east of this position, near α =18h 49m 22s, δ = −0o 57′ 22′′ (J2000),
continuum-subtracted H2 2.12 µm images show an ∼ 30′′ diameter area containing a cluster
of small (< 5′′) clumps of H2 emission. The ISOCAM CVF spectrum of this region shows
strong H2 S(2) through S(7) H2 line emission (Reach et al. 2002). Longer wavelength ISO
LWS spectra toward this location show thermal H2O, OH, high-J 12CO (Reach & Rho 1998)
and [O I] 63 µm (Reach & Rho 1996) emission. This so-called broad molecular line region
(see Reach & Rho 1999) is also the site of OH 1720 MHz maser (Frail et al. 1996) and broad
(FWHM ∼ 20 km s−1) 12CO (J = 2 − 1), CS (J = 3 − 2), CS (J = 5 − 4), and HCO+
(J = 1 − 0) emission (Reach & Rho 1999). The existence of strong and broad molecular
line emission has led to the conjecture that the supernova blast wave is encountering dense
(n(H2) ∼ 104 – 105 cm−3) cores present within the molecular cloud from which the progenitor
of 3C391 formed. The region that we have mapped with Spitzer (Figure 3) is centered on
the broad molecular line region.
The extinction to 3C391 is very high due to its location in the Galactic plane and its
9 kpc distance (Caswell et al. 1971; Radhakrishnan et al. 1972). Column densities inferred
from the spectral analysis of ROSAT data yield a foreground column density of 2 – 3.6 ×
1022 cm−2 (Rho & Petre 1996). Since this analysis has been applied by averaging spectra
obtained over relatively large areas of the remnant, it is likely that the higher value applies
to the denser broad molecular line region of interest here. Thus, assuming a foreground
column density of 3.6 × 1022 cm−2, a line-of-sight visual extinction of Av = 19 magnitudes
is derived (Reach et al. 2002).
2.4.
IC443
IC443 is a well-studied SNR located in the outer Galaxy at a distance of approximately
1.5 kpc (Georgelin 1975; Fesen & Kirshner 1980; Fesen 1984). It is one of the more striking
examples of the interaction of a SNR with the interstellar molecular medium. IC443C, the
position that we have targeted with Spitzer, is one of several shocked gas regions within the
– 6 –
SNR first identified by DeNoyer (1978) based on the presence of high velocity HI emission.
These shocked gas regions were later detected to have broad OH absorption and CO emission
lines (DeNoyer 1979a,b). More extensive observations of CO and HCO+ emission in IC443
by Dickman et al. (1992) revealed many more regions of broad molecular line emission and
showed that they formed an expanding, tilted ring. This ring of shocked gas presumably
delineates where the SNR shock has encountered the molecular interstellar material. This
ring is also seen in 2 µm H2 emission (Burton et al. 1988; Richter et al. 1995b).
Based upon the kinematics of the observed CO and HCO+ emissions, IC443C is believed
to be located on the front side of the expanding ring (Dickman et al. 1992). The unshocked
molecular gas in this direction has a vLSR of –3 km s−1. The broad velocity emission from
the shocked molecular gas extends from the velocity of the unshocked gas to a vLSR of –70
km s−1 (Snell et al. 2005). The region of most intense molecular broad-line emission is about
1 arcminute in size.
Emission from the shocked interstellar gas in IC443 has been extensively studied and
detected in numerous atomic and molecular tracers, including HI (Braun & Strom 1986)
vibrationally excited H2 (Burton et al. 1988), the 63 µm line of [OI] (Burton et al. 1990), the
pure rotational lines of H2 (Richter et al. 1995a; Cesarsky et al. 1999) and numerous molecular
species (DeNoyer & Frerking 1981; Ziurys et al. 1989; van Dishoeck et al. 1993). Most
recently, the detection of water was reported in several of these shocked regions, including
IC443C (Snell et al. 2005). Most studies have concluded that a combination of shocks are
needed to explain the emission seen in these atomic and molecular tracers. Snell et al. (2005)
suggested the presence in IC443C of both a fast J-shock and either a slow C-shock or slow
J-shock. Given the observed ratio of H2 vibrational lines, Richter et al. (1995b) derived 2.1
µm extinctions of 1.3 and 1.6 mag for two slit positions in IC443C, corresponding to visual
extinctions AV of 12 and 15 mag.
3. Observations and results
W44, W28, 3C391, and IC443C were each observed using the IRS as part of a Cycle
2 General Observer Program, the Short-Low (SL), Short-High (SH) and Long-High (LH)
modules being used to obtain complete spectral coverage over the wavelength range available
to IRS. The spectral resolving power of IRS, R = λ/∆λ, ranges from 60 to 127 for the SL
module and is ∼ 600 for the SH and LH modules. Spectral line maps of size ∼ 1′ × 1′ were
obtained by stepping the slit perpendicular to its long axis by one-half its width, and – in
the case of SH and LH – parallel to its long axis by 4/5 of its length. Table 1 summarizes
the observational details for each source; note that the map center positions are not in all
– 7 –
cases identical to the (0,0) positions defined in the captions to Figures 10 – 13.
The data were processed at the Spitzer Science Center (SSC), using version 13.2 of the
processing pipeline, and then reduced further using the SMART software package (Higdon
et al. 2004), supplemented by additional routines (Neufeld et al. 2006a), that provide for the
removal of bad pixels in the LH and SH data, the calibration of fluxes obtained for extended
sources, the extraction of individual spectra at each position sampled by the IRS slit, and the
creation of spectral line maps from the set of extracted spectra. No background subtraction
could be performed, since no off-source measurements were made; while background sub-
traction might have affected the measured continuum fluxes, it would be expected to have
no effect upon the derived intensities for spectral lines. There is no evidence for fringing in
any of the spectra.
For several transitions observed in the LH and SH modules, the resultant line maps
exhibited striping perpendicular to the long axis of the slit. We determined this effect to
be an artifact created by the standard algorithm used in the IRSCLEAN software to obtain
interpolated data for bad pixels; that algorithm interpolates in the dispersion direction,
causing a systematic underestimate in the intensity when a bad pixel falls close to the
central wavelength of a spectral line. The striping was successfully removed by applying a
correction factor, computed separately for each position along the slit and for each affected
spectral line. The required correction factors were obtained by considering – separately for
each of the five positions along the slit – the cumulative distribution functions for the line
fluxes obtained over the entire map. In other words, for each five possible positions along
the slit, we plotted the fraction of line flux measurements that were smaller than a given
value, as a function of that value. Making the assumption that the cumulative distribution
functions should be similar except for the highest flux values, we computed the necessary
correction factors needed to bring the five distribution functions into agreement.
In order to improve the signal-to-noise ratio beyond that obtained at a single position,
we have computed average spectra for each of the sources we observed. The resultant spec-
tra, plotted in Figures 5 – 8, are the averages of all spectra within a circular region, the
contributing spectra being weighted by a Gaussian taper (HPBW of 25′′) from the center of
each synthesized beam. The central positions are given in the figure captions.
The spectra plotted in Figures 5 – 8 reveal a total of 25 spectral lines, detectable in
one or more source. These include 15 fine structure transitions of the species Ne+, Ne++ (2
transitions), Si+, P+, S, S++ (2 transitions), Cl+, Fe+ (5 transitions), and Fe++; 8 transitions
of H2 (the S(0) through S(7) pure rotational transitions), and the R(3) and R(4) transitions
of HD (detected toward IC443C and reported previously by Neufeld et al. 2006b). Beam
averaged intensities are listed in Table 2. In addition to these emission lines, the 6.2, 7.7,
– 8 –
8.6, 11.3 and 12.6 µm PAH emission bands are readily apparent in the spectra of W44, W28,
and 3C391, and the 11.3 µm micron feature is detected in IC443C. A discussion of the PAH
emission features is deferred to a future paper.
The observed line transitions represent a wide variety of excitation conditions. In Figure
9, we show their position in a scatter plot, where the horizontal axis shows the appearance
potential – equal to zero for neutral species and equal to the ionization potential of X(n−1)+
for the case of ion Xn+ – and the vertical axis shows the critical density at which the rate
of collisional de-excitation equals the spontaneous radiative rate. In computing the critical
density, we assumed the dominant collision partner to be electrons for ionized species and
hydrogen molecules for neutral species.
In Figures 10 – 13, again for each source, we present maps of a representative selection
of six strong spectral lines: the H2 S(0), [SI] 25 µm, H2 S(2), [SiII] 35 µm, [NeII] 12.8 µm,
and [SIII] 33 µm lines. As demonstrated by the Principal Component Analysis described
in §4 below, each of the other transitions shows a spatial variation that is almost identical
to one of the transitions plotted in Figures 10 – 13. For example, except for a constant
scaling factor, the [FeII] 26 µm line maps are visually indistinguishable from those shown
here for [SiII] 35 µm. For the maps obtained with the SL module, the effective integration
time and resultant signal-to-noise ratio is the same everywhere in the map. For the SH and
LH modules, the effective integration time is constant, except in overlap regions between
adjacent pointings where the integration time is larger by a factor two.
4. Principal Component Analysis
A cursory inspection of Figures 10 – 13 reveals both similarities and differences in the
spatial distributions of the various emission lines that we have observed. The method of Prin-
cipal Component Analysis (PCA) provides a valuable tool for describing those similarities
and differences. The value of PCA in analysing multitransition maps was first demonstrated
by Ungerechts et al. (1997), who applied the method to molecular line maps of the Orion
Molecular Cloud and gave a comprehensive description of PCA that will not be repeated
here (see also Heyer & Schloerb 1997). To summarize, PCA creates an orthogonal basis
set of maps – known as principal components (PC) – such that each observed map can be
represented as a linear combination of the PC. The key feature of the new basis set is that
the PC are obtained in decreasing order of amplitude: the first PC contains as much of the
spatial variation as possible, the second PC - constructed orthogonal to the first - contains
as much of the remaining variation as possible, and so forth.
– 9 –
Following Ungerechts et al. (1997), we applied the Principal Component Analysis to
spectral line maps that had been mean-subtracted and divided by the standard deviation,
the latter being dominated by structure in the source rather than instrumental noise. This
procedure gives equal weighting to each line instead of allowing the brightest lines to domi-
nate the analysis. Because this analysis involves data obtained with all three modules (i.e.
SL, SH, and LH), it was applied only to the intersection of the regions mapped by the dif-
ferent modules. Applying PCA to the four sources we have observed, we find that the first
PC contains 58 − 73% of the total spatial variation in the maps, the first and second PCs
contain together 82 − 88%, and the first three PCs contain 90 − 93% of the information. In
Figures 14 – 17 we present the results of the PCA for each source. The upper panel shows
the fraction of the variance accounted for by each component (asterisks), along with the
cumulative fraction for the first n components as a function of n (open squares). The lower
panels show the coefficients for the 1st, 2nd and 3rd PC needed to approximate the maps of
each transition; the coefficients for the 1st and 2nd PC are shown in the left lower panel and
those for the 2nd and 3rd PC are shown in the right lower panel. (These figures were referred
to as h-plots by Ungerechts et al. 1997.) Because the PCA is applied to mean-subtracted
line maps, the coefficients may be either positive or negative. While the 3rd PC contains
information about the source – as evidenced by the fact that transitions with similar coeffi-
cients for the 2nd PC tend to show similar coefficients for the 3rd – the 4th and higher PCs
are likely dominated by noise.
The principal components are normalized such that quadrature sum of the coefficients
for each transition is unity. Thus, if a particular transition were completely accounted for by
the 1st and 2nd PC, its position in the lower left plot would lie on the circle of unit radius;
similarly, if a particular transition were completely accounted for by the 2nd and 3rd PC,
its position in the lower right plot would lie on the circle of unit radius. The proximity of
most points in the lower left plot to the circle of unit radius is a reflection of the fact that
the first two PCs account for ∼ 85% of the variance in the observed line maps. In the limit
where two points actually lie on the unit circle, the cosine of the angle between the two radii
measures the degree of linear correlation; points lying next to each other on the unit circle
would be perfectly correlated (cos θ = +1), while those lying exactly on opposite sides of the
circle would be perfectly anticorrelated (cos θ = −1). For example, in W44, the lower left
panel indicates that the [FeII] 26 µm and [SiII] transitions are very well correlated with each
other, strongly anticorrelated with both observed [SIII] transitions, and almost uncorrelated
with the H2 S(0) transition.
In W44, the PCA yields a clear separation into five groups of spectral lines: (1) lines of
S, and H2 (J > 2); (2) the H2 S(0) line; (3) lines of ions with appearance potentials less than
13.6 eV (Fe+, Si+, P+); (4) lines of ions with appearance potentials greater than 13.6 eV
– 10 –
(Ne+, Ne++, Fe++), not including S++; (5) lines of S++. In the other sources, some of these
groups share a common distribution, but in no case do the spectral lines within a group
defined above fail to be strongly correlated with each other. In W28, lines in groups (1) and
(2) share a similar distribution, which is markedly different from that of groups (3) – (5). In
3C391, lines in groups (1), (2), (3) and (5) show distributions that are readily distinguishable
from each other, but group (4) shows a distribution that is very similar to that of group (3).
In IC443, groups (1), (3), and (4) are readily distiguishable from each other; group (2) (the
H2 S(0) line) has a similar distribution to that of group (1); and group (5) ([SIII] transitions)
is not detected. These behaviors suggest that the line emissions we have observed originate
in at least five physically distinct components (some of which have similar distributions – as
projected onto the sky – in sources other than W44). The implications of these groupings
are discussed further in §5 below.
5. Discussion
5.1. Line emission from neutral species: H2 and atomic sulphur
Each of the sources we observed shows strong emission in the H2 S(0) through S(7) lines.
In Figure 18, we present rotational diagrams derived from the H2 line fluxes listed in Table
1. Here we corrected for extinction, adopting the foreground extinction estimates given
in §2 and the interstellar extinction curves of Weingartner & Draine (2001; "Milky Way,
RV = 3.1"). The rotational diagrams bear a strong similarity to those obtained previously
from Spitzer observations of the Herbig-Haro objects HH54 and HH7 (Neufeld et al. 2006a).
In each case, a positive curvature in the rotational diagrams indicates the presence of an
admixture of gas temperatures, while a characteristic zigzag pattern – more pronounced for
low-lying transitions – indicates a nonequilibrium ortho-to-para ratio (OPR), smaller1 than
3. In Neufeld et al. (2006a), we argued that the nonequilibrium H2 ortho-to-para ratios were
consistent with shock models in which the gas is warm for a time period shorter than that
required for reactive collisions between H and para-H2 to establish an equilibrium OPR.
As in Neufeld et al. (2006a), we obtained a two-component fit to the rotational diagrams,
in which a warm component with temperature, Tw, ortho-to-para ratio, OPRw, and column
density Nw coexists with a hot component with temperature, Th, ortho-to-para ratio, OPRh,
and column density Nh. The resultant parameters, listed in Table 3, are broadly similar to
1Except in IC443C, where – given the uncertainties – the data are consistent with an equilibrium OPR
of 3.
– 11 –
those obtained for HH7 and HH54, and are consistent with the presence of nondissociative
shocks2 of velocity 10 – 20 km s−1. Our two-component fits are based upon a simple LTE (lo-
cal thermodynamic equilibrium) treatment of the H2 excitation, a treatment that is accurate
for the lower-lying transitions but may be an oversimplification in the case of the highest
transitions (S(6) and S(7)). In a future paper, we will present a refined analysis, including
non-LTE effects, together with maps of the various physical parameters derived from the H2
(and, for IC443C, the HD) maps.
While the H2 S(1) through S(7) lines show a clear morphological similarity to each
other, a result that is apparent from inspection of the spectral line maps and is confirmed
by the PCA, the H2 S(0) line is typically more extended than – and, in W44 and 3C391,
essentially uncorrelated with – the other H2 rotational lines. The [SI] 25 µm line, by contrast,
is always well correlated with the H2 S(1) through S(7) lines, despite the fact that its upper
state energy is very similar to that of H2 S(0). These behaviors are demonstrated clearly
in Figure 19, which present scatter diagrams showing correlations between [SI] and H2 S(0).
Each point in Figure 19 shows the mean intensities within a 5′′ × 5′′ square region. The
results shown for W28 are very revealing: while [SI] and H2 S(0) are well correlated in W28
(with a linear correlation coefficient of 0.88), the best-fit linear regression yields a positive
x-intercept, indicating that there is an extended component of H2 S(0) emission (with typical
intensity ∼ 2.5 × 10−5erg cm−2 s−1 sr−1) that is unassociated with measurable [SI] emission.
Figure 19 suggests an explanation for why [SI] 25 µm and H2 S(0) are reasonably well-
correlated in IC433C and W28 but not in W44 and 3C391: in W28 and IC443C, the shock
excited contribution dominates the variations in the extended component, whereas in W44
and 3C391 it does not.
The much lower [SI]/H2 S(0) line ratio in the extended component could indicate a much
lower abundance of atomic sulfur, or - more likely - a lower gas density, the critical density for
the [SI] 25 µm line being much greater than that of the H2 S(0) line (see Fig. 9). The extended
H2 S(0) emission in these sources has an intensity comparable to that measured by Falgarone
et al. (2005), who observed H2 pure rotational emission along a long sight-line within the
Galaxy that did not intersect any region of high-mass star formation. Falgarone et al. argued
that the observed H2 emission originated in a warm gas component (unassociated with
supernova remnants or regions of star formation) that is present within the cold, quiescent
2Given the magnetic field strengths typical of the interstellar medium, such shocks are expected to be of
"continuous" or "C-type" in the designation of Draine (1980), and are qualitatively different from the "jump"
or "J-type" shock fronts associated with faster, dissociative shock waves. C-type shocks are characterized by
continuous variations in the gas velocity, large ion-neutral drift velocities, and lower peak gas temperatures
than those attained behind J-type shocks of the same velocity.
– 12 –
ISM. Alternatively, or in addition, the H2 S(0) emission within supernova remnants might
originate within low density photodissociation regions that are irradiated and heated by
UV radiation from nearby fast shocks.
In contrast to the continuum radiation incident
upon typical PDRs, the UV radiation emitted by fast shocks is dominated by spectral lines.
Although detailed calculations will be needed to model PDRs that are irradiated by UV
emission from fast shocks, we speculate that a line-dominated spectrum may be particularly
effective in exciting H2 rotational emissions since – except for spectral lines that happen
to coincide with an H2 Lyman or Werner band transition – line radiation can heat the gas
without photodissociating molecules.
It is also noteworthy that the [SI] emission is better correlated with the warm H2 than
with the ionized species in these sources. Haas et al. (1991) had previously suggested that
[SI] was a tracer of fast dissociative shocks, rather than the slower nondissociative shocks
responsible for the H2 emissions. At least in the supernova remnants that we have mapped
with Spitzer, the PCA strongly suggests that [SI] traces slow nondissociative shocks.
The atomic sulfur abundance can be estimated from the [SI] 25 µm line flux. Unfortu-
nately, the critical density for the [SI] 25 µm line is probably large in neutral gas (∼ 106 cm−3
according to Hollenbach & McKee 1989, Table 5), so the derived SI abundance depends on
both the gas density and the adopted rate coefficient for excitation of SI in collisions with
H2. IC443C presents the best case for determining the atomic sulfur abundance, because
a measurement of the HD R(4)/R(3) line allows the gas pressure to be determined for this
case (Neufeld et al. 2006b). Adopting the H2 column densities and temperatures obtained
from a two-component fit to the rotational diagram, assuming the warm and hot gas compo-
nents to be at a common pressure adjusted to match the HD R(4)/R(3) line ratio (Neufeld
et al. 2006b), and making use of Hollenbach & McKee's (1989; hereafter HM89) estimate
of the rate coefficient for excitation of [SI] fine structure collisions3, we derive an atomic
sulfur abundance of ∼ 10−6 relative to H2. This value corresponds to ∼ 4% of the solar
abundance of sulfur (1.4 × 10−5 relative to H nuclei; Asplund, Grevesse, & Sauval 2005).
The uncertainty in this estimate is large, not least because of the absence of a reliable esti-
mate for the excitation rate coefficient. Owing to the presence of a significant UV radiation
field produced by fast dissociative shocks in the vicinity of the warm molecular component,
Snell et al. (2005) argued that sulfur could be significantly ionized within the nondissociative
molecular shocks that are responsible for the pure rotational emissions from H2; thus the
observed SI abundance might be entirely consistent with a negligible depletion of sulfur onto
grains and a negligible fraction of the sulfur nuclei being bound within gas-phase molecules.
3This estimate is actually for excitation by atomic hydrogen, but – since we are not aware of any estimate
in the literature for collisional excitation of SI by H2 – we adopt it for excitation by H2 as well.
– 13 –
Further discussion of the atomic sulfur abundance, and its implications for the chemistry of
interstellar sulfur molecules, is deferred to a future paper.
5.2. Line emission from ionized species
The PCA shows a very tight correlation between the [FeII], [PII] and [SiII] line intensities
in all the regions we observed, a result confirmed by the correlation plots shown in Figures 20
([FeII] versus [SiII] intensity) and 21 ([PII] versus [SiII] intensity). These close correlations
are unsurprising, given the similar appearance potentials (all < 13.6 eV) of Fe+, P+ and Si+:
all three ions are expected behind shocks that are fast enough to cause significant ionization,
and can also be produced in the ionizing "precursors" of fast shocks, where ultraviolet
radiation propagates upstream and can photoionize the gas before it even enters the shock
front. For shocks propagating in molecular gas, the calculations of HM89 suggest that a
shock velocity of at least 35 kms−1 is required to produce detectable fine structure emissions
from [FeII] and [SiII].
One additional feature of the data, however, which is not revealed by our PCA of mean-
subtracted line maps, is that the correlation plots typically show positive intercepts on the
horizontal [SiII] axis. As in the case of H2 S(0), this behavior implies the presence of an
extended gas component within which [SiII] is enhanced relative to [FeII] and [PII].
Four ions of appearance potential greater than 13.6 eV are detected in one or more
region: Ne+, Ne++, Fe++ and S++. Such ions, which have been observed widely in supernova
remnants (e.g. Raymond et al. 1997), are produced only in faster shocks. For example, when
∼ 80 kms−1 are needed to ionize Ne
shocks propagate in molecular gas clouds, velocities >
(HM89). These lines are generally well-correlated with each other, with the curious exception
of the two [SIII] lines (19 and 33 µm); in W44, the [SIII] lines are anticorrelated with the
other emissions from ionized species. We speculate that this behavior could represent an
excitation effect, the critical density for the [SIII] transitions being smaller than those for
[NeII] and [NeIII]. The [SIII] 18.7/33.5 µm and [NeIII] 15.6/36.0 µm line ratios both provide
useful constraints on the electron density in the emitting region (Alexander et al. 1999). The
observed [SIII] 18.7/33.5 µm line ratio was 0.78, 0.54 and 0.55 respectively in W44, W28,
and 3C391, the three sources in which [SIII] emissions were detected. All three values are
consistent with the low density limit obtained by Alexander et al. (1999; 0.6 at a temperature
of 104 K, or 0.7 at a temperature of 2 × 104 K). Conservatively estimating the likely error
on the ratio as ±35%, assuming that the gas temperature is at least 104 K, and making use
of the results presented by Alexander et al. (1999; their Figure 3), we obtain upper limits
on the electron density of 600, 200, and 200 cm−3 for W44, W28, and 3C391. For all three
– 14 –
sources, our best estimate of the [SIII] 18.7/33.5 µm line ratio is somewhat smaller than – but
probably consistent with – that obtained by Oliva et al. (1989) from ISO observations of the
young supernova remnant RCW 103 (0.89 ± 0.18). The observed [NeIII] 15.6/36.0 µm ratio
was ∼ 11 in both W28 and 3C391 (and was consistent with 11 in W44 and IC443C where
the 36.0 µm transition was not detected). Once again, this value is similar to that observed
by Oliva et al. (1989) toward RCW 103 (10 ± 2), and is consistent with the low density limit
obtained by Alexander et al. (1999) (11 at temperature of both 104 K and 2 × 104 K), but
in this case the implied upper limit on the electron density is ∼ 104 cm−3 for both W44 and
3C391. Thus the observed 18.7/33.5 µm and [NeIII] 15.6/36.0 µm line ratios are entirely
consistent with a picture in which the [SIII] emissions originate in a spatially distinct gas
component of lower density than that responsible for the observed [NeIII] emissions. In a
future paper, we will present a more detailed analysis of the emission from ionized species
in these sources, in the context of theoretical models for fast shocks and PDRs.
6. Summary
1) We have carried out spectroscopic mapping observations, covering the 5.2−37 µm spectral
region, toward the supernova remnants W44, W28, IC443C, and 3C391, with the use of the
Infrared Spectrograph (IRS) of the Spitzer Space Telescope. In each case, a region ∼ 1′ × 1′
has been mapped. Except in 3C391, which is more distant and has a much smaller angular
size than the other objects we observed, the mapped regions cover less than 1 percent of the
total radio-emitting area of the supernova remnant.
2) These observations, performed using the Short-Low, Short-High and Long-High modules
of the IRS, have led to the detection of a total of 15 fine structure transitions of Ne+, Ne++
(2 transitions), Si+, P+, S, S++ (2 transitions), Cl+, Fe+ (5 transitions), and Fe++; the S(0)
– S(7) pure rotational lines of molecular hydrogen; and the R(3) and R(4) transitions of
hydrogen deuteride (reported previously by Neufeld et al. 2006b).
3) We have performed a principal component analysis (PCA) of the spectral line maps
obtained from our observations, with the goal of characterizing the differences and similarities
between the spatial distributions of the various line emissions. In W44, the PCA reveals that
the observed emission lines fall into five distinct groups, each of which exhibits a distinct
spatial distribution : (1) lines of S and H2 (J > 2); (2) the H2 S(0) line; (3) lines of ions with
appearance potentials less than 13.6 eV; (4) lines of ions with appearance potentials greater
than 13.6 eV, not including S++; (5) lines of S++. In the other sources, some of these groups
share a common distribution, but in no case do the spectral lines within a group defined
above fail to be strongly correlated with each other. This behavior suggests that the line
– 15 –
emissions we have observed originate in at least five physically distinct components (some of
which have similar distributions – as projected onto the sky – in sources other than W44).
4) The observed lines of S and H2 (J > 2) likely originate in molecular material subject to a
slow, nondissociative shock that is driven by the overpressure within the supernova remnant.
For the case of IC443C, where the gas pressure is constrained by the HD R(4)/R(3) ratio
(Neufeld et al. 2006b), we estimate the atomic sulfur abundance as ∼ 10−6 relative to H2
(with a large uncertainty), a value corresponding to ∼ 4% of the solar abundance of sulfur.
Given the possible presence of S+ in the [SI] emitting-region, this observed atomic sulfur
abundance might be entirely consistent with a negligible depletion of sulfur onto grains and
a negligible fraction of the sulfur nuclei being bound within gas-phase molecules.
5) The H2 S(0) line is typically more extended than any of the other spectral lines we
have observed. In W28, it shows clear evidence for both a shock-excited component and a
low-density diffuse emission component.
6) The fine structure emissions from singly- and doubly-charged ions originate primarily in
dissociative shock fronts with a range of shock velocities. Those ions with appearance poten-
tials greater than 13.6 eV originate in faster shocks than – and show a different distribution
from – ions with appearance potentials less than 13.6 eV. In W44 the S++ emissions show
a very different distribution from those of other ions of similar appearance potential (e.g.
Ne+), a behavior which may reflect the lower critical density for the observed S++ emission
lines.
This work, which was supported in part by RSA agreement 1263841, is based on ob-
servations made with the Spitzer Space Telescope, which is operated by the Jet Propulsion
Laboratory, California Institute of Technology, under a NASA contract. D.A.N. and P.S.
gratefully acknowledge the additional support of grant NAG5-13114 from NASA's Long Term
Space Astrophysics (LTSA) Research Program. We thank the anonymous referee for a very
careful reading of the manuscript and a number of valuable suggestions.
REFERENCES
Alexander, T., Sturm, E., Lutz, D., Sternberg, A., Netzer, H., & Genzel, R. 1999, ApJ, 512,
204
Asplund, M., Grevesse, N., & Sauval, A. J. 2005, ASP Conf. Ser. 336: Cosmic Abundances
as Records of Stellar Evolution and Nucleosynthesis, 336, 25
– 16 –
Braun, R., & Strom, R. G. 1986, A&AS, 63, 345
Burton, M. G., Geballe, T. R., Brand, P. W. J. L.,& Webster, A. S. 1988, MNRAS, 231, 617
Burton, M. G., Hollenbach, D. J., Haas, M. R., & Erickson, E. F. 1990, ApJ, 355, 197
Cesarsky, D., Cox, P., Pineau des Forets, G., van Dishoeck, E. F., Boulanger, F., & Wright,
C. M. 1999, A&A, 348, 945
Caswell, J.L., Dulk, G.A., Goss, W.M., Radhakrishnan, V. & Green, A.J. 1971, A&A, 12,
271
Chevalier, R. A. 1999, ApJ, 511, 798
Claussen, M. J., Frail, D. A., Goss, W. M., & Gaume, R. A. 1997, ApJ, 489, 143
Cox, D. P., & Smith, B. W. 1974, ApJ, 189, L105
Cox, D. P., Shelton, R. L., Maciejewski, W., Smith, R. K., Plewa, T., Pawl, A., & R´ozyczka,
M. 1999, ApJ, 524, 179
DeNoyer, L. K. 1978, MNRAS, 183, 187
DeNoyer, L. K. 1979a, ApJ, 228, L41
DeNoyer, L. K. 1979b, ApJ, 232, L165
DeNoyer, L. K. & Frerking, M. A. 1981, ApJ, 246, L37
Dickman, R. L., Snell, R. L., Ziurys, L. M., & Huang, Y.-L. 1992, ApJ, 400, 203
Draine, B. T. 1980, ApJ, 241, 1021
Dubner, G. M., Vel´azquez, P. F., Goss, W. M., & Holdaway, M. A. 2000, AJ, 120, 1933
Draine, B. T., & McKee, C. F. 1993, ARA&A, 31, 373
Falgarone, E., Verstraete, L., Pineau Des Forets, G., & Hily-Blant, P. 2005, A&A, 433, 997
Fesen, R. A. & Kirshner, R. P. 1980, ApJ, 242, 1023
Fesen, R. A. 1984, ApJ, 281, 658
Frail, D. A., Kulkarni, S. R., & Vasisht, G. 1993, Nature, 365, 136
– 17 –
Frail, D.A., Goss, W.M., Reynoso, E.M., Giacani, E.B., Green, A.J., and Otrupcek, R. 1996,
AJ, 111, 1651
Georgelin, Y. M. 1975, Ph.D. dissertation, Universit´e de Provence, Marseille, France
Giacani, E. B., Dubner, G. M., Kassim, N. E., Frail, D. A., Goss, W. M., Winkler, P. F., &
Williams, B. F. 1997, AJ, 113, 1379
Giveon, U., Sternberg, A., Lutz, D., Feuchtgruber, H., & Pauldrach, A. W. A. 2002, ApJ,
566, 880
Goudis, C. 1976, Ap&SS, 40, 91
Haas, M. R., Hollenbach, D., & Erickson, E. F. 1991, ApJ, 374, 555
Heyer, M. H., & Schloerb, F. P. 1997, ApJ, 475, 173
Higdon, S. J. U., et al. 2004, PASP, 116, 975
Hollenbach, D., & McKee, C. F. 1989, ApJ, 342, 306 (HM89)
Jones, L. R., Smith, A., & Angellini, L. 1993, MNRAS, 265, 631
Le Bourlot, J., Pineau des Forets, G., & Flower, D. R. 1999, MNRAS, 305, 802
Long, K. S., Blair, W. P., Matsui, Y., & White, R. L. 1991, ApJ, 373, 567
Leahy, D. A. 2004, AJ, 127, 2277
McKee, C. F., & Ostriker, J. P. 1977, ApJ, 218, 148
Neufeld, D. A., et al. 2006a, ApJ, 649, 816
Neufeld, D. A., et al. 2006b, ApJ, 647, L33
Oliva, E., Moorwood, A. F. M., Drapatz, S., Lutz, D., & Sturm, E. 1999, A&A, 343, 943
Radhakrishnan, V., Goss, W.M., Murray, J.D. & Brooks, J.W. 1972, ApJS, 24, 49
Raymond, J. C., Blair, W. P., Long, K. S., Vancura, O., Edgar, R. J., Morse, J., Hartigan,
P., & Sanders, W. T. 1997, ApJ, 482, 881
Reach, W.T. & Rho, J.-H. 1996, A&A, 315, L277
Reach, W.T. & Rho, J.-H. 1998, ApJ, 507, L93
– 18 –
Reach, W.T. & Rho, J.-H. 1999, ApJ, 511, 836
Reach, W. T., & Rho, J. 2000, ApJ, 544, 843
Reach, W.T., Rho, J.-H., Jarrett, T.H., & Lagage, P.-O. 2002, ApJ, 564, 302
Reach, W. T., J. Rho, and T. H. Jarrett 2005, ApJ, 618, 297 (RRJ05)
Reynolds, S.P. & Moffett, D.A. 1993, AJ, 105, 2226
Rho, J., Petre, R., Schlegel, E. M., & Hester, J. J. 1994, ApJ, 430, 757
Rho, J.-H. & Petre, R. 1996, ApJ, 467, 698
Rho, J., & Petre, R. 1998, ApJ, 503, L167
Rho, J., & Borkowski, K. J. 2002, ApJ, 575, 201
Richter, M. J., Graham, J. R., Wright, G. S., Kelly, D. M. & Lacy, J. H. 1995a ApJ, 449,
L83
Richter, M. J., Graham, J. R. & Wright, G. S. 1995b, ApJ, 454, 277
Seta, M., et al. 1998, ApJ, 505, 286
Snell, R. L., Hollenbach, D., Howe, J. E., Neufeld, D. A., Kaufman, M. J., Melnick, G. J.,
Bergin, E. A. & Wang, Z. 2005, ApJ, 758, 773
Ungerechts, H., Bergin, E. A., Goldsmith, P. F., Irvine, W. M., Schloerb, F. P., & Snell,
R. L. 1997, ApJ, 482, 245
van den Bergh, S. 1973, A&A, 28, 469
van Dishoeck, E. F., Jansen, D. J., & Phillips, T. G. 1993, A&A, 279, 541
Vel´azquez, P. F., Dubner, G. M., Goss, W. M., & Green, A. J. 2002, AJ, 124, 2145
Weiler, K. W., & Sramek, R. A. 1988, ARA&A, 26, 295
Weingartner, J. C., & Draine, B. T. 2001, ApJ, 548, 296
Wilner, D.J., Reynolds, S.P., & Moffett, D.A. 1998, AJ, 115, 247
Wolszczan, A., Cordes, J. M., & Dewey, R. J. 1991, ApJ, 372, L99
Wootten, A. 1981, ApJ, 245, 105
– 19 –
Yusef-Zadeh, F., Wardle, M., & Roberts, D. A. 2003, ApJ, 583, 267
Ziurys, L. M., Snell, R. L., & Dickman, R. L. 1989, ApJ,341, 857
This preprint was prepared with the AAS LATEX macros v5.2.
– 20 –
Table 1. Details of the observations
Source
Date
Module Observing Map center position in equatorial (J2000)
timea (s)
and Galactic (decimal degrees) coordinates
2005 Apr 24
W44
2005 Apr 24
W44
2005 Oct 11
W44
2005 Apr 19
W28
2005 Apr 19
W28
2005 Apr 19
W28
2005 Apr 17
3C391
2005 Apr 17
3C391
3C391
2005 Apr 18
IC443C 2005 Mar 14
IC443C 2005 Mar 14
IC443C 2005 Mar 14
SL
SH
LH
SL
SH
LH
SL
SH
LH
SL
SH
LH
5552
6625
2650
5552
6625
2650
5552
6625
2650
5552
6625
2650
(R.A., Dec.) = (18h56m28.4s, +01◦29′59.0′′)
(l, b) = (34.8407, −0.4361)
(R.A., Dec.) = (18h01m52.3s, −23◦19′26.0′′)
(l, b) = (6.6856, −0.2956)
(R.A., Dec.) = (18h49m21.9s, −00◦57′22.0′′)
(l, b) = (31.8447, 0.0252)
(R.A., Dec.) = (06h17m42.8s, 22◦21′29.0′′)
(l, b) = (189.2953, 3.0250)
Map size
(arcsec)
58 × 57
57 × 58
59 × 56
58 × 57
57 × 58
59 × 56
58 × 57
57 × 58
59 × 56
58 × 57
57 × 58
59 × 56
aOne cycle was used in each case, with a 60 sec ramp time for LH and SL, and a 30 sec ramp time for SH.
– 21 –
Table 2. Beam-averaged line intensities, in a 25′′ (HPBW) Gaussian beam
Transition
Wavelength
(µm)
Intensity (10−6 erg cm−2 s−1 sr−1)
W44
W28
3C391
IC443C
3/2
−2 P 0
H2 S(0) J = 2 − 0
H2 S(1) J = 3 − 1
H2 S(2) J = 4 − 2
H2 S(3) J = 5 − 3
H2 S(4) J = 6 − 4
H2 S(5) J = 7 − 5
H2 S(6) J = 8 − 6
H2 S(7) J = 9 − 7
HD R(3) J = 4 − 3
HD R(4) J = 5 − 4
S 3P1 −3 P2
Si+ 2P 0
P+ 3P2 −3 P1
Cl+ 3P1 −3 P2
Fe+ 4F9/2 −6 D9/2
Fe+ 4F7/2 −4 F9/2
Fe+ 4F5/2 −4 F7/2
Fe+ 6D7/2 −6 D9/2
Fe+ 6D5/2 −6 D7/2
Ne+ 2P 0
−2 P 0
Ne++ 3P1 −3 P2
Ne++ 3P0 −3 P1
S++ 3P2 −3 P1
S++ 3P1 −3 P0
Fe++ 5D3 −5 D4
1/2
1/2
3/2
7.6
113
135
324
247
808
...a
436
28.2188
17.0348
12.2786
9.6649
8.0251
6.9095
6.1086
5.5112
28.502
23.034
25.2490
34.8141
32.8709
14.3678
5.3403
17.9363
24.5186
25.9882
35.3491
12.8149
15.5545
36.0135 < 2.6 (3σ)
22.4
5.3
156
17.1
1.7
149
1.7
0.7 ± 0.12
1.6 ± 0.5
107
4.7
18.713
33.480
22.925
34.4
43.7
1.2 ± 0.2
36.2
275
392
622
233
589
...a
231
16.2
415
4.4
3.3
151
8.7
2.5
44.7
11.6
247
38.4
3.4 ± 1.1
40.4
74.2
3.5
12.7
60.3
81.7
151
143
661
...a
272
26.8
835
11.9
8.6
388
42.0
12.5
160.4
47.2
538
202
18.0
76.9
140
7.1
12.9
359
499
1930
896
2331
491
1087
5.2
2.1
30.3
97.3
2.8 ± 0.4b
< 1.2 (3σ)
31.6 ± 5.0
2.5
< 1.2 (3σ)
11.3
10.3
15.5
4.2
< 0.9 (3σ)
< 1.4 (3σ)
< 4 (3σ)
< 0.5 (3σ)
aThe H2 S(6) line intensity cannot be determined reliably, because the line is blended
with a strong 6.2 µm PAH feature.
bError bars, where given, are 1σ statistical errors. When error bars are not given,
the uncertainty is dominated by systematic errors, which we estimate to be ≤ 25% (see
Neufeld et al. 2006a).
– 22 –
Table 3. H2 parameters
W44 W28
3C391
IC443C
Rotational state
J=2
J=3
J=4
J=5
J=6
J=7
J=8
J=9
Two-parameter fits
Tw (K)
log10 (Nw/cm−2)
OPRw
Th (K)
log10 (Nh/cm−2)
OPRh
log10 [(Nw + Nh)/cm−2]
log10 (Column density in cm−2)
19.98
19.73
19.52
20.06
19.32
18.71
19.52
18.63
18.40
17.79
18.32
17.84
...a
17.33
17.39
16.96
19.97
19.28
18.54
18.41
17.60
17.78
...a
16.79
20.36
19.83
19.09
18.70
17.69
17.66
...a
16.63
347
20.39
1.58
1134
19.42
...a
20.44
322
20.94
0.93
1039
19.51
...a
20.95
266
20.56
0.65
1099
19.31
...a
20.58
617
20.67
2.42
1225
19.60
3.85
20.70
aThe H2 S(6) line intensity cannot be determined reliably, be-
cause the line is blended with a strong 6.2 µm PAH feature. In
these cases, the OPR cannot be determined for the hot compo-
nent.
– 23 –
W44 OH region E
1 31 00
)
0
0
0
2
J
(
n
o
i
t
a
n
i
l
c
e
D
30 30
30 00
29 30
29 00
28 30
18 56 32 30 28 26 24
Right ascension (J2000)
)
0
0
0
2
J
(
n
o
i
t
a
n
i
l
c
e
D
1 35
30
25
20
15
10
05
Spitzer/IRS
field
18 57 00
18 56 00
18 55 00
Right ascension (J2000)
Fig. 1 – Finder chart for W44. The region mapped by the SH module (black square in upper
panel: white square with arrow in lower panel) is shown relative to the OH maser emission
and radio continuum emission in W44. Upper left panel: the region mapped by the SH
module is marked on Figure 5e of Claussen et al. (1997), which shows the location of the OH
maser spots within W44 OH Region E as well as the 1720 MHz radio continuum intensity
(contours, labeled in mJy/beam). Lower right panel (from Claussen et al. 1997; their Figure
2): the location of the OH maser regions is shown on the 1442 MHz radio continuum image
of Giacani et al. (1997). For an assumed distance to the source of 2.5 kpc, the map in the
lower right panel has a linear size ∼ 25 × 25 pc, and the region mapped by Spitzer has a
linear size ∼ 0.7 × 0.7 pc.
– 24 –
W28 OH region F
-23 18 30
)
0
0
0
2
J
(
n
o
i
t
a
n
i
l
c
e
D
45
19 00
15
30
45
20 00
18 01 56 55 54 53 52 51 50
Right ascension (J2000)
-23 10
)
0
0
0
2
J
(
n
o
i
t
a
n
i
l
20
30
40
c
e
50D
Spitzer/IRS
field
18 03 00 18 02 00 18 01 00 18 00 00
17 59 00
Right ascension (J2000)
Fig. 2 – Finder chart for W28. The region mapped by the SH module (black square in upper
panel: white square with arrow in lower panel) is shown relative to the OH maser emission
and radio continuum emission in W28. Upper left panel: the region mapped by the SH
module is marked on Figure 4f of Claussen et al. (1997), which shows the location of the OH
maser spots within W28 OH Region F as well as the 1720 MHz radio continuum intensity
(contours, labeled in mJy/beam). Lower right panel (from Claussen et al. 1997; their Figure
1): the location of the OH maser regions is shown on the 327 MHz radio continuum image
of Frail et al. (1993). For an assumed distance to the source of 1.9 kpc, the map in the lower
right panel has a linear size ∼ 40 × 30 pc, and the region mapped by Spitzer has a linear
size ∼ 0.55 × 0.55 pc.
– 25 –
3C 391
Spitzer / IRS
)
0
0
0
2
(
n
o
i
t
a
n
i
l
c
e
D
18h 49m 40s
35s
30s
25s
20s
15s
10s
Right Ascension (2000)
Fig. 3 – Finder chart for 3C391. Radio continuum (1446 MHz) image of 3C391 obtained with
a spatial resolution of 6′′ (from Reynolds & Moffett 1993). The Galactic plane is indicated by
the dashed diagonal line. The box centered on the "broad molecular line region" at α = 18h
49m 21.s9, δ = −0o 57′ 22′′ (J2000) delineates the area mapped with the SH module. For an
assumed distance to the source of 9 kpc, the entire map has a linear size ∼ 20 × 20 pc, and
the region mapped by Spitzer has a linear size ∼ 2.5 × 2.5 pc.
– 26 –
Fig. 4 – Finder chart for IC443. An optical image of IC443 with contours of the 1420 MHz
continuum emission overlaid (Leahy 2004). The insert shows an expanded region of the
remnant with the distribution of H2 v = 1 − 0 emission obtained by Burton et al. (1988).
The region mapped by the SH module (black square) is shown in the insert overlaid on the
contours of the H2 emission. For an assumed distance to the source of 1.5 kpc, the insert has
a linear size ∼ 9 ×3.5 pc, and the region mapped by Spitzer has a linear size ∼ 0.45 ×0.45 pc.
– 27 –
Fig. 5 – Average spectra observed for a 25′′ (HPBW) diameter circular aperture centered
at α = 18h 56m 28.s4, δ = +0o 29′ 59′′ (J2000) in W44 (corresponding to offset (0′′, 0′′) in
Figure 10 below). The spectra in the upper two panels were obtained with the SL module,
those in the middle two panels were obtained with the SH module (λ ≤ 19.5 µm) and those
in the lower two panels with the LH module (λ ≥ 19.5 µm). (For each of the SH and LH
spectra, two panels are shown with different vertical scales.) Vertical dotted lines demark
the wavelengths of spectral lines listed in Table 2 (regardless of whether the line is actually
detected in this particular source.) The spectra shown here are not background-subtracted,
since no off-source measurements were made, and thus the continuum flux level must be
regarded as somewhat uncertain.
– 28 –
Fig. 6 – same as for Fig. 5, except for a 25′′ (HPBW) diameter circular aperture centered
at α = 18h 01m 52.s3, δ = −23o 19′ 25′′ (J2000) in W28 (corresponding to offset (0′′, 0′′) in
Figure 11 below).
– 29 –
Fig. 7 – same as for Fig. 5, except for a 25′′ (HPBW) diameter circular aperture centered
at α = 18h 49m 21.s9, δ = −0o 57′ 22′′ (J2000) in 3C391 (corresponding to offset (0′′, 0′′) in
Figure 12 below).
– 30 –
Fig. 8 – same as for Fig. 5, except for a 25′′ (HPBW) diameter circular aperture centered at
α = 06h 17m 42.s5, δ = +22o 21′ 29′′ (J2000) in IC443C (corresponding to offset (−24′′, −20′′)
in Figure 13 below).
– 31 –
Fig. 9 – Excitation conditions for the observed transitions. The horizontal axis shows the
appearance potential – equal to zero for neutral species and equal to the ionization potential
of X(n−1)+ for the case of ion Xn+ – and the vertical axis shows the critical density, ncr, at
which the rate of collisional de-excitation equals the spontaneous radiative rate. For neutral
species, the critical densities apply to de-excitation by H2 molecules at a temperature of
500 K; for ionized species, the critical densities apply to de-excitation by electrons at a
temperature of 104 K. The values adopted for ncr are from the papers of Giveon et al.
(2002), Le Bourlot, Pineau des Forets & Flower (1999) and HM89.
– 32 –
Fig. 10 – H2 S(0), H2 S(1), [SI], [SiII], [NeII], and [SIII] emission line intensities observed
toward W44. The inset boxes demark the regions within which each transition was mapped.
The horizontal and vertical axes show the R.A. (∆α cos δ) and declination (∆δ) offsets in
arcsec relative to α = 18h 56m 28.s4, δ = +0o 29′ 59′′ (J2000).
– 33 –
Fig. 11 – H2 S(0), H2 S(1), [SI], [SiII], [NeII], and [SIII] emission line intensities observed
toward W28. The inset boxes demark the regions within which each transition was mapped.
The horizontal and vertical axes show the R.A. (∆α cos δ) and declination (∆δ) offsets in
arcsec relative to α = 18h 01m 52.s3, δ = −23o 19′ 25′′ (J2000).
– 34 –
Fig. 12 – H2 S(0), H2 S(1), [SI], [SiII], [NeII], and [SIII] emission line intensities observed
toward 3C391. The inset boxes demark the regions within which each transition was mapped.
The horizontal and vertical axes show the R.A. (∆α cos δ) and declination (∆δ) offsets in
arcsec relative to α = 18h 49m 21.s9, δ = −0o 57′ 22′′ (J2000).
– 35 –
Fig. 13 – H2 S(0), H2 S(1), [SI], [SiII], and [NeII] emission line intensities observed toward
IC443C. The inset boxes demark the regions within which each transition was mapped. The
horizontal and vertical axes show the R.A. (∆α cos δ) and declination (∆δ) offsets in arcsec
relative to α = 06h 17m 44.s2, δ = +22o 21′ 49′′ (J2000).
– 36 –
Fig. 14 – Results of PCA analysis for W44. Upper panel: fraction of the variance accounted
for by each component (asterisks), along with the cumulative fraction for the first n
components as a function of n (open squares). Left lower panel: coefficients for the first
and second principal components needed to approximate the maps of each transition. Right
lower panel: coefficients for the second and third principal components. Where multiple
transitions of a single ion have been observed, the labels distinguish between the different
transitions by giving the wavelength rounded to the nearest micron.
– 37 –
Fig. 15 – same as for Fig. 14, except for W28.
– 38 –
Fig. 16 – same as for Fig. 14, except for 3C391.
– 39 –
Fig. 17 – same as for Fig. 14, except for IC443C.
– 40 –
Fig. 18 – H2 rotational diagrams, computed for 25′′ (HPBW) diameter circular apertures
centered at the positions given in the captions to Figures 5 – 8. As usual, the quantity
plotted on the vertical axis is the logarithm of the column density per magnetic substate,
log10(N/gJgN ), where gJ = (2J + 1), J is the rotational quantum number, gN = (2I + 1) is
the nuclear spin degeneracy, and I is the nuclear spin (equal to 3 for states of odd J and 1
for states of even J). Square symbols indicate the measured values, while the lines show the
best two-component fit to the observations (see Table 3). The J = 8 column density cannot
be measured reliably in W44, W28, or 3C391, the H2 S(6) line being blended with a strong
6.2 µm PAH emission feature.
– 41 –
Fig. 19 – Scatter diagrams showing correlations between [SI] and H2 S(0). Each point shows
the mean intensities within a 5′′ × 5′′ square region.
– 42 –
Fig. 20 – Scatter diagrams showing correlations between the [FeII] 26 µm and [SiII] fine
structure lines. Each point shows the mean intensities within a 5′′ × 5′′ square region.
– 43 –
Fig. 21 – Scatter diagrams showing correlations between the [PII] 26 µm and [SiII] fine
structure lines. Each point shows the mean intensities within a 5′′ × 5′′ square region.
|
astro-ph/0105499 | 1 | 0105 | 2001-05-29T16:03:38 | Spectroscopy and orbital periods of four cataclysmic variable stars | [
"astro-ph"
] | We present spectroscopy and orbital periods Porb of four relatively little-studied cataclysmic variable stars. The stars and their periods are: AF Cam, Porb = 0.324(1) d (the daily cycle count is slightly ambiguous); V2069 Cyg (= RX J2123.7+4217), 0.311683(2) d; PG 0935+075, 0.1868(3) d; and KUV 03580+0614, 0.1495(6) d. V2069 Cyg and KUV 03580+0614 both show HeII lambda 4686 emission comparable in strength to H beta. V2069 Cyg appears to be a luminous novalike variable, and the strong HeII suggests it may be an intermediate polar. The period of KUV 03580+0614 is similar to members of the SW Sex-type novalike variables, and it shows the phase-dependent absorption in the Balmer and He I lines typical of this subclass. AF Cam shows absorption features from a K-type secondary, as expected given its rather long orbital period. The secondary spectrum and the outburst magnitude both suggest that AF Cam is about 1 kpc distant. The spectrum of PG 0935+075 resembles that of a dwarf nova at minimum light, with a noticeable contribution from an M-dwarf secondary star. The secondary spectrum and a tentative outburst magnitude both suggest a distance near 500 pc. | astro-ph | astro-ph |
Mon. Not. R. Astron. Soc. 000, 000 -- 000 (1994)
Printed 27 October 2018
(MN plain TEX macros v1.6)
Spectroscopy and orbital periods of four cataclysmic
variable stars
John R. Thorstensen and Cynthia J. Taylor
Department of Physics and Astronomy, Dartmouth College, Hanover, New Hampshire, 03755 USA
ABSTRACT
We present spectroscopy and orbital periods Porb of four relatively little-studied
cataclysmic variable stars. The stars and their periods are: AF Cam, Porb = 0.324(1)
d (the daily cycle count is slightly ambiguous); V2069 Cyg (= RX J2123.7+4217),
0.311683(2) d; PG 0935+075, 0.1868(3) d; and KUV 03580+0614, 0.1495(6) d. V2069
Cyg and KUV 03580+0614 both show HeII λ4686 emission comparable in strength
to Hβ. V2069 Cyg appears to be a luminous novalike variable, and the strong HeII
suggests it may be an intermediate polar. The period of KUV 03580+0614 is similar
to members of the SW Sex-type novalike variables, and it shows the phase-dependent
absorption in the Balmer and He I lines typical of this subclass. AF Cam shows
absorption features from a K-type secondary, as expected given its rather long orbital
period. The secondary spectrum and the outburst magnitude both suggest that AF
Cam is about 1 kpc distant. The spectrum of PG 0935+075 resembles that of a dwarf
nova at minimum light, with a noticeable contribution from an M-dwarf secondary
star. The secondary spectrum and a tentative outburst magnitude both suggest a
distance near 500 pc.
Key words: stars: novae -- stars: cataclysmic variables -- binaries: close -- stars: indi-
vidual -- stars: variables -- stars: fundamental parameters
1
INTRODUCTION
Cataclysmic variable stars (CVs) are close binary systems in
which a red dwarf transfers matter onto a white dwarf, gen-
erally by overflowing its Roche critical lobe. Warner (1995)
gives an excellent comprehensive review of CVs.
The orbital period of a cataclysmic is a fundamental
observable which correlates with evolutionary state and out-
burst type. In addition, the Ritter & Kolb (1998) catalog,
which is heavily used by the CV community, includes only
those systems with known or suspected binary periods, with
the result that a system remains largely 'beneath the radar'
until a period is measured. Photometry yields incontrovert-
ible orbital periods when eclipses are present, but other mod-
ulations can masquerade as orbital periods. Radial-velocity
spectroscopy is therefore the most reliable technique for de-
termining orbital periods of non-eclipsing CVs. The long cu-
mulative exposures needed to find reliable periods are useful
for characterizing the stars in other ways as well.
We present here spectra and radial velocity periods for
four CVs. Section 2 details the techniques common to the
studies, Section 3 gives background information and results
for the individual stars, and Section 4 contains a brief dis-
cussion.
2 TECHNIQUES
All the observations (summarized in Table 1) are from the
MDM Observatory on Kitt Peak, Arizona. The 1998 Jan-
uary observations of KUV 03580+0614 are from the 1.3
c(cid:13) 1994 RAS
m McGraw-Hill telescope and Mark III spectrograph, and
all others are from the 2.4 m Hiltner telescope and modu-
lar spectrograph. We observed comparison lamps frequently,
and checks of the λ5577 night-sky line show that our wave-
length scale is typically stable to < 10 km s−1. When the
weather appeared photometric, we observed flux standards
in twilight. Even so, our absolute fluxes are not expected
to be accurate to much better than 30 per cent, because
of occasional clouds and variable losses at the spectrograph
slits (1 arcsec at the 2.4 m., 2.2 arcsec at the 1.3 m). Fur-
thermore, for unknown reasons the modular spectrograph
produces wavelike distortions in the continua; these appear
to average out in sums of many exposures. Table 2 contains
measurements of the average fluxed spectra.
The data were reduced to counts vs. wavelength at the
observatory, using standard procedures within IRAF. Ra-
dial velocities of emission lines were measured using convo-
lution algorithms described by Schneider & Young (1980)
and Shafter (1983). For absorption velocities, we used the
cross-correlation algorithms of Tonry & Davis (1979), as im-
plemented in the rvsao package (Kurtz & Mink 1998). To
search for periods we used the 'residual-gram' method de-
scribed by Thorstensen et al. (1996), and to test alias choices
in doubtful cases we used the Monte Carlo method explained
by Thorstensen & Freed (1985). With the rough period es-
tablished, we fit the velocities with sinusoids, the parameters
of which are in Table 3.
We note one simple innovation in these otherwise stan-
dard procedures. As part of the spectral reduction process,
2
J. R. Thorstensen & C. J. Taylor
Table 1. Observing Log
Run
Nights N exposure
V2069 Cyg
1997 Sep
1997 Dec
1998 Sep
1999 Jun
1999 Oct
AF Cam
2000 Jan
3
3
1
3
4
5
PG 0935+075
3
1996 Apr
2000 Jan
4
KUV 03580+0614
1997 Dec
1998 Jan
1999 Oct
3
4
4
(s)
360
480
480
300
240
23
12
3
9
7
Range
(A)
4000 -- 7500
4000 -- 7500
4040 -- 7500
4230 -- 7580
4230 -- 7550
∆λ
(A)
3.7
3.7
3.7
3.6
3.9
32
360
4210 -- 7560
3.6
27
21
41
7
57
480
480
480
900
480
4208 -- 6780
4210 -- 7560
4000 -- 7500
4845 -- 6865
4230 -- 7560
2.7
3.6
3.7
4.5
3.6
Exposure times listed are typical values for each run.
∆λ is the FWHM resolution determined from fits to
night-sky features.
IRAF can estimate the noise in each spectral bin from
the readout noise, detector gain, background, and so on.
We have modified the convolution algorithm to compute
a counting-statistics uncertainty in the radial velocities by
propagating these noise estimates through the calculation.
These a priori estimates allow the data to be more optimally
weighted in period finding and curve fitting.
3 THE INDIVIDUAL STARS
3.1 AF Camelopardalis
This object is listed in the General Catalog of Variable stars
(Kholopov et al. 1988) as a U Gem-type dwarf nova with
13.4 ≤ mpg ≤ 17.6. It is the only star here with a useful
period estimate in the literature. Early indications of a very
short period (Szkody & Mateo 1986, Howell and Szkody
1988) prompted an emission-line radial velocity study by
Szkody and Howell (1989). They obtained very limited data
which indicated a 5 -- 6 h period, leading Ritter & Kolb (1998)
to tabulate Porb = 0.23 : in their catalogue.
We observed this object in 2000 January, in quiescence.
The mean spectrum (Fig. 1) shows the absorption features
of a late-type secondary star, as would be expected given
the relatively long Porb. The emission lines are conspicuous
and fairly narrow. We quantified the secondary star's con-
tribution by taking K-type main-sequence spectra from the
library of Pickles (1998), scaling them by logarithmically-
spaced factors, and subtracting them from the mean fluxed
AF Cam spectrum, which had been smoothed to match the
lower resolution of Pickles' spectra. The resulting spectra
were plotted and examined by eye to find the best cancella-
tion of the secondary star features. We found acceptable sub-
tractions for stars of type K4, K5, and K7, and a marginal
match was possible for M0. The secondary contribution was
30 -- 40 per cent of the light near λ6500. For the secondary
alone we find V = 18.4, with an uncertainty estimated from
the decomposition and the flux calibration of perhaps ±0.5
mag.
Table 2. Spectral Features.
Wavelength Identification E.W. Flux FWHM
(A)
(A)
(A)
AF Cam (2000 January)
4339
4473
4860
5876
5893
6564
6680
7066
Hγ
HeI λ4471
Hβ
HeI λ5876
NaD
Hα
HeI λ6678
HeI λ7067
33 107:
9:
28:
125
33
37:
7:
−1.8 :
. . .
220
37
4
23
20
3:
V2069 Cyg (1997 December)
4340
4641
4684
4922
5014
5411
5780
5874
5889
5895
6562
6678
7064
Hγ
CIII/NIII
HeII λ4686
HeI λ4921
HeI λ5015
HeII λ5411
DIB
HeI λ5876
Na DI
Na DII
Hα
HeI λ6678
HeI λ7067
10
3:
15
2:
2:
2.5:
-0.4
49
-0.5
-0.9
32
3.8
2.7
PG 0935+075 (2000 January)
4340
4861
4921
5018
5168
5875
6562
6676
Hγ
Hβ
HeI λ4921
HeI λ5015
FeI λ5169
HeI λ5876
Hα
HeI λ6678
93
80
14
16:
9:
23
101
10
KUV 03580+0614 (1999 October)
6
3
10
8.6
0.6:
0.8
-0.4
1.3:
0.4:
0.4
23
2
1.3:
Hγ
CIII/NIII
HeII λ4686
Hβ
HeI λ5015
HeII λ5411
DIB
HeI λ5876
Na D1
Na D2
Hα
HeI λ6678
HeI λ7067
4338
4647
4684
4859
5011
5407
5780
5872
5890
5896
6562
6678
7063
120
35:
180
26
25
30
. . .
13
. . .
. . .
365
42
31
129
138
20:
25
16:
40:
199
18
158
67
220
185
13
14
. . .
22
. . .
. . .
303
27
15
18:
16:
15
15:
10:
16
19
24:
13
26
13
16:
12:
20:
4
3.2
6.3
15
16
15
18:
18
28:
18:
23
20:
25
21
29
20
19
. . .
20
5.8
22:
3.4
4.1
25
25
24:
Wavelengths are as observed. Positive equivalent
widths denote emission. Line fluxes are given in
units of 10−16 erg cm−2 s−1, and their esti-
mated accuracy is ±30 per cent. The secondary
absorption features in AF Cam are not tabulated,
with the exception of Na D. The HeI lines in
PG0935+075 were strongly affected by their cen-
tral absorption, and the Na D1 and HeI λ5876 lines
generally overlapped.
We cross-correlated the λλ5020 − 5860 and λλ5900 −
6500 regions against a composite of several G- and K-type
IAU radial velocity standards, and obtained velocities for
30 of our 32 target spectra, with typical estimated errors
∼ 20 km s−1. We measured the Hα emission line by con-
volving with the derivative of Gaussian, optimized for a 16 A
c(cid:13) 1994 RAS, MNRAS 000, 000 -- 000
Four Cataclysmics
3
Table 3. Fits to the Radial Velocities
Data
T0
(HJD)
P
(d)
K
γ
σ
(km s−1) (km s−1) (km s−1)
AF Cam (absn)
AF Cam (emn)
AF Cam (avg)
2451552.671(4) 0.3242(12)
2451552.497(4) 0.3230(12)
0.3236(9)
. . .
104(5)
58(3)
. . .
V2069 Cyg
2451066.783(2) 0.311683(2)
125(5)
20(4)
15(3)
. . .
12(3)
PG 0935+075
2451552.774(2)
0.1868(3)
KUV 03580+0614 2451470.668(3)
0.1495(6)
86(7)
67(8)
−20(5)
−53(6)
21
12
. . .
20
22
42
N
30
32
53
42
105
Fits are of the form v(t) = γ + K sin[2π(t − T0)/P ], and σ is the uncertainty of a typical
measurement judged from the scatter around the best fit. N is the number of velocities
used.
Fig. 1. Mean flux-calibrated spectrum of AF Cam, from 2000
January. The vertical scale is uncertain by perhaps 30 per cent.
FWHM. The counting-statistics velocity uncertainties here
were typically 10 km s−1. Fig. 2 shows periodograms for the
emission and absorption velocities, both of which favor a pe-
riod near 0.32 d, but with the inevitable daily cycle-count
aliases. The Monte Carlo tests give discriminatory powers
(defined in Thorstensen & Freed 1985) of 0.96 and 0.83 for
the emission and absorption time series respectively, with
correctness likelihoods somewhat higher than this, so the
cycle count is reasonably secure, but frequencies differing by
1 cycle d−1 cannot be excluded absolutely. The data span
7.8 h of hour angle, ordinarily enough to decide daily cycle
counts unambiguously, but the sampling is somewhat sparse
and the period is awkwardly close to an integer submultiple
of one day. On the face of it, Szkody & Howell's (1989) 5 -- 6
hr period favors the ∼ 4 cycle d−1 alias, rather than the ∼ 3
cycle d−1 we adopt; however, their data do not have enough
hour angle span to influence the choice.
Fig. 3 shows the velocities folded on the combined best
period. The sine fits (Table 3) show that the emission line
velocity modulation lags the absorption line velocities by
0.54±0.02 cycles, suggesting as usual that the emission lines
do not trace the white dwarf motion with any precision.
The secondary velocity amplitude suggests a fairly low
orbital inclination, as might be expected from the single-
peaked emission lines. As a purely illustrative example, the
observed 104 km s−1 velocity amplitude of the secondary
would be expected for a 0.6 M⊙ secondary orbiting a 0.7
M⊙ white dwarf with an inclination of 35 degrees. If the late-
type features arise preferentially on the hemisphere facing
c(cid:13) 1994 RAS, MNRAS 000, 000 -- 000
Fig. 2. Period searches of the absorption (top) and emission (bot-
tom) velocities of AF Cam. The figure of merit shown is the in-
verse of (1/N ) Pi(o − c)2/σ2
i , where o is the observed value, c
is the value computed from the best-fit sinusoid at the trial fre-
quency, and σi is the estimated uncertainty of that particular
velocity.
away from the white dwarf, the true K is even smaller, and
the inclination still lower.
The spectral type found here (K4 -- M0) is similar to that
of other relatively long-period CVs (Beuermann et al. 1998).
AF Cam appears very similar to the apparently brighter
system CH UMa. The orbital period of CH UMa is 0.343
d (Friend et al. 1990), only a few per cent longer than AF
Cam. Becker et al. (1982) estimate the spectral type of CH
UMa's secondary as K4 -- M0, identical to our estimate for
AF Cam.
4
J. R. Thorstensen & C. J. Taylor
Fig. 3. Velocities of AF Cam folded on the best period. All data
are plotted twice for continuity. The round dots show the absorp-
tion velocities, and the crosses the Hα emission velocities. The
best-fit sinusoids are overplotted.
The detection of the secondary and the orbital period
constrain the distance. Beuermann et al. (1999) tabulate ab-
solute magnitudes, colors, and estimated radii of a sample
of nearby K and M dwarfs (their Table 3). Scaling these
data, we find that hypothetical 1 R⊙ stars of type K4 and
type M0 should respectively have MV = +6 and +7.2.
The orbital period and Roche constraints (Beuermann et
al. 1998, eqn. 1) yield (R2/R⊙) = f (q)0.92(M2/M⊙)(1/3) at
AF Cam's period, where the subscript 2 refers to the sec-
ondary star, and the function f (q) is within 3 percent of
unity for q = M2/M1 ≤ 1. Because the secondary is likely
to be modified by mass transfer (see, e. g., Beuermann et
al. 1998), we do not assume a main-sequence mass-radius
relation, but rather use the range of evolutionary models cal-
culated by Baraffe & Kolb (2000) as a guide; this may not
cover all possibilities, but it is arguably better than guessing.
At Porb = 8 h their models span 0.36 ≤ M2/M⊙ ≤ 0.99, but
the more massive secondaries are calculated to have spec-
tral types distinctly earlier than observed here; more real-
istically, M2 ≤ 0.7 M⊙. Secondaries ranging from 0.36 to
0.7 M⊙ yield 0.65 < R2/R⊙ < 0.82, corresponding to stars
0.7 ± 0.3 magnitude fainter than otherwise identical stars
with R = R⊙. Combining these calculations and propagat-
ing the uncertainties in quadrature yields MV = 7.3 ± 0.7
for the secondary. Our detection of the secondary therefore
yields yields m − M = 11.1 ± 0.9.
Warner (1987, 1995) gives a relation for dwarf novae be-
tween MV at maximum light, orbital inclination, and orbital
period. The inclination is uncertain, but likely to be fairly
low as noted earlier. Taking i = 30 degrees, these relations
give MV (max) = 3.0 for AF Cam. For the apparent mag-
nitude we use the GCVS outburst magnitude mpg = 13.4,
which should be similar to V for typical outburst colors.
This then yields m − M = 10.4, in reasonable agreement
with the distance estimated from the secondary. The actual
distance is likely to be less, because of extinction corrections
at this low Galactic latitude (b = 2◦.2).
3.2 V2069 Cygni
V2069 Cyg (= RX J2123.7+4217) was discovered through
its X-ray emission by Motch et al. (1996). They presented a
spectrum and a 2.8-hour session of time-resolved CCD pho-
tometry, which showed flickering but no apparent periodic-
ity. It is listed as 'Cyg6' in Downes, Webbink, and Shara's
(1997) catalog and atlas of CVs (hereafter DWS97).
The spectrum (Fig. 4) appears similar to that in Motch
et al. (1996). The He II λ 4686 line is notably strong. The
equivalent width of NaD (which does not show orbital mo-
tion) and the continuum slope both suggest that the redden-
ing is not negligible. The Hα radial velocities were measured
using the derivative of a gaussian as the convolution func-
tion, optimized for 14 A FWHM. The period search (Fig. 5)
yielded a unique choice of cycle count for the 760-day span
of the observations. The resulting period, 0.311683(2) d, or
7.48 h, is determined with the greatest accuracy of those
reported here. The folded velocities (Fig. 6) and fits (Table
3) show why this went so well -- the velocity amplitude K
was was large, and the scatter relatively small, making the
periodic modulation very conspicuous.
V2069 Cyg shows several features in common with V405
Aur (= RX J0558+5353). They are: (1) relatively nar-
row, single-peaked lines; (2) strong HeII λ4686 emission; (2)
Porb > 4 h; (3) a quiet, large-amplitude emission-line radial
velocity curve. V405 Aur has proven to be a very interesting
DQ Her star (= intermediate polar; see Harlaftis & Horne
1999 and references therein). Although Motch et al.'s (1996)
brief light curve did not show obvious periodic pulsation, a
more sensitive search might be rewarded with success.
3.3 PG 0935+075
PG 0935+075 was discovered as an ultraviolet-excess object
in the Palomar-Green survey (Green, Schmidt, & Liebert
1986). Ringwald (1993) obtained a spectrum showing the
strong, broad Balmer and He I emission typical of a cat-
aclysmic binary. DWS97 list the object as 'Leo7'. It is
evidently a little-studied dwarf nova, since it appears at
B = 13.0 in the PG survey, but is much fainter in Ring-
wald's spectrum and in the observations described below.
We observed this object only in 1996 April and 2000
January. The 1996 April observations indicated a periodicity
near 0.19 d, but the daily cycle count was ambiguous. The
2000 January observations, though fewer in number, were ar-
ranged to span 7.35 h of hour angle and therefore decided the
daily cycle count. The periods derived from the two observ-
ing runs differ by almost twice their mutual standard devia-
tion, but are reasonably consistent, with a weighted average
of 0.1868(3) d. Because the two observing runs were so far
apart, there is no unique choice of cycle count between them,
but if the two runs are phase coherent the allowed periods
are expressed by (1371.151 ±0.003 d)/(7345 ±50), where the
denominator is an integer. Because of the mediocre agree-
ment of the individual runs' periods, the cycle-count uncer-
tainty is chosen to yield periods within ±4σ of the weighted
average. Fig. 7 summarizes the period search for the veloci-
ties, and Fig. 8 shows the folded velocities.
The mean spectrum (Fig. 9, Table 2) closely resembles
those of dwarf novae at minimum light. There is a contribu-
tion from an M-dwarf secondary star, which confirms that
the luminosity was fairly low when the data were taken.
c(cid:13) 1994 RAS, MNRAS 000, 000 -- 000
Four Cataclysmics
5
Fig. 4. Mean spectrum of V2069 Cyg. Note the relatively narrow
emission lines and the strength of HeII λ4686.
Fig. 6. Hα radial velocities of V2069 Cyg, folded on the best
period. Different symbols represent different observing runs as
follows: solid dots, 1997 Sept.; filled squares, 1997 Dec.; filled
triangles, 1998 Sept.; stars, 1999 June; and open circles, 1999
Oct.
Fig. 5. Period search of the Hα emission velocities of V2069 Cyg.
The vertical axis is the mean of the inverse square residual at each
frequency. A single frequency is strongly preferred.
It would appear from the spectrum, period, and the mini-
mal information available on variability that this object is a
UGSS.
To quantify the secondary star contribution we used the
subtraction technique described earlier, but this time with
a library of M-dwarf spectra classified by Boeshaar (1976)
and observed with the same instrumental setup. We found
the secondary contribution to be type M3 ± 1, with the
secondary contributing 45 ± 15 per cent of the continuum
at 6500 A. The secondary contribution has V = 19.7 ± 0.5.
Unfortunately, this was too faint for us to measure radial
velocities of the secondary.
Again, we can estimate a distance from the secondary
contribution. Adapting the data from Table 3 of Beuermann
et al. (1999), we find that a hypothetical 1R⊙ star of type M3
± 1 would have MV = 9.2 ± 1.0. The Roche lobe constraint
at this period yields R2/R⊙ = 0.64f (q)(M2/M⊙)(1/3). Turn-
ing again to the evolutionary models of Baraffe & Kolb
(2000), we find 0.17 ≤ M2/M⊙ ≤ 0.6 in the 4 -- 5 hour pe-
riod range, which gives 0.35 ≤ R2/R⊙ ≤ 0.54, ignoring the
slight variation in f (q). This makes the secondary 1.8 ± 0.5
mag fainter than a 1R⊙ star of the same spectral class, so
we estimate MV = 11.0 ± 1.2 for the secondary, which in
c(cid:13) 1994 RAS, MNRAS 000, 000 -- 000
Fig. 7. Period search of Hα radial velocities of PG0935+075.
The full periodogram shows ringing from the unknown number
of cycle counts between the widely-spaced observing runs, so the
curve shown here is formed by connecting the local maxima of
the periodogram with straight lines.
turn yields m − M = 8.7 ± 1.3. The uncertainty in the sec-
ondary's spectral type dominates the error budget. For com-
parison, Warner's (1987) maximum light relationship yields
MV (max) = 4.6 at this period. Taking the inclination cor-
rection as zero and V at minimum light as 13.0 (both quite
uncertain) yields m − M = 8.4. The good agreement should
probably be interpreted as supporting the dwarf nova clas-
sification of this star (that is, on the one occasion it was
seen to be bright, it was about as bright as expected for
a dwarf nova), rather than as corroborating the distance.
At b = +40.2, the extinction is unlikely to be significant.
If m − M = 8.7, PG0935+075 lies some 350 pc from the
Galactic plane.
6
J. R. Thorstensen & C. J. Taylor
Fig. 8. Hα radial velocities of PG0935+075 folded on the best
period. The period used reflects an essentially arbitrary choice
of cycle count between observing runs. Filled triangles are from
1996 April, and filled squares from 2000 January.
Fig. 10. Period search of Hα radial velocities of KUV 03580+
0614. The curve shown here is formed by connecting the local
maxima of the periodogram with straight lines.
Fig. 9. Mean spectrum of PG0935+075 from 1999 October. Note
the M dwarf features.
3.4 KUV 03580+0614
Wegner & Boley (1993) discovered that KUV 03580+0614
shows emission lines, and flagged it as a cataclysmic variable
candidate. DWS97 list it as 'Tau2'.
The 1997 December velocities indicated a period near
0.15 d, but the daily cycle count was not securely estab-
lished. The 1999 October data are slightly more exten-
sive and the velocities show less scatter; they unambigu-
ously confirm the 0.15 d period. Sinusoidal fits to velocities
from the two observing runs starting near this frequency
gave essentially the same period, the weighted average being
0.1495 ± 0.0006 d. A period search of the combined veloci-
ties yields a slightly longer best-fit period near 0.1502 d. The
cycle count between the observing runs is not determined,
but the periods within ±3σ = 0.0018 d of the best overall
period can be expressed as (671.772 ± 0.005 d)/(4472 ± 53),
the denominator being integer. Figures 10 and 11 show the
period search and the folded radial velocities.
Fig. 12 shows the mean spectrum from 1999 October;
the 1997 spectrum appeared generally similar with a slightly
Fig. 11. Hα radial velocities of KUV 03580+0614 folded on
the best period. The period used reflects an essentially arbitrary
choice of cycle count between observing runs. Filled squares are
from 1997 December, filled triangles from 1998 January, and solid
circles are from 1999 October.
lower flux level. The spectrum is that of a novalike vari-
able, with HeII λ4686 similar in strength to Hβ; the con-
tinuum is quite blue, approximately Fλ ∝ λ−1.7. Fig. 13
shows a single-trailed representation of the 1999 October
data, prepared using phase-averaging techniques (Taylor,
Thorstensen, & Patterson 1999) The He I features especially
show the distinctive phase-dependent absorption character-
istic of the SW Sex stars (Thorstensen et al. 1991, Taylor
et al. 1999, Dhillon, Marsh, & Jones 1998). In the SW Sex
stars the Balmer emission velocities typically lag the ex-
pected white-dwarf motion by ∼ 0.2 cycle. Assuming this
to be the case here, we infer that white dwarf superior con-
junction should occur near φ = 0.3 in the phase convention
we are using. In most SW Sex stars the phase of white dwarf
superior conjunction is marked by eclipses, and the absorp-
tion reaches maximum strength approximately opposite the
eclipse, at 'phase 0.5' in the eclipse ephemeris. This should
correspond to φ ∼ 0.8 in the present phase convention. The
c(cid:13) 1994 RAS, MNRAS 000, 000 -- 000
Four Cataclysmics
7
ACKNOWLEDGMENTS
We thank Joe Patterson and Bob Fried for communicat-
ing that KUV 03580+0614 does not eclipse, and Patrick
Schmeer for a discussion of the variability of PG0935 +075.
The US National Science Foundation supported this work
through grants AST 9314787 and AST 9987334. The MDM
Observatory staff provided cheerful and efficient help.
REFERENCES
Baraffe, I., Kolb, U. 2000, MNRAS, 318, 354
Becker, R. H., Wilson, A. S., Pravdo, S. H., Chanan, G. A. 1982,
MNRAS, 201, 265
Beuermann, K., Baraffe, I., Kolb, U., Weichhold, M. 1998, A&A,
339, 518
Beuermann, K., Baraffe, I., Hauschildt, P. 1999, A&A, 348, 524
Boeshaar, P. 1976, Ph. D. Thesis, Ohio State University
Dhillon, V. S., Marsh, T. R., Jones, D. H. P. 1998, MNRAS, 291,
694
Downes, R. A., Webbink, R. F., Shara, M. M. 1997, PASP, 109,
345
Friend, M. T., Martin, J. S., Connon Smith, R., Jones, D. H. P.
1990, MNRAS, 246, 654
Green, R. F., Schmidt, M., Liebert, J. 1986, ApJS, 61, 305
Harlaftis, E. T., Horne, K. 1999, MNRAS, 305, 437
Hellier, C. 1998, PASP, 110, 420
Howell, S. B., Szkody, P. 1988, PASP, 100, 224
Kholopov, P. N. et al., 1988, General Catalogue of Variable Stars,
Moscow: Nauka Publishing House.
Kurtz, M. J., Mink, D. J. 1998, PASP, 110, 934
Motch, C., Haberl, F., Guillout, P., Pakull, M., Reinsch, K.,
Krautter, J. 1996, A&A, 307, 459
Patterson, J., Skillman, D. R. 1994, PASP, 106, 1141
Pickles, A. J. 1998, PASP, 110, 863
Ringwald, F. A. 1993, PhD thesis, Dartmouth College
Ritter, H., Kolb, U. 1998, A&AS, 129, 83
Schneider, D. P., Young, P. 1980, ApJ, 238, 946
Shafter, A. W. 1983, ApJ, 267, 222
Szkody, P., Mateo, M. 1986, AJ, 92, 483
Szkody, P., Howell, S. B., 1989, AJ, 97, 1176
Taylor, C. J., Thorstensen, J. R., Patterson, J. 1999, PASP, 111,
184
Thorstensen, J. R., Freed, I. W. 1985, AJ, 90, 2082
Thorstensen, J. R., Patterson, J. Thomas, G., Shambrook, A.
1996, PASP, 108, 73
Tonry, J., Davis, M. 1979, AJ, 84, 1511
Warner, B. 1987, MNRAS, 227, 23
Warner, B. 1995, Cataclysmic Variable Stars (Cambridge: Cam-
bridge U. Press)
Wegner, G., Boley, F. I. 1993, AJ, 105, 660
This paper has been produced using the Royal Astronomical
Society/Blackwell Science TEX macros.
Fig. 12. Mean spectrum of KUV 03580+0614 from 1999 October.
Note the strong continuum and HeII λ4686 emission.
observed absorption is strongest around φ ∼ 0.73, in fair
agreement with expectation. Close examination of Fig. 13
also shows a red-to-blue drift of the absorption features as
they strengthen, a behaviour seen in other SW Sex stars.
The orbital period is also similar to other examples of the
class.
A search for eclipses by R. Fried and J. Patterson
(private communication, 2000) proved negative. Thus KUV
03580+0614 joins the ranks of stars which behave spectro-
scopically like SW Sex stars, but which do not eclipse (e.g.,
Taylor et al. 1999).
4 DISCUSSION
The spectra and orbital periods presented here show all
these objects to be fairly typical examples of their classes.
The two dwarf novae (AF Cam and PG0935+075) are of
interest largely because their secondaries are detected, which
allows us to determine distances and further characterise CV
secondaries. PG0935+075 should be monitored for further
outbursts to confirm its variability type.
The two novalikes, V2069 Cyg and KUV 03580+0614,
are potentially more interesting as individuals. V2069 Cyg
shows similarities with V405 Aur, which has proven to be a
very interesting DQ Her star. SW Sex stars frequently show
interesting 'permanent superhumps' (which might more ac-
curately be called persistent superhumps) in their light
curves (Patterson & Skillman 1994). Thus more thorough
time-series photometry of these stars may prove interesting.
KUV 03580+0614 is a good example of a star showing
the phase-dependent absorption of the SW Sex phenomenon,
but no eclipses. The growing number of non-eclipsing SW
Sex stars presents a challenge to scenarios which require the
line of sight to graze the rim of the disk (e.g., Hellier 1998).
If the light being absorbed arises near the disk center, the
absorbing material must be rather far from the disk plane (to
be visible at non-eclipsing inclinations) and not azimuthally
symmetric (to be visible only at certain phases). How this
material gets there is unknown.
Fig. 13. (separate jpg file) Single-trailed representation of the KUV 03580+0614 spectra from 1999 October. The greyscale in the lower
panel is chosen to emphasize weaker features. The upper panel shows the same data, scaled to make visible the behaviour of the stronger
emission lines. The scale is negative (black = bright). In each panel, the data are ordered by phase and shown twice for continuity. The
individual spectra were divided by a fitted continuum. Note the phase-dependent absorption features, especially prominent in the He I
lines (λλ5876, 4921, 5015, and 6678).
c(cid:13) 1994 RAS, MNRAS 000, 000 -- 000
This figure "thorfig13.jpg" is available in "jpg"(cid:10) format from:
http://arxiv.org/ps/astro-ph/0105499v1
|
astro-ph/0506011 | 1 | 0506 | 2005-06-01T08:45:17 | Direct detection of exo-planets: GQ Lupi | [
"astro-ph"
] | We present a comparison of our VLT/NACO K-band spectrum of the GQ Lupi companion with the new GAIA-dusty model atmosphere grid for T=2000 and 2900 K and log g from 0 to 4. Then, we discuss the mass estimate for GQ Lup companion. | astro-ph | astro-ph |
Direct detection of exo-planets: GQ Lupi⋆
Ralph Neuhauser1, Eike Guenther2, and Peter Hauschildt3
1 Astrophysikalisches Institut und Universitats-Sternwarte, Schillergasschen 2,
D-07745 Jena, Germany [email protected]
2 Thuringer Landessternwarte Tautenburg, D-07778 Tautenburg, Germany
3 Hamburger Sternwarte, Gojenbergsweg 112, D-21029 Hamburg, Germany
1 Introduction: GQ Lupi and its companion
Since several years, we have been searching for sub-stellar companions around
young (up to 100 Myrs) nearby (up to 150 pc) stars, both brown dwarfs and
giant planets. Young sub-stellar objects are self-luminous due to ongoing
contraction and accretion, so that young stars are good targets. We have
found several brown dwarf companion candidates and confirmed three of
them by both proper motion and spectroscopy: TWA-5 B, HR 7329 B, and
GSC 8047 B, all being few Myr young late M-type brown dwarfs.
With K-band imaging using VLT/NaCo, Subaru/CIAO, and HST/PC,
we detected a 6 mag fainter object 0.7′′ west of the classical T Tauri star GQ
Lup, which is a clearly co-moving companion (≥ 10 σ), but orbital motion
was not yet detectable. The NaCo K-band spectrum yielded ∼ L1-2 (M9-L4)
as spectral type. At 140±50 pc distance (Lupus I cloud), it can be placed into
the H-R diagram. For more details, see Neuhauser et al. (2005). According
to our own calculations following Wuchterl & Tscharnuter (2003), it has 1
to 3 Mjup, but according to Baraffe et al. (2002) and Burrows et al. (1997)
models, it is anywhere between 3 and 42 Mjup. The latter models are not
applicable for the young GQ Lup and its companion, while the former does
take into account the formation and collapse; the Wuchterl & Tscharnuter
(2003) convection model is calibrated on the Sun.
2 Comparison with model atmospheres
We use the new so-called GAIA-dusty grid (Brott & Hauschildt et al., in
preparation), which is an updated version of the models from Allard et al.
(2001). The updates include improved molecular dissociation constants as
well as more dust species and their opacities. In addition, the models were
computed for a convective mixing length parameter of 1.5 times the pressure
scale height Hp. It uses spherical symmetry, which is most important for
young and sub-stellar objects as GQ Lup A and b with low gravities. The
Allard et al. (2001) AMES-dusty sometimes had problems at very low gravity.
⋆Proceedings ESO Workshop on The power of optical/IR interferometry
2
Ralph Neuhauser, Eike Guenther, and Peter Hauschildt
We compare our K-band spectrum (see Neuhauser et al. 2005) with the
GAIA grid for temperatures Teff = 2000 and 2900 K and for gravities of
log g = 0, 2, and 4 (g in cgs units). See figure 1 for the comparison.
Fig. 1. Comparison of our GAIA-dusty model atmosphere grid with the observed
spectrum overplotted. The high T=2900 K in the upper two spectra do not repro-
duce the water steam absorption in the blue. The model with T=2000 K and log g
=2 fits best (bottom), see text. H2O, Na, and 12CO are indicated.
The model spectrum with Teff = 2000 K is much better than the hotter
temperature (where there is no water vapour absorption in the blue part),
again indicating an early L spectral type. For the gravities, log g = 2 fits
best (log g = 4 is better than log g = 0). This gravity is fully consistent with
the gravity-sensitive CO index measured in our spectrum to be 0.862 ± 0.035,
yielding log g = 2.5 ± 0.8 according to Gorlova et al. (2003). We conclude
log g ≃ 2.0 to 3.3. The good fit indicates that the new spherically symmetric
GAIA-dusty model is better for low gravities than the former AMES-dusty.
It is important to note that the GAIA-dusty models are applicable: They
are independant of age and stand-alone, even without interior models like
the Baraffe et al. (2002) models. They are used, however, as outer boundary
conditions by, e.g., Baraffe et al. (2002) to provide their evolution models with
a more realistic description of how an interior model loses energy through the
atmosphere. Models of the solar atmosphere usually used are also stand-alone
and independant of the interior and age.
Direct detection of exo-planets: GQ Lupi
3
3 Discussion: Mass estimate for the GQ Lup companion
At 140 ± 50 pc distance, with Teff = 2050 ± 450 K and the flux of the
companion (K = 13.10 ± 0.15 mag), we can estimate its radius to be 1.2 ±
0.5 Rjup. With this radius and log g ≃ 2.0 to 3.3, we can estimate its mass
to be ≤ 1 Mjup (for log g ≃ 4 and 2 Rjup, its ∼ 6 Mjup).
The Wuchterl & Tscharnuter (2003) calculations (Fig. 4 in Neuhauser et
al. 2005) indicate ∼ 1 to 3 Mjup, co-eval with the star at ∼ 1 Myr.
Mohanty et al. (2004a) measured gravities for isolated young brown dwarfs
and free-floating planetary mass objects. Their coolest objects have spectral
type M7.5 and gravities as low as log g = 3.125 (GG Tau Bb). This lead
Mohanty et al. (2004b) to mass estimations as low as ∼ 10 Mjup. GQ Lup
A is younger than the Mohanty et al. Upper Sco objects. Its companion is
at least as late in spectral type, probably even cooler. An object younger
and cooler must be lower in mass. The GQ Lup companion is fainter than
the faintest Mohanty et al. object (USco 128, ∼ 9 Mjup), so that the mass
estimate for the GQ Lup companion is ≤ 8 Mjup.
According to Burrows et al. (1997) and Baraffe et al. (2002), the mass
of the companion could be anywhere between 3 and 42 Mjup, as given in
Neuhauser et al. (2005). However, these models are not valid at the young age
of GQ Lup, so that they are not applicable. According to both Mohanty et al.
(2004b) and Close et al. (2005), the Baraffe et al. (2002) models overestimate
the masses of young sub-stellar objects below ∼ 30 Mjup, but underestimate
them above ∼ 40 Mjup. This is consistent with our results.
Hence, according to all valid estimations, the mass of the GQ Lup com-
panion is ∼ 1 to 8 Mjup, i.e. significantly below ∼ 13 Mjup, hence almost
certainly a planet imaged directly, to be called GQ Lup b.
Acknowledgements. We would like to thank Subu Mohanty and Gibor
Basri for very fruitfull discussion about GQ Lupi b. Especially, we would like
to acknowledge Gibor Basri for pointing us to the Mohanty et al. work.
References
1. Allard F., Hauschildt P.H., Alexander D.R., Tamanai A., Schweitzer A., 2001,
ApJ, 556, 357
2. Baraffe I., Chabrier G., Allard F., Hauschildt P.H. 2002, A&A 382, 563
3. Burrows A., Marley M., Hubbard W. et al. 1997, ApJ 491, 856
4. Close L.M., Lenzen R., Guirado J.C., et al., 2005, Nature 433, 286
5. Gorlova N.I., Meyer M.R., Rieke G.H., Liebert J., 2003, ApJ 593, 1074
6. Mohanty S., Basri G., Jayawardhana R., Allard F., Hauschildt P., Ardila D.,
2004a, ApJ 609, 854
7. Mohanty S., Jayawardhana R., Basri G., 2004b, ApJ 609, 885
8. Neuhauser R., Guenther E.W., Wuchterl G., Mugrauer M., Bedalov A.,
Hauschildt P., 2005, A&A 435, L13
9. Wuchterl G., Tscharnuter W.M. 2003, A&A 398, 1081
|
astro-ph/9501102 | 1 | 9501 | 1995-01-26T11:48:00 | The Tully-Fisher relation for low surface brightness galaxies - implications for galaxy evolution | [
"astro-ph"
] | We present the B band Tully-Fisher relation for Low Surface Brightness (LSB) galaxies. These LSB galaxies follow the same Tully-Fisher relation as normal spiral galaxies. This implies that the mass-to-light ratio (M/L) of LSB galaxies is typically a factor of 2 larger than that of normal galaxies of the same total luminosity and morphological type. Since the dynamical mass of a galaxy is related to the rotation velocity and scale length via M \propto V^2 h, at fixed linewidth LSB galaxies must be twice as large as normal galaxies. This is confirmed by examining the relation between scale length and linewidth for LSB and normal galaxies. The universal nature of the Tully-Fisher relation can be understood if LSB galaxies are galaxies with low mass surface density, \sigma. The mass surface density apparently controls the luminosity evolution of a galaxy such as to keep the product \sigma M/L constant. | astro-ph | astro-ph | Mon. Not. R. Astron. Soc. , { ( )
Printed January
The Tully{Fisher relation for low surface brightness
galaxies { implications for galaxy evolution
M.A. Zwaan,
J.M. van der Hulst,
W.J.G. de Blok
and S.S. McGaugh
Kapteyn Astronomical Institute, P.O. Box , AV Groningen, The Netherlands
Institute of Astronomy, Madingley Road, Cambridge CB HA, UK
Accepted date. Received date
5
9
n
a
J
6
2
2
0
1
1
0
5
9
/
h
p
-
o
r
t
s
a
ABSTRACT
We present the B band Tully{Fisher relation for Low Surface Brightness (LSB)
galaxies. These LSB galaxies follow the same Tully{Fisher relation as normal
spiral galaxies. This implies that the mass-to-light ratio (M=L) of LSB galax-
ies is typically a factor of larger than that of normal galaxies of the same
total luminosity and morphological type. Since the dynamical mass of a galaxy
is related to the rotational velocity and scale length via M / V
h, at (cid:12)xed
linewidth LSB galaxies must be twice as large as normal galaxies. This is con-
(cid:12)rmed by examining the relation between scale length and line width for LSB
and normal galaxies. The universal nature of the Tully{Fisher relation can be
understood if LSB galaxies are galaxies with low mass surface density, (cid:22)(cid:27). The
mass surface density apparently controls the luminosity evolution of a galaxy
such as to keep the product (cid:22)(cid:27) M=L constant.
Key words: galaxies: fundamental parameters { galaxies: evolution { galaxies:
spiral { galaxies: distances and redshifts { cosmology: distance scale.
INTRODUCTION
It has become clear in recent years that disk galaxies exhibit a large range in central surface
brightness and scale length: the two parameters that describe an exponential disk. Freeman's
( ) result that the central disk surface brightness of spiral galaxies falls within a narrow
range of (cid:22)
( ) = : (cid:6) : mag=ut
, has now been superseded (e.g., de Jong and van der
B
Kruit ; Davies ). At the faint end of the surface brightness range we (cid:12)nd the so-
called Low Surface Brightness (LSB) galaxies; galaxies with (cid:22)
( ) (cid:25) mag=ut
or fainter.
B
L M.A. Zwaan, J.M. van der Hulst, W.J.G. de Blok and S.S. McGaugh
The selection e(cid:11)ects imposed by the night sky and their low surface brightnesses make LSB
galaxies di(cid:14)cult to detect. Hence, they are underrepresented in catalogues and have until
recently been studied in less detail than normal, High Surface Brightness (HSB) galaxies,
i.e., those galaxies with surface brightnesses approximately equal to the Freeman value.
The Tully{Fisher relation (Tully and Fisher , hereafter T{F), the relation between
absolute magnitude and rotational velocity, is well-established for HSB spiral galaxies. In
this paper we examine whether LSB galaxies follow the same relationship, despite their
much lower central surface brightness (see also Sprayberry et al. ). The position of LSB
galaxies in the luminosity-line width plane is particularly interesting because it will give
information about the mass-to-light ratio (M=L) for these galaxies. This parameter is still
relatively uncertain for LSB galaxies since a detailed study of rotation curves has only been
performed for one LSB galaxy (Bosma, Athanassoula and van der Hulst ).
DATA
. The samples
We selected the data from several samples from the literature, which we brie(cid:13)y discuss here.
McGaugh (McGaugh and Bothun ) and de Blok (de Blok, van der Hulst and Bothun
) studied LSB galaxies with central surface brightnesses (cid:22)
( ) (cid:25) mag=ut
. These
B
galaxies are selected from the UGC (Nilson ) and the LSB catalogue of Schombert et
al. ( , hereafter SBSM). Bulge and disk components are deconvolved from the surface
brightness pro(cid:12)les and exponential functions are (cid:12)t to the disk components. The disk mag-
nitudes are calculated by integrating the surface brightness pro(cid:12)les to in(cid:12)nity. However, the
disk magnitudes ignore any contribution of light from the bulge (which is nevertheless small
for most LSB galaxies in this sample), and will introduce extra light at larger radii. Fortu-
nately, the data necessary to calculate sky-limited aperture magnitudes are available. We
calculated these and used them in our derivation of absolute magnitudes. The inclinations
are derived from the photometric data.
Knezek (Knezek ) investigated giant LSB galaxies with mean blue surface bright-
nesses fainter than mag=ut
. The consequence of this selection criterion is that galaxies
are included with very bright central surface brightnesses. From the total sample we selected
only those galaxies with central surface brightnesses fainter than mag=ut
. We calculat-
ed the disk magnitudes from the published structural parameters; aperture magnitudes are
Tul ly{Fisher relation for low surface brightness galaxies { implications for galaxy evolution
L
Table . Properties of samples
()
()
()
()
()
Sample
N
hW
i
hM
i
h(cid:22)
( )i
R;
B
B
i
b;i
McGaugh
(cid:6)
-. (cid:6) .
. (cid:6) .
De Blok
(cid:6)
-. (cid:6) .
. (cid:6) .
Knezek
(cid:6)
- . (cid:6) .
. (cid:6) .
Total
(cid:6)
-. (cid:6) .
. (cid:6) .
Notes:
() sample name
() number of galaxies in sample
() mean corrected linewidth measured at % of peak intensity in km s
(cid:0)
() mean corrected absolute blue magnitude
() mean blue central surface brightness
not available. We determined the inclinations from the ma jor-to-minor axis ratios from the
UGC. Large errors in inclination corrected linewidth are hence unavoidable.
From these samples, we selected all galaxies with inclinations larger than
and for
(cid:14)
which a linewidth at % of the peak intensity is available. The linewidths were obtained
from the catalogue by SBSM, from LEDA
or from Schneider et al. ( , ). Table
?
presents the characteristics of each sample, including standard deviations.
. Corrections
The two parameters which de(cid:12)ne the T{F relation, the magnitude and the velocity width,
need to be corrected for inclination and absorption e(cid:11)ects. In the following we brie(cid:13)y describe
the corrections which we have applied.
We corrected the magnitudes for Galactic and internal extinction to face-on orientation,
as outlined in Tully and Fouqu(cid:19)e ( , hereafter TFq). The optical depth (cid:28) and the param-
eter f , which describes to what extent the obscuring matter is mixed with the light, are not
well determined for LSB galaxies. Hence, we adopt the standard values derived by TFq for
normal spiral galaxies: (cid:28) = : and f = :.
We corrected the line widths for random motion e(cid:11)ects and inclination as outlined in
TFq. The parameter which describes the importance of random motion is not well known
for LSB galaxies, so once again the value for HSB galaxies is used: W
= km s
.
t
(cid:0)
?
Lyon-Meudon Extragalactic Database
L M.A. Zwaan, J.M. van der Hulst, W.J.G. de Blok and S.S. McGaugh
Figure . Tully{Fisher relation for the sample of LSB galaxies. The long and short dashes represent the (cid:27) and (cid:27) range of
the (cid:12)t to the Broeils sample. One galaxy that enters the sample twice is connected.
For determining the distances to the galaxies we used a Hubble constant of km s
Mpc
.
(cid:0)
(cid:0)
We corrected the redshifts for Galactic rotation and a Virgocentric (cid:13)ow of km s
.
(cid:0)
RESULTS
In Fig. we present the B band T{F relation for the sample of LSB galaxies. As a compar-
ison we use a sample of (cid:12)eld HSB galaxies compiled by Broeils ( ). We used the same
corrections as described in section .. The slope of the T{F relation for these galaxies is
(cid:0): , which is in good agreement with slopes found for cluster samples (e.g., Pierce and
Tully ). The dispersion for the Broeils sample is . mag. This is considerably larger
than what is found for cluster samples, probably due to uncertainties in distances. The (cid:27)
and (cid:27) ranges around the (cid:12)t to the Broeils sample are represented in Fig. by long and
short dashes, respectively. Obviously, the LSB galaxies are indistinguishable from normal
galaxies in the luminosity-line width plane, that is, the T{F relation for LSB galaxies is
identical to that for HSB galaxies.
. Morphological type dependence
For normal spiral galaxies, di(cid:11)erent morphological types have di(cid:11)erent positions in the
luminosity-line width plane (e.g., Kraan-Korteweg, Cameron and Tammann ). In order
Tul ly{Fisher relation for low surface brightness galaxies { implications for galaxy evolution
L
Figure . Tully{Fisher relations binned according to morphological type. The solid circles are LSB galaxies, the open circles
are HSB galaxies from Broeils. The line is a (cid:12)t to all LSB galaxies.
to test whether LSB galaxies follow a similar trend, we made T{F plots binned according
to morphological type, both for the LSB and the HSB sample. These are shown in Fig. .
Only those LSB galaxies which have been unambiguously classi(cid:12)ed have been plotted. From
this (cid:12)gure, it is clear that LSB galaxies follow the same trend with morphological type as
the HSB galaxies.
DISCUSSION
We have shown that LSB and HSB galaxies obey the same, apparently fundamental, T{F
relation. Below we discuss the consequences of this result.
. Mass-to-light ratios
The fact that LSB galaxies follow a normal T{F relation has important implications for
the M=L of these galaxies. We derive a simple relation between the luminosity, line width,
L M.A. Zwaan, J.M. van der Hulst, W.J.G. de Blok and S.S. McGaugh
central surface brightness and M=L to illustrate this. The mass M is proportional to V
h
max
and the total luminosity L is proportional to (cid:6)
h
, where (cid:6)
is the central surface brightness,
h the scale length of the disk and V
the maximum rotational velocity. From these two
max
relations it follows that
M
M
(cid:6)
V
/
/
;
()
max
h
L
so that
V
max
L /
;
()
(cid:6)
(M=L)
where one recognises the T{F relation, L / V
. However, LSB and HSB galaxies are
max
observed to have similar luminosities at a (cid:12)xed line width, in spite of the di(cid:11)erence in (cid:6)
.
This requires that the di(cid:11)erence in surface brightness be compensated by a di(cid:11)erence in
M=L so as to keep the product (cid:6)
(M=L)
constant. The mean central surface brightness
of our sample is h(cid:22)
( )i = :, i.e., . mag fainter than typical of the Broeils sample. In
B
order to account for this factor di(cid:11)erence in (cid:6)
, M=L must be a factor of greater than
for normal spiral galaxies of similar total luminosity. This is in good agreement with the
typical values of M=L for LSB galaxies found by van der Hulst et al. ( ).
. Scale lengths
The normal T{F relation for LSB galaxies and the implication for M=L should be considered
further. Since M / V
h at any given location in the T{F diagram (i.e., at (cid:12)xed luminosity
max
and line width), M=L depends solely on h. Therefore, LSB galaxies which have a higher
M=L must also have scale lengths larger by a factor of than those of the HSB counterparts
with the same line width. The scale length is the preferred parameter for making a fair
comparison between the sizes, as D
severely underestimates the size of a LSB galaxy. The
values of the scale lengths of the Broeils sample can be derived from D
which is related
to the scale length via D
= : h assuming that these galaxies can be represented by
exponential disks with central surface brightnesses of : mag=ut
.
In Fig. we show a plot of the scale length of LSB and HSB galaxies versus line width.
A strong relationship between scale length and line width exists and there is a signi(cid:12)cant
segregation between the LSB and HSB galaxies. Although a minor di(cid:11)erence between the
slopes is present, both groups of galaxies follow a similar trend. In order to derive a reliable
value for the o(cid:11)set between the two samples, we made double regression (cid:12)ts, requiring
Tul ly{Fisher relation for low surface brightness galaxies { implications for galaxy evolution
L
Figure . Scale length versus velocity width, the solid circles are LSB galaxies, the open circles are HSB galaxies from Broeils.
the slopes to be equal. The di(cid:11)erence between the intercepts of the (cid:12)ts is . dex, which
means that LSB galaxies have scale lengths which are . times larger than those of normal
galaxies with the same line width. This is in good agreement with the factor of found for
the di(cid:11)erence in M=L for this sample of galaxies.
. Mass surface density
In addition to comparing M=L ratios and scale lengths at (cid:12)xed line width, one can consider
the implications of a universal T{F relation for the mass surface density. From L / V
max
and V
/ M=h it follows that L / M
=h
, or that
max
M=L / h
=M = = (cid:22)(cid:27) ;
()
where (cid:22)(cid:27) is the average mass surface density. In other words, galaxies with similar mass surface
densities have similar M=L ratios, i.e., the product (cid:22)(cid:27) M=L is constant. The result that LSB
galaxies have higher M=L ratios than HSB galaxies with the same rotational velocity implies
that LSB galaxies have lower mass surface densities. Another way of phrasing this result
is to combine Eq. () and () to (cid:6)
/ (cid:22)(cid:27)
; the surface brightness of a galaxy apparently
is determined by its mass surface density, i.e., LSB galaxies must be less dense than HSB
galaxies. The implication is that the mass surface density may be an important parameter
controlling the luminosity evolution of a galaxy. Less dense galaxies evolve more slowly,
L M.A. Zwaan, J.M. van der Hulst, W.J.G. de Blok and S.S. McGaugh
forming relatively fewer stars in a Hubble time, which results in lower surface brightness
disks and higher M=L ratios. This makes sense given that the stability of a disk, and hence
the star formation activity, is primarily controlled by the mass surface density (e.g., Toomre
; Goldreich and Lynden-Bell ; Kennicutt ; van der Hulst et al. , ). In
this light, the T{F relation simply describes the stability of galaxies.
CONCLUSIONS
We have shown that the B band Tully{Fisher relation for LSB (cid:12)eld galaxies is indistin-
guishable from the relation for their high surface brightness counterparts. This implies that
the Tully{Fisher relation is rather fundamental to spiral disks. Basic theoretical considera-
tions require that LSB galaxies have M=L ratios which are a factor of larger than those
of normal spiral galaxies of comparable total luminosity and morphological type. This dif-
ference implies that at a (cid:12)xed rotational velocity, LSB galaxies should have twice as large
scale lengths as normal galaxies. This is con(cid:12)rmed by the data. The universal nature of the
Tully{Fisher relation requires that galaxies of similar mass surface density have similar M=L
ratios and it is suggestive that the mass surface density of galaxies is a crucial parameter
controlling their luminosity evolution.
ACKNOWLEDGEMENTS
We thank T.S. van Albada for helpful comments on this paper.
REFERENCES
Broeils A.H. , PhD thesis, University of Groningen
de Blok W.J.G., van der Hulst J.M., Bothun G.D., , MNRAS, submitted
Bosma A., Athanassoula E., van der Hulst J.M., , A&A, ,
Davies J.I., , MNRAS, ,
de Jong R.S., van der Kruit P.C., , A&AS, ,
Freeman K.C., , ApJ, ,
Goldreich P., Lynden-Bell D., , MNRAS, ,
Kennicutt R.C., , ApJ, ,
Knezek P.M., , PhD thesis, University of Massachusetts
Kraan-Korteweg R.C., Cameron L.M., Tammann G.A., , ApJ, ,
McGaugh S.S., Bothun G.D., , AJ, ,
Nilson P., , Uppsala General Catalog of Galaxies, Uppsala Astron. Obs. Ann.,
Pierce M.J., Tully R.B., , ApJ, ,
Tul ly{Fisher relation for low surface brightness galaxies { implications for galaxy evolution
L
Schneider S.E., Thuan T.X., Magri C., Wadiak J.E., , ApJS, ,
Schneider S.E., Thuan T.X., Mangum J.G., Miller J., , ApJS, ,
Schombert J.M., Bothun G.D., Schneider S.E., McGaugh S.S., , AJ, ,
Sprayberry D., Bernstein G.M., Impey C.D., Bothun G.D., , ApJ, in press
Toomre A., , ApJ, ,
Tully R.B., Fisher J.R., , A&A, ,
Tully R.B., Fouqu(cid:19)e P., , ApJS, ,
van der Hulst J.M., Skillman E.D., Kennicutt R.C., Bothun G.D., , A&A, ,
van der Hulst J.M., Skillman E.D., Smith T.R., Bothun G.D., McGaugh S.S., de Blok W.J.G., , AJ, ,
This paper has been produced using the Blackwell Scienti(cid:12)c Publications L
T
X style (cid:12)le.
a
E
|
astro-ph/0303441 | 1 | 0303 | 2003-03-19T18:13:40 | Distinct Abundance Patterns in Multiple Damped Ly-alpha Galaxies: Evidence for Truncated Star Formation? | [
"astro-ph"
] | (abridged) Following our previous work on metal abundances of a double damped Ly-alpha system with a line-of-sight separation ~2000 km/s (Ellison & Lopez 2001), we present VLT UVES abundances of 3 new systems spanning a total of \~6000 km/s at z~2.5 toward the southern QSO CTQ247. These abundances are supplemented with echelle observations of another `double' damped Ly-alpha system in the literature. We propose a definition in terms of velocity shift of the sub-class 'multiple damped Ly-alpha system', which is motivated by its possible connection with large-scale structure. We find that the abundance ratio alpha/Fe is systematically low in multiple systems compared with single systems, and with a small scatter. The same behavior is found in 2 more single DLA systems taken from the literature that show evidence of belonging to a galaxy group. After a careful investigation of possible sources of systematic errors, we conclude that the low alpha/Fe ratios in multiple DLAs have a nucleosynthetic origin. We suggest that they could be explained by reduced star formation in multiple damped Ly-alpha systems, possibly due to environmental effects. | astro-ph | astro-ph | Astronomy & Astrophysics manuscript no. lopez
(DOI: will be inserted by hand later)
March 31, 2017
Distinct Abundance Patterns in Multiple Damped Lyα Galaxies:
Evidence for Truncated Star Formation?⋆
Sebastian Lopez1 and Sara L. Ellison2
,
3
1 Departamento de Astronom´ıa, Universidad de Chile, Casilla 36-D, Santiago, Chile
e-mail: [email protected]
2 European Southern Observatory, Casilla 19001, Santiago 19, Chile
3 Pontificia Universidad Cat´olica de Chile, Casilla 306, Santiago 22, Chile
e-mail: [email protected]
Received / Accepted
Abstract. Following our previous work on metal abundances of a double damped Lyα system with a line-of-sight
separation ∼ 2 000 km s−1(Ellison & Lopez 2001), we present VLT UVES abundances of 3 new systems spanning
a total of ∼ 6 000 km s−1 at z ∼ 2.5 toward the southern QSO CTQ247. These abundances are supplemented with
echelle observations of another 'double' damped Lyα system in the literature. We propose a definition in terms of
velocity shift of the sub-class 'multiple damped Lyα system', which is motivated by its possible connection with
large-scale structure. We find that the abundance ratio [S/Fe] is systematically low in multiple systems compared
with single systems, and with a small scatter. The same behavior is found in 2 more single DLA systems taken
from the literature that show evidence of belonging to a galaxy group. Although [Si/Fe] ratios are also generally
lower in multiple DLAs than in single DLAs, the effect is less striking since the scatter is larger and there are
a number of low [Si/Fe] DLAs in the literature. We suggest that this can be explained with a combination of
detection bias and, to a lesser extent, the scatter in ionization corrections for different absorbers. We investigate
whether consistently low α/Fe ratios could be due to dust depletion or ionization corrections and find that the
former effect would emphasize the observed trend of low α/Fe in multiple systems even further. Ionization may
have a minor effect in some cases, but at a level that would not change our conclusions. We thus conclude that
the low α/Fe ratios in multiple DLAs have a nucleosynthetic origin and suggest that they could be explained by
reduced star formation in multiple damped Lyα systems, possibly due to environmental effects. There seems to
be independent evidence for this scenario from the mild odd-even effect and from the relatively high N/α ratios
we observe in these multiple systems.
Key words. Quasars: general -- quasars: absorption lines -- quasars: individual: CTQ247 -- galaxies: evolution --
galaxies: clusters: general
3
0
0
2
r
a
M
9
1
1
v
1
4
4
3
0
3
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
1. Introduction
Damped Lyman Alpha systems (DLAs) in the spec-
tra of high-redshift quasars have promised to be a
powerful probe of early metal enrichment in young
galaxies. However, despite the considerable effort in-
vested to develop this technique, its full potential has
yet to be realized. Although extensive observing cam-
paigns at intermediate resolution have swollen the list
of known DLAs (Ellison et al. 2001b, Lopez et al. 2001
and Peroux et al. 2001 are some of the most recent DLA
surveys), only a fraction have been followed-up at high
resolution to study their chemical properties, investigate
Send offprint requests to: S. Lopez
⋆ The work presented here is based on data obtained at the
ESO Very Large Telescopes, programs 68.A-0361(A) and 66.A-
0660(A).
dust content, photoionization and metallicity evolution.
With the present sample of some 60 abundance mea-
surements some broad trends have been identified (e.g.,
Prochaska & Wolfe 2002), but the global picture of star
formation (SF) history deduced from these patterns of el-
emental abundances -- likely affected by dust depletion pat-
terns (Hou, Boissier & Prantzos 2001) -- is still unclear. In
particular, extracting coherent scenarios of SF histories in
individual DLAs, or subsets of absorbers remains a chal-
lenging prospect with current data.
One of the great advantages of studying high redshift
galaxies in absorption is the lack of selection bias asso-
ciated with cosmological dimming: galaxies are selected
based only on gas cross section, irrespective of their intrin-
sic luminosity. However, this feature may turn out to be a
double-edged sword in that extracting coherent chemical
enrichment patterns is almost certainly hindered by the
2
Sebastian Lopez and Sara L. Ellison: Abundances in Multiple DLAs
Table 1. VLT Observations of CTQ 247
Mode
Wavelength
Exp. Time
Observing Date
FORS2 (GRISM 600B)
FORS2 (GRISM 600R)
UVES Dichroic (390+580)
UVES Dichroic (437+860)
[nm]
345-590
525-745
328-445,476-684
376-500,667-1040
[sec]
600
400
13 500
10 800
Dec. 18 2000
Dec. 18 2000
Oct. 10, 11 and Nov.17, 2001
Nov. 17 2001
mixed morphological representation. Undoubtedly, some
of the scatter in relative abundances comes from vari-
able dust depletion, although system-to-system differences
persist even after correcting for grain fraction (Vladilo
2002). In order to compare elemental trends in DLAs
in a similar manner to, for example, local stellar abun-
dance studies, a significant refinement in our approach is
required. Such a refinement might include isolating par-
ticular subsets of DLAs based either on their morphology
or some other observed property (e.g. spin temperature;
Chengalur & Kanekar 2000).
In this paper, we investigate whether a distinct enrich-
ment trend exists for absorption line systems with close
companions in velocity space, which we will refer to as
'Multiple Damped Lyman Alpha systems', or MDLAs.
2. The Sub-class 'MDLA'
Lopez et al. (2001) identified the existence of three
damped Lyα absorbers
spanning a total of v =
5 900 km s−1 at z ∼ 2.5 toward the southern quasar
CTQ 247, the first ever example of a triple DLA; Ellison
et al. (2001b) discovered a double absorption systems for
which v = 1 800 km s−1 at z ∼ 2 toward Q2314−409;
a similar velocity span as in CTQ 247 is covered by the
two DLAs toward Q2359−02 at z ∼ 2 (Wolfe et al. 1986).
Statistically, these are extremely improbable events if the
proximity of the systems is purely by chance, and only
these few cases are reported in the literature. It is there-
fore feasible to postulate that in these cases the multiple
absorption systems may be in some way related, rather
than chance alignments. However, such large velocities
are difficult to reconcile with today's galaxy groups and
clusters -- i.e., virialized entities -- or even with superclus-
ters. However, at z > 0 current work on small angular
fields shows evidence for very large structures on comoving
scales as large as d ∼ 100 h−1
100 Mpc (∆v ∼ 10 000 km s−1 if
purely due to Hubble flow) at various redshifts. Evidence
for large structure on many Mpc scales has been found
in absorbing gas both across (e.g., Williger et al. 2002;
d ∼ 100 h−1 Mpc at z ∼ 1.3 ) and along the line of sight
(Quashnock, vanden Berk & York 1996; d ∼ 100 h−1
in super-clusters (d ∼ 20 h−1 Mpc
Mpc at z ∼ 3),
at z ∼ 0.8; Haines, Campusano & Clowes 2003) and in
Lyman-break galaxies (Steidel et al. 1998; d ∼ 10 h−1
Mpc at z ∼ 3 ). Large structure has also been reproduced
in CDM N-body simulations (d ∼ 20 h−1 Mpc at z ∼ 1;
Evrard et al. 2002). It therefore seems a feasible (but not
exclusive) possibility that absorption systems separated
by several thousand km s−1 may be associated with very
large scale structure.
Motivated both by the uniqueness of MDLAs and the
possibility of large-structure, we have embarked on a pro-
gram to measure abundances in these systems. We will
define an MDLA as 2 or more DLAs spanning a velocity
range
500 < ∆v < 10 000 km s−1.
(1)
While the upper limit is set to match the limits of the
largest structures known at high redshift, the lower veloc-
ity roughly corresponds to the largest asymptotic veloci-
ties in massive spirals (e.g. Rigopoulou et al. 2002)1. Note
that this a working definition only; although it might be
possible that MDLAs probe some kind of large structure
(but see § 5), our definition does not necessarily imply that
they host the strong C iv absorption used by Quashnock,
vanden Berk & York (1996) or they are associated with
the LBGs identified by Steidel et al. (1996).
From an instrumental point of view,
identifying
MDLAs in Lyα -- in contrast to using metal lines -- will
certainly limit the identification of systems separated by
∆v <∼ 1 000 km s−1. If the discovery spectrum has poor
S/N, obtaining line parameters for close MDLAs will re-
quire both the detection of metal lines, and a good con-
tinuum estimation2.
For N (H i) criterion, we will relax the canonical limit
to include systems with N (H i) > 1020 cm−2. Not only
do systems down to this limit continue to exhibit clear
damping wings, but in the following we show that in such
range ionization does not yet play a significant role when
deriving abundances from low-ions.
The first in-depth study of an MDLA was presented by
Ellison & Lopez (2001; hereafter Paper I). There, the two
DLAs toward Q2314−409 were noted to have low [S/Fe]
ratios, a result which we suggested was due to the pos-
sible impact of environment on the chemical abundances
1 This lower limit will probably exclude the majority
of bound galaxy satellites since most local spirals' satel-
less than 300 km s−1,
lites have velocity differences of
Zaritsky et al. (1997)
2 Two N (H i) = 2 × 1020 cm−2 DLAs separated by ∆v =
1 000 km s−1 will mimic a single system in FWHM ∼ 5 A data
if S/N <
∼ 20, but for ∆v ∼ 500 km s−1 they will do so unless
S/N >
∼ 50.
Sebastian Lopez and Sara L. Ellison: Abundances in Multiple DLAs
3
in DLA protogalaxies. In the present work, we extend the
study of abundances in MDLAs by investigating the triple
DLA toward CTQ247 with new echelle data, and includ-
ing literature abundances of the double system toward
Q2359−02. These new abundances reinforce our previous
suggestion of peculiar relative abundances in MDLAs.
After presenting the new data in § 3, the discussion
of the abundances is given in § 4, where we attempt a
statistical comparison between the 2 populations, and as-
sess possible systematic effects that may bias our result.
A discussion on the nature of MDLAs is outlined in § 5
3. Data Analysis
3.1. Observations and Data Reduction
CTQ 247 was observed in service mode with the UVES
instrument at the ESO Kueyen telescope in October and
November 2001 under good seeing conditions (0.6′′ −1.0′′).
With the 390+580 and 437+860 dichroic modes we cov-
ered from 328 to 1020 nm with two gaps at 576-583 nm
and 852-866 nm. The exposure times were 13 500 seconds
for dichroic 1 and 10 800 seconds for dichroic 2 (Table 1).
After the usual fashion of bias-subtracting and flat-
fielding of the individual CCD frames, the echelle orders
were extracted and reduced interactively with the UVES
pipeline routines (Ballester et al. 2000). The reference Th-
Ar spectra used for wavelength calibration were taken af-
ter each science exposure. The wavelength values were
converted to vacuum heliocentric values and each order
of a given instrumental configuration was binned onto a
common linear wavelength scale of 0.04 A pixel−1. The
reduced orders were then added with a weight according
to the inverse of the flux variances. Finally, the flux values
were normalized by a continuum that was defined using
cubic splines over featureless spectral regions. The spec-
tral resolution is FWHM ∼ 6.7 km s−1, while the typical
signal-to-noise ratio per pixel is S/N∼ 35 − 40.
In addition to the UVES data, a lower resolution
(FWHM ≈ 3 A) spectrum of CTQ 247 was obtained
on December 18 2000 using FORS2 at the ESO Kueyen
Telescope. This spectrum was used to better define the
quasar continuum in the spectral region around the
damped Lyα lines.
3.2. Column densities
We used FITLYMAN and VPFIT3 to fit the line profiles
with theoretical Voigt profiles. All fits were unconstrained
in redshift, Doppler width and column density, unless oth-
erwise stated. In general, we prefer this approach over
the apparent optical depth method (AODM; Savage &
Sembach 1991) when velocity components are not resolved
as is sometimes the case. In addition, when dealing with
transitions that lie in the Lyα forest, such as the S ii
triplet, fitting provides important information to the ex-
tent of possible blending. Despite this, however, we did use
3 Available at http://www.ast.cam.ac.uk/rfc/vpfit.html
the AODM for transitions where the fit failed due to line
blending of many components or poor S/N. We adopted
the up-to-date f -values listed in Prochaska et al. (2001)
and the solar abundances by Grevesse & Sauval (1998),
with updates for O and N by Holweger (2001). Table 2
lists the column densities and derived abundances of all
elements covered by our observations.
The H i column density was determined by fitting 3
components to the damped Lyα and Lyβ profiles in the
FORS2 spectrum. The choice of the low resolution spec-
trum reduces the uncertainties inherent to determining
N (H i) from echelle data, such as badly defined contin-
uum where the damping wings extend over more than one
order. We obtained N (H i) = 1021.13, 1021.09, and 1020.47
cm−2, for the 3 systems centered at zabs = 2.5505, 2.5950
and 2.6215. The theoretical profiles are superposed on the
data in Fig. 1. The internal fit errors were 0.02 dex for all
3 H i measurements but we believe a more realistic error
is given by ±0.1, which is shown by the dotted curve in
the figure. Although the noisy data around Lyβ constrains
H i only marginally, this error seems quite safe, given the
higher S/N at Lyα.
We next give a short description of the metal line fits
in each of the 3 DLAs which we henceforth refer to as
CTQ 247A,B and C.
CTQ247A (zabs = 2.5505): For this DLA we have identi-
fied nonsaturated transitions of N i, Si ii, S ii, Cr ii, Fe ii,
and Zn ii, all of which show 4 velocity components. More
(weaker) components are revealed by stronger transitions,
e.g. Fe ii λ2344, but these do not contribute significantly
to the total column densities. Therefore, in order to avoid
introducing too many parameters at intrinsically weak
transitions in relatively noisy parts of the spectrum, we
proceeded by fitting 4 components. Fig. 2 shows the non-
saturated transitions used and their fits. From an analysis
of the weaker components we estimate that the obtained
column densities encompass ∼ 90 % of the total column
density; exclusion of weaker components will lead to only
a ∼ 0.05 dex difference at log N = 15 cm−2. The S ii 1260
triplet lines fall on the red wings of the damped Lyα lines;
consequently, they were fitted simultaneously with the H i
fixed at the values found for the FORS data, hence the
sub-unity values of the continuum in Fig. 2. In addition,
we determine an AODM measurement for Mn ii and an
upper limit for P ii.
CTQ247B (zabs = 2.5950): This DLA has a relatively sim-
ple velocity structure. Although the strongest Si ii and
Fe ii transitions have 5 velocity components, these span
only ∼ 80 km s−1. As expected for such a high H i, O i is
saturated and we consider it no further here. Conversely,
N i which is often barely detected exhibits a large equiva-
lent width in this system. In fact, two of the λ1134 triplet
transitions are saturated. However, the triplet is relatively
unblended, as are two of the S ii triplet lines. Fig. 3 shows
the nonsaturated transitions used and their fits. There is
4
Sebastian Lopez and Sara L. Ellison: Abundances in Multiple DLAs
Fig. 1. Normalized flux and 1σ VLT FORS2 spectrum of CTQ 247 at FWHM = 3 A resolution showing the Lyα and
Lyβ lines of the triple DLA. Overlaid is a synthetic Voigt profile of three systems with N (H 1)= 1021.13, 1021.09, and
1020.47 cm−2, centered at zabs = 2.5505, 2.5950, and 2.6215. The dotted lines are ±0.1 dex deviations for all N (H i)
values.
an absorption feature at the expected position of P ii λ963
but since we suspect contamination by a Lyα forest line we
only provide an upper limit on N (P ii). As for CTQ247A,
we also provide an AODM measurement for Mn ii.
only values obtained from echelle data. In cases where
numbers are present in several references, we have pre-
ferred the values by Prochaska et al. (2001) for consis-
tency.
CTQ247C (zabs = 2.6215): This is the lowest H i DLA
of the triplet. For this DLA the bulk of the low ions are
divided into 2 'sub-clumps' of 3 and 2 velocity components
separated by ∼ 120 km s−1. Fig. 4 shows the nonsaturated
transitions used and their fits. Since this is the DLA with
the lowest H i of the 3, fewer species are detected and limits
are not very stringent. For example, for sulphur our 3σ
detection limit is log N (S ii) < 14.3. We have attempted
a fit for the N i triplet, but a combination of many weak
components and Lyα blending makes this a very uncertain
result and we regard it as an upper limit due to possible
contamination. On the other hand, the low H i allows us
to profile fit 2 O i transitions, for which each one of the
5 components shows at least one nonsaturated line. Also,
we report one of the few existing measurements of Al ii in
DLAs (line positions were held fixed at the Si ii values),
and one AODM measurement for Al iii.
4. Abundances in MDLAs
In Paper I we focused on the relative α to Fe peak abun-
dances in order to establish whether MDLAs may ex-
hibit distinct SF histories. Similarly, in Fig. 5 the solid
circles depict [α/Fe] vs. [Fe/H] for the 3 MDLAs pre-
sented in this paper (CTQ247A, B, and C), the dou-
ble DLA Q2314-409A, B (Paper I), and Q2359−02 A,
B (Prochaska & Wolfe 1999). The stars are DLAs taken
from the literature. Due to the potential problem of blend-
ing between S lines and the Lyα forest, we have compiled
In the following we show statistically that the sub-
class MDLA is drawn from a different distribution to sin-
gle DLAs. We then discuss to which extent dust and ion-
ization effects might produce such distinction, and what
possible systematic effects may be present in the 2 sam-
ples.
4.1. Low α/Fe in MDLAs
As expected in DLAs, all points in Fig. 5 fall over the
[α/Fe] = 0 mark, as both S and Si are α-chain elements
which are created mainly in massive stars on much shorter
scales than the long-lived iron producing stars. The ra-
tio of [α/Fe]
is usually interpreted as an indicator of
star formation history and the rate at which the differ-
ent elements are released into the ISM. In addition to
the MDLAs in Fig. 5, we have also plotted DLAs with
other neighboring galaxies (usually detected via Lyman
break imaging) in the field. These systems are Q0201+112
(Ellison et al. 2001a) which has 4 identified galaxies in
the vicinity, and Q0000-262 (Molaro et al. 2001) for which
Steidel et al. (1996) found 2 Lyman break galaxies within
∆z = 0.05 of the absorber.
The remarkable property to note in Fig. 5 is that all
MDLAs systematically show low [S/Fe] values with a small
scatter (< 0.2 dex). To quantify this impression, we have
performed a modified KS test that provides the likelihood
of MDLAs with [S/Fe] measurements to be drawn from
a different distribution to single DLAs. We created 1000
realizations of a simulation which recreates the [S/Fe] dis-
Sebastian Lopez and Sara L. Ellison: Abundances in Multiple DLAs
5
CTQ247A
CTQ247B
CTQ247C
N(X)
21.13 ± 0.10
14.55 ± 0.03
[X/H]
...
−2.51 ± 0.10
N(X)
21.09 ± 0.10
15.07 ± 0.02
[X/H]
...
N(X)
20.47 ± 0.10
−1.95 ± 0.10
< 14.36
[X/H]
...
< −2.04
−2.02 ± 0.10
−1.97 ± 0.12b
...
−2.04 ± 0.12
...
< −1.33
...
...
15.19 ± 0.02
12.88 ± 0.07
12.33 ± 0.03a
13.99 ± 0.06
...
< 14.34
...
...
13.60 ± 0.02
−2.37 ± 0.10
...
...
...
...
...
...
X
H i
N i
O i
Al ii
Al iii
Si ii
P ii
S ii
Cr ii
Mn ii
Fe ii
Fe iii
Ni ii
Zn ii
...
...
12.81 ± 0.05a
15.32 ± 0.04
< 13.15
...
...
...
−1.37 ± 0.11
< −1.54
...
...
12.63 ± 0.04a
15.59 ± 0.03
< 12.98
...
...
...
−1.06 ± 0.10
< −1.67
−1.51 ± 0.12
14.82 ± 0.06
13.20 ± 0.03
−1.62 ± 0.10
12.86 ± 0.02a −1.80 ± 0.10
−1.68 ± 0.12
14.95 ± 0.06
...
...
13.94 ± 0.02
12.44 ± 0.05
−1.44 ± 0.10
−1.36 ± 0.11
−1.10 ± 0.11
15.19 ± 0.05
13.37 ± 0.04
−1.41 ± 0.11
12.87 ± 0.01a −1.75 ± 0.10
−1.44 ± 0.10
15.15 ± 0.02
< 13.66
13.86 ± 0.31
12.68 ± 0.02
...
−1.48 ± 0.33
−1.08 ± 0.10
Table 2. Abundance measurements (and 3σ upper limits) for CTQ247. a uses AODM; b N (Al) = N (Al ii) + N (Al iii)
tribution of literature DLAs and MDLAs including the 1
σ quoted errors. For each DLA, an error was drawn at
random from a Gaussian distribution and added to the
observed value4. A KS test was then run on each of the
1000 datasets and the likelihood that the KS probability
PKS was greater than a certain confidence level was calcu-
lated from the number of realizations. In this fashion, we
overcome 2 problems associated with a normal KS statis-
tics: the small number of MDLAs we have at our disposal
and the exclusion of measurement errors.
According to our simulations, the likelihood p that
the two samples are drawn from different distributions is
p = 1.00 for PKS > 95 % (i.e. every one of the 1000 realiza-
tions gives a PKS > 95 %), p = 1.00 for PKS > 98 %, and
even as high as 0.97 for PKS = 99 %. This indicates that
the consistently low [S/Fe] values in MDLAs and other
multiple systems are statistically significant. Since Si and
S are both produced in oxygen burning reactions, [S/Si]
= 0 in all observed Galactic disk stars (Chen et al. 2002)
and we may expect to see similar patterns between S and
Si in DLAs. However, the MDLA [Si/Fe] values have a
slightly larger dispersion and there are a number of low
literature [Si/Fe] values, which renders the two popula-
tions less distinguishable. This may be explained because
low N (S ii) lines are easily contaminated in the Lyα for-
est, whereas low N (Si ii) can be much more easily mea-
sured. If we consider only the literature DLAs with Si and
S measurements, the KS probability that the two [Si/Fe]
samples are consistent is 11%, as opposed to 32% if all
are included. There is an obvious outlier in the [Si/Fe]
distribution associated with system Q2359−02A; in the
following section we argue that this system has anoma-
lously large dust content compared with the rest of the
MDLAs and literature DLAs. Discounting this exception-
4 We treat the upper limit toward DLA B in Q2314−409
as a detection; a lower value than this does not affect the KS
probability
ally dust depleted system reduces the KS probability even
further of all [Si/Fe] measurements to 9%. Therefore, al-
though there is a hint that [Si/Fe] may be low in MDLAs,
the evidence is not as convincing as that for S. We explore
this further in the following section.
4.2. How dust and/or ionization might affect the
observed ratios
Before discussing the implications of the observed low
[α/Fe] ratios in terms of a nucleosynthetic origin, we must
first rule out other (non-intrinsic) systematics. In Paper
I, we have already discussed the issue of dust and pho-
toionization in the double DLA toward Q2314-409. Here,
we embark upon a similar assessment for the CTQ247
triplet. In particular, we investigate whether systematics
may cause (a) the α/Fe ratio to be lower in MDLAs than
in single DLAs; and (b) the α under-abundance in MDLAs
to be more evident in S rather than Si.
First, we consider dust depletion. Although disentan-
gling dust depletion from pure nucleosynthesic effects in
DLAs is still matter of debate, some broad trends can
be established: (1) In no instance is S known to be de-
pleted in the Galactic ISM (e.g. Savage & Sembach 1996).
Conversely, Si is mildly depleted in most environments and
Fe is very easily incorporated into dust. Therefore, dust
depletion will cause an increase in both [S/Fe] and [Si/Fe],
the effect being most pronounced in the former ratio, so
[S/Si] > 0. (2) Zn is undepleted; thus, assuming the same
nucleosynthetic origin as Fe, one expects [Zn/Fe] > 0 in
dusty systems.
Both in CTQ247A and B the [Zn/Fe] shows a signifi-
cant departure from solar, indicating that dust is present.
This is also true for Q2314−409, the other MDLA for
which a Zn abundance is determined; in all cases [Zn/Fe] >
0.3, typical of other DLAs. Therefore, any downward cor-
rection due to dust would further emphasize the low α/Fe
in these MDLAs. It is noteworthy that Q2359−02A, which
6
Sebastian Lopez and Sara L. Ellison: Abundances in Multiple DLAs
Fig. 3. Same as in Fig. 2 but for CTQ247B (zabs =
2.5950)
has the largest [Si/Fe] of all, also shows an unusually large
dust content, [Zn/Fe]=+0.88, explaining its anomalously
large [Si/Fe]. Not only is Q2359−02A relatively highly de-
Fig. 2. UVES spectrum showing nonsaturated transitions
in CTQ247A (zabs = 2.5505). The solid line in the P ii
and Mn ii panels indicates the velocity region used for the
AODM.
Sebastian Lopez and Sara L. Ellison: Abundances in Multiple DLAs
7
Fig. 5. Relative ratios of α/Fe for MDLAs (solid circles),
the DLAs toward Q0201+1120 and Q0000−262 (solid tri-
angles) and DLAs taken from the literature (stars). All
ratios have been corrected to the following solar val-
ues:
log
(Fe/H)⊙ = −4.50 (Grevesse & Sauval 1998). The data
points shown in this plot are listed in Table 3.
log (S/H)⊙ = −4.80, log (Si/H)⊙ = −4.44,
Fig. 4. UVES spectrum showing nonsaturated transitions
in CTQ247C (zabs = 2.6215)
pleted compared to other MDLAs, but also in comparison
with other DLAs which mostly have [Zn/Fe] < 0.6.
Can low dust content in MDLAs compared with sin-
gle DLAs be responsible for lower [S,Si/Fe] in the for-
mer? Fig. 6 shows [S/Si] vs. [S/Fe] for the (few) systems
where both elements are available, including 3 MDLAs.
It can be seen that apart from one DLA with an ex-
tremely large [S/Fe], there is no strong trend for sys-
tems with lower [S/Si] (supposedly less dusty) to show
lower [S/Fe], suggesting that low S/Fe ratios are not
strongly effected by dust. Indeed, the large S abundance
measured in Q0307−49 might be due to Lyα blending
(Bonifacio et al. 2001) and must be considered with cau-
tion. Therefore, low S/Fe is not obviously correlated with
low dust content, at least not to the accuracy we can
Fig. 6. [S/Si] vs. [S/Fe] in MDLAs (circles), single DLAs
(stars), and in the DLA toward Q0000−262 (triangle).
presently measure this ratio. In conclusion, it does not
seem from the present sample that a general trend of low
α/Fe values be driven by any effect of dust, nor do MDLAs
have atypical dust content.
8
Sebastian Lopez and Sara L. Ellison: Abundances in Multiple DLAs
QSO
[Fe/H]
[Zn/Fe]
[Si/Fe]
[S/Fe]
Ref.
Q0013−004
Q0100+13
Q0149+33
Q0201+3634
Q0216+0803A
Q0216+0803B
Q0255+00A
Q0255+00B
Q0307−49
Q0336−01
Q0347−38
Q0741+474
Q0836+11
Q0957+33A
Q0957+33B
HE1104−18
Q1108−07
Q1210+17
Q1215+33
Q1223+17
Q1331+17
Q1409+095A
Q1759+75
Q1946+7658
Q2206−19
Q2206−19
Q2230+02
Q2231−00
Q2237−060
Q2243−603
Q2343+1232
Q2344+12
Q2348−01A
Q2348−01B
CTQ247A
CTQ247B
CTQ247C
Q2314−409A
Q2314−409B
Q2359−02A
Q2359−02A
-1.83
-1.90
-1.77
-0.87
-0.97
-1.06
-1.44
-2.05
-1.96
-1.80
-1.80
-1.93
-1.40
-1.45
-1.87
-1.60
-2.12
-1.15
-1.70
-1.84
-2.06
-2.30
-1.21
-2.53
-0.86
-2.61
-1.17
-1.31
-2.17
-1.25
-1.23
-1.83
-1.39
-2.24
-1.68
-1.44
-2.37
-1.32
-1.87
-1.66
-1.88
1.09
0.31
0.10
0.59
...
...
...
...
...
...
−0.10
...
...
...
...
0.57
...
0.25
0.41
0.22
...
...
...
0.45
...
0.45
0.54
...
0.13
...
...
...
...
0.32
0.36
...
...
...
...
...
...
0.48
0.28
0.46
0.30
0.50
0.50
...
0.41
...
0.46
0.24
0.25
0.45
0.37
0.55
0.32
0.27
0.22
0.25
0.61
0.28
0.40
0.30
0.44
0.30
0.41
0.44
0.36
0.38
0.47
0.09
0.69
0.26
0.31
0.38
0.33
0.27
0.00
0.88
0.29
0.79
0.44
...
...
...
...
...
0.27
1.00
0.39
0.51
0.25
...
...
0.56
0.53
...
...
...
...
0.76
...
0.46
...
...
...
...
...
...
0.40
0.40
...
...
...
0.17
0.34
<1.04
0.32
<0.21
...
...
m
b,c
a,b
a,n
d
d
a
a
e,h
a
a,h,i
a
a
a
a
f
a
a
a
a
a,b,o
k
a
a
a
a
a,b
b
c,d
g
c,l
a
a
a
p
p
p
q
q
b
b
Table 3. Abundance ratios: aProchaska et al. (2001);
bProchaska & Wolfe (1999); cLu, Sargent & Barlow (1998);
dLu et al. (1996); eDessauges-Zavadsky et al. (2001);
f Lopez et al. (1999); gLopez et al. (2002); hBonifacio et al.
(2001); iLevshakov et al. (2002); jGe, Bechtold & Kulkarni
(2001); kPettini et al. (2002); lProchaska, Gawiser & Wolfe
(2001); mCenturion et al. (2000); nProchaska & Wolfe
(1996); oDessauges-Zavadsky, D'Odorico & Prochaska (in
prep.); pThis paper; qPaper I
We now turn to ionization. Given the large N (H i)
of CTQ247A and B,
it is highly unlikely that ioniza-
tion corrections are required (Viegas 1995). Nonetheless,
Prochaska et al (2002c) found significant corrections in a
DLA with log N (H i) = 20.8, so there are occasional cases
where photoionization plays a part even in relatively large
column density systems. In order to constrain the ion-
ization parameter for CTQ247A and B we use CLOUDY
(Ferland 1993) models with a log N (H i) = 21.1, a Haardt-
Madau ionizing spectrum (Haardt & Madau 1996) at z =
2.3 and a metallicity of −1.5 solar, and use the ob-
served ratios to constrain the ionization parameter. For
CTQ247A, R ≡ log N (Al iii)/N (Al ii) provides the most
stringent limit on log U , whereas the ratio of Fe iii/Fe ii
provides most information on the ionization of CTQ247B.
In both cases, the limiting ionization parameter is found to
be log U < −3.2, see upper panel of Fig. 7. Consequently,
this limits the correction to [S/Fe] to be less than 0.07 dex.
Note that a correction this extent is in agreement with the
subsolar [S/Si] observed for the 3 MDLAs in Fig. 6 and
brings this ratio into line with solar, within the measure-
ment errors.
CTQ247C has a significantly lower H i column density.
Our spectrum does cover the Fe iii λ1122 transition but
the limit is not very meaningful. However, we can com-
pare Al ii and Al iii whose ratio has been shown to exhibit
a steady trend with N (H i), a higher fraction of Al iii be-
ing present for lower N (H i) (Vladilo et al. 2001). We find
R = −0.55 which is typical of moderate column density
DLAs, log N (H i) > 20.5. Repeating the above CLOUDY
models with the appropriate H i confirms that the cor-
rections to [S/Fe] will not exceed 0.1 dex in this MDLAs
(bottom panel in Fig. 7). Therefore, we would need to
make a small upward correction to [S/Fe] of up to about
0.1 dex but this does not alter our conclusions.
We note that the MDLAs toward Q2359−02 have
both high R ≈ −0.3, indicating relatively high ionization.
According to our CLOUDY models, however, the correc-
tions for [S/Fe] are still negligible for the corresponding
ionization parameter, while [Si/Fe] requires downward cor-
rections of 0.1 -- 0.2 dex.
From a more general point of view, Fig. 7 tells us that
for a typical log N (H i) = 20.5 DLA the corrections are
downward for Si/Fe and upward for S/Fe, whereas at the
low end of H i column densities the [Si/Fe] ratio is signif-
icantly more sensitive to ionization than [S/Fe] and both
corrections are downward (cf. Prochaska et al. 2002c).
Ionization will therefore require downward corrections in
[Si/S] but these will not necessarily contribute to the wider
spread of MDLA [Si/Fe] values we observe in Fig. 5.
We can quantify the effect of ionization on the compar-
ative dispersion between [Si/Fe] and [S/Fe] by performing
Monte-Carlo simulations with a grid of CLOUDY mod-
els. To this end, we constructed 2 samples of N = 500
CLOUDY models each, with −4 < log U < −3 ('low-
ionization' sample) and −4 < log U < −2 ('high+low-
ionization' sample), and H i column density and metal-
licity values uniformly distributed in the ranges 19.4 <
log N (H i) < 22.0, −2.4 < log(Z/Z⊙) < −0.8 5. To simu-
late our observations we selected randomly 1 000 sets of 20
5 Of course, uniformly distributed values of H i and Z do
not represent the observed distribution of DLAs, this range is
purely for the purpose of investigating the effects of ionization.
Sebastian Lopez and Sara L. Ellison: Abundances in Multiple DLAs
9
Fig. 7. Ratio of doubly to singly ionized species as a
function of ionization parameter U from CLOUDY models
with a Haardt & Madau (Haardt & Madau 1996) ionizing
spectrum. Top panel: Model with log N (H i) = 21.1 and
gas metallicity log Z = log Z⊙ − 1.5 compared with data
points for CTQ247A (open square) and CTQ247B (filled
squares). Middle panel: The measured Al iii/Al ii ratio
for CTQ247C, and a model using log N (H i) = 20.5 and
metallicity log Z = log Z⊙ − 2.5. Bottom panel: A model
using log N (H i) = 20.0, the column density cutoff in our
definition of MDLAs, and metallicity log Z = log Z⊙−1.5.
In all panels the dashed lines indicate solar-corrected ra-
tios while the solid lines are for column density ratios.
DLAs each from the two samples. Their [X/Fe] distribu-
tions are shown in the upper panel of Fig. 8. The difference
between the scatters in [Si/Fe] and [S/Fe] was calculated
as ∆S ≡ σ[Si/Fe] − σ[S/Fe], where σ[X/Fe] is the stan-
dard deviation of each dataset. The bottom panel of the
figure shows the ∆S distribution for both samples. For the
low+high-ionization sample the probability P (∆S > 0)
of having a wider scatter in [Si/Fe] is roughly 0.7, while
P (∆S > 0) = 0.85 if the DLAs are drawn exclusively from
the low-ionization sample. This is an indication that ion-
ization does affect the distributions differentially, although
the differences are generally below measurement errors.
4.3. Other possible systematic effects
Finally, other possible systematic effects we have to be-
ware of are: (1) Column densities: our column densities
come from χ2 fits whereas the majority of literature DLAs
have AODM measurements. However, it is well established
that both approaches give essentially same results pro-
vided the lines are nonsaturated as is our case. (2) Solar
abundances: solar abundances have undergone some sig-
Fig. 8. Simulated [S/Fe] and [Si/Fe] distributions (top
panel) and difference between their scatters (bottom
panel). The solid (dotted) histogram shows DLA simula-
tions drawn from a sample of DLAs with −4 < log U < −3
(−4 < log U < −2 )
nificant revisions in recent years; all the points in figures 5
and 9 have been put on a common solar scale. (3) Atomic
data: some f -values have changed in recent years, most
notably Fe ii λ1611. However, this should not be an issue
as long as very few systems rely on only one transition
6. (4) Blending: S ii transitions are normally in the forest
and therefore particularly susceptible to contamination.
This is especially a worry if the AODM is used since it
would cause larger N (S) if contamination is present. Since
40% of literature S measurements comes from the UCSD
database which uses the AODM, in theory we could have
compared our data with overestimations of S/Fe. However,
checking in Prochaska et al. (2001) on a case-by-case ba-
sis shows that this is unlikely to be a problem because the
possibility of contamination was carefully assessed in each
case. Even excluding [S/Fe] points determined by AODM
6 e.g., Q2206−19A is the only DLA in the sample that relies
solely on Fe ii λ1611
10
Sebastian Lopez and Sara L. Ellison: Abundances in Multiple DLAs
from Fig. 5 still shows the low S trend in MDLAs. (5)
Instrument: almost all literature abundances come from
HIRES data whereas we have used UVES at a slightly
higher resolution. However, some high α/Fe UVES values
in Table 3 and the low α/Fe HIRES value in the MDLAs
toward Q2359−02 show that a possible bias due to differ-
ent data type is unfounded.
In summary, we have found that MDLAs exhibit signif-
icantly lower [S/Fe], and to a less striking extent, [Si/Fe]
ratios than single DLAs taken from the literature. We have
argued that these low values are not driven by ionization
differences or dust depletion, although an observational
bias may be partially responsible for the clearer abun-
dance distinction in S compared to Si. Having excluded
the main systematic effects, these results therefore imply
a nucleosynthetic origin for differences in abundance ratios
between MDLAs and single DLAs.
5. Discussion
5.1. The Nature of MDLAs
[α/Fe] of MDLAs are
If the typically low values of
truly nucleosynthetic in origin, this may indicate that
these systems either represent a particular galaxy type
or that they share a common evolutionary or environmen-
tal link. The first issue to address is whether the 'mul-
tiplicity' is indeed due to more than one distinct galaxy.
Schaye (2001) has suggested that some DLAs may be as-
sociated with winds, in which case MDLAs may be caused
by outflowing material from the parent DLA. Indeed,
Nulsen, Barcons & Fabian (1998) have shown that large
scale galactic winds can produce column densities equal
to those of DLAs. This could be tested by studying abun-
dances in systems that are effectively transverse multiple
systems, that is DLAs observed in more than one line of
sight separated by many tens of kpc. The closest exam-
ple we have of this in the current dataset is the case of
Q0201+1120 (Ellison et al. 2001a) which is located in a
concentration with at least four other galaxies and indeed
shows low [S/Fe], possible evidence that the abundance
trend is symptomatic of a group environment rather than
simply line of sight velocities within a single object.
However, this evidence remains vague in our context
of MDLAs as these imaging studies deal only with DLAs
without line-of-sight companions. Deep imaging and sub-
sequent spectroscopy of the fields toward CTQ247 and
Q2314-409 as well as control fields around QSOs har-
boring single DLAs is therefore required in order to
pursue a comparison of possibly different environments.
At higher redshifts, finding evidence for cluster envi-
ronments of single DLAs (and even the DLA absorbers
themselves) has proven difficult
(Gawiser et al. 2001;
Prochaska et al. 2002a), so much work is still to be done
in this direction. Similarly, although several DLA imaging
studies have been carried out (e.g., Le Brun et al. 1997),
we lack studies of the clustering properties in the DLA
fields at z < 1.0 and there are currently no known MDLAs
at low redshift.
5.2. Truncated star formation in MDLAs?
As suggested in Paper I, the low α/Fe abundances ob-
served in MDLAs might be due to suppressed SF in the
galaxies harboring the absorbing gas. We now raise the
question as to whether environment may cause these ob-
served abundance anomalies and speculate upon the pos-
sibility that MDLAs arise in 'protogalactic groups'.
There is a clear trend of galaxies in rich environments
out to z <∼ 0.5 to exhibit suppressed SF, not just in dense
cluster cores, but also in clusters out to 2-3 virial radii
(Balogh et al. 1997; Lewis et al. 2002) and perhaps out to
several Mpc (Pimbblet et al 2001). The line diagnostics
of these galaxies indicate that radial trends are consis-
tent with an age sequence in which the last episode of
star formation happened most recently in the outermost
galaxies (Balogh et al. 1999). However, the physical mech-
anisms governing this effect are not clear; violent processes
such as ram pressure stripping (e.g., Couch et al. 2001
and references therein) and passive exhaustion of gas sup-
ply (e.g., Balogh et al. 1999) are the two main possibil-
ities. The recent finding by Lewis et al. (2002) that lo-
cal density is the dominant parameter in suppressing SF
rates indicates that extreme cluster-scale processes such
as ram-stripping by the ICM play a minor role. These
results also point to the possibility that reduction in
SF need not take place within the gravitational bounds
of existing virialized clusters and may occur already in
looser groups before they are accreted onto larger struc-
tures. This ties in with cluster histories at lower redshifts,
because although Dressler et al. (1999) find a significant
post-starburst population in clusters, there is evidence
that this fraction is not greater than among the field pop-
ulation (Balogh et al. 1999).
If the same physical processes are on-going at high
redshifts, then suppressed SF may be evident in galaxies
associated with early proto-clusters, such as those identi-
fied by Steidel et al (1998), even though canonical, viri-
alized clusters have yet to form. Although the abundance
ratios for MDLAs that we have discussed here are distinct
from 'field' DLAs, they have typical metallicities for their
redshift (c.f. Pettini et al 1999) and moderate dust con-
tent. These properties are indications that some chemical
maturity has indeed taken place, and MDLAs are simi-
lar to other DLAs in terms of their general enrichment.
However, the low [α/Fe] is an indication that what sets
MDLAs apart is that they have been evolving quiescently,
without any further major SF episodes, over the last ∼
500 Myrs.
5.3. Other evidence for quiescent star formation
If the low α/Fe ratios observed in MDLAs are to be
explained by quiescent SF, then we should expect to
Sebastian Lopez and Sara L. Ellison: Abundances in Multiple DLAs
11
Dessauges-Zavadsky et al 2002). If these high ratios are
not due to dust depletion (Fe is more depleted in the local
ISM than Mn) or ionization, they show a mild odd-even
effect. Analogously, [Al/Si] = 0.07±0.12 in CTQ247C, the
highest value ever measured, and [Al/Si] = −0.28 ± 0.18
in B2314−409B (Paper I), which is among the highest
values after the z = 2.154 MDLA toward Q2359−02
(Prochaska & Wolfe 2002). Altogether, the odd-even ef-
fect seems to be mild in those MDLAs where we can
measure it. Since models of massive star yields give good
agreement with observations of Mn/Fe in Galactic stars
(Goswami & Prantzos 2000; and references therein) and
also DLAs (Ledoux, Bergeron & Petitjean 2002), the lack
of odd-even effect at low metallicity in MDLAs indicates
again that high-mass stars have not been the dominant
metal pollutant in MDLAs. This may be interpreted as
further evidence that there have not been several epochs
of SF building on the enriched interstellar media from pre-
vious episodes; further support of suppressed SF.
Secondary
6. Summary
Fig. 9. The N/α ratio in MDLAs (circles), 'field' DLAs
(stars) and H ii regions (dots; see Pettini et al. 2002 for
references). The triangles are for the 2 DLAs with com-
panion LBGs.
see its signature in other elemental ratios that also
act as 'cosmic clocks'. The most promising possibil-
ity is the N/α ratio which traces the primary and
secondary contributions from stars of different masses
(Henry, Edmunds and Koeppen 2000). Fig. 9 shows the
N/α ratio in MDLAs, 'field' DLAs and H ii regions (Pettini
et al. 2002). We only used measurements of O or S in the
plot and corrected S values by the solar S/O ratio (1.54
dex). Prochaska et al. (2002b) have suggested a bimodal
distribution of N/α and although this needs to be con-
firmed by more data, it is clear that a number of DLAs
exhibit low ratios. For MDLAs, the ratios are all high,
with no points (N/O) < −1.9, supporting the quiescent
SF scenario. It perhaps indicates more chemical maturity
than average, but also that the role of massive stars in
polluting the ISM of MDLAs may be less important.
The other established trend with metallicity in
Galactic stars is the odd-even effect. Although this
is not technically a chronometer in the same sense
as N and the alpha elements, there is a correlation
with metallicity which may be due to metallicity de-
pendent yields (Goswami & Prantzos 2000). At our dis-
posal we have [Mn/Fe] = −0.12 ± 0.08 (CTQ247A)
and [Mn/Fe] = −0.31 ± 0.03 (CTQ247B), the for-
mer of which is among the highest values in DLAs
of this metallicity (Ledoux, Bergeron & Petitjean 2002;
We have studied a sample of 7 DLAs with line-of-sight
companions, 'multiple' DLAs in our nomenclature, and 2
DLAs arising in transverse groups. We have shown that
the relative abundances in these two sub-samples are un-
usual as compared with single DLAs. In particular, the
[S/Fe] and [Si/Fe] ratios are statistically lower than liter-
ature DLAs, although the effect is more striking in [S/Fe].
We suggest that this difference is due to (1) an obser-
vational bias against measuring low S column densities
because weak S ii are usually lost in the forest; and (2),
to a lesser degree, the larger scatter induced in [Si/Fe]
by photoionization. Besides low α/Fe ratios, MDLAs also
exhibit a mild odd-even effect and relatively high N/α
ratios. We interpret these results as truncated star forma-
tion in MDLAs. If the multiplicity of these DLAs is due
to grouping in physical space, a thesis we cautiously sup-
port, environment may be the cause of the quiescent SF
scenario, just as is observed in more nearby clusters and
groups of galaxies.
Acknowledgements. We are grateful to the anonymous ref-
eree for many qualified comments and suggestions on a
first version of this paper. We also thank Eric Gawiser
for fruitful discussions and comments, Michael Rauch for
allowing us
to use the FORS data of CTQ247, Max
Pettini
for providing us with the template for Fig. 9,
Mirka Dessauges-Zavadsky for communicating abundances
in advance of publication, and Jason Prochaska, whose
database at http://kingpin.ucsd.edu/~hiresdla/ we have
consulted. SL acknowledges support from the Chilean Centro
de Astrof´ısica FONDAP No. 15010003, and from FONDECYT
grant No3 000 001.
References
Ballester, P., Modigliani, A., Boitquin, O., Cristiani, S.,
Hanuschik, R., Kaufer, A., Wolf, S., 2000, ESO Messenger,
101, 31
12
Sebastian Lopez and Sara L. Ellison: Abundances in Multiple DLAs
Balogh, M., Morris, S., Yee, H. K. C., Carlberg, R. G.,
Peroux, C. Storrie-Lombardi, L. J., McMahon, R. G., Irwin,
Ellingson, E., 1999, ApJ, 527,54
M. & Hook, I. M., 2001, AJ, 121, 1799
Balogh, M., Morris, S., Yee, H. K. C., Carlberg, R. G.,
Pettini, M., Ellison, S. L., Bergeron, J. & Petitjean, P., 2002,
Ellingson, E., 1997, ApJL, 488, L75
A&A, 391, 21
Bonifacio, P., Caffau, E., Centurion, M., Molaro, P. & Vladilo,
Pettini, M., Ellison, S. L., Steidel, C. C., Bowen, D. V. 1999,
G., 2001, MNRAS, 325, 767
ApJ, 510, 576
Centurion, M., Bonifacio, P., Molaro, P. & Vladilo, G. 2000,
Pimbblet, K, Smail, I., Edge, A., Couch, W., O' Hely, E.,
ApJ, 536, 540
Chen, Y. Q., Nissen, P. E., Zhao, G. & Asplund, M., 2002,
A&A, 390, 225
Chengalur, J. N. & Kanekar, N., 2000, MNRAS, 318, 303
Couch, W., Balogh, M., Bower, R., Smail, I., Glazebrook, K.,
& Taylor, M., 2001, ApJ., 549, 820
Dessauges-Zavadsky, M., Prochaska, J., & D'Odorico, S., 2002,
A&A, 391, 801
Dessauges-Zavadsky, M., D'Odorico, S., McMahon, R. G.,
Molaro, P., Ledoux, C., Peroux, C. & Storrie-Lombardi,
L. J, 2001, A&A, 370, 426
Dressler, A., Smail, I., Poggianti, B., Butcher, H., Couch, W.,
Ellis, R., Oemler, A., 1999, ApJS, 122, 51
Ellison, S. L., & Lopez, S., 2001, A&A, 380, 117, Paper I
Ellison, S. L., Pettini M., Steidel C. C., Shapley A., 2001a,
ApJ, 549, 770
Ellison, S. L., Yan, L., Hook, I., Pettini, M., Wall, J., Shaver,
P., 2001b, A&A, 379, 393
Evrard, A. E., MacFarland, T. J., Couchman, H. M. P.,
Colberg, J. M., Yoshida, N., White, S. D. M., Jenkins, A.,
Frenk, C. S., Pearce, F. R., Peacock, J. A. & Thomas, P.
A., 2002, ApJ, 573, 7
Ferland, G. J. 1993, University of Kentucky, Physics
Department Internal Report
Gawiser, E., Wolfe, A. M., Prochaska, J. X., Lanzetta, K. M.,
Zabludoff, A., 2001, MNRAS, 327, 588
Prochaska, J X., Gawiser, E., Wolfe, A. M., Quirrenbach, A.,
Lanzetta, K. M., Chen, H.-W, Cooke, J., & Yahata, N.
2002a, AJ, 123, 2206
Prochaska, J. X., Henry, R., O'Meara, J., Tytler, D., Wolfe, A.,
Kirkman, D., Lubin, D., Suzuki, N., 2002b, ApJ, accepted
Prochaska, J. X., Howk, J. C., O'Meara, J. M., Tytler, D.,
Wolfe, A., Kirkman, D., Lubin, D. & Suzuki, N. 2002c,
ApJ, 571, 693
Prochaska, J. X., & Wolfe, A., 2002, ApJ, 566, 68
Prochaska, J. X., Wolfe, A. M., Tytler, D., Burles, S., Cooke,
J., Gawiser, E., Kirkman, D., O'Meara, J. M., & Storrie-
Lombardi, L. ApJS, 2001, 137, 21
Prochaska, J. X., Gawiser, E. & Wolfe, A. M., 2001, ApJ, 552,
99
Prochaska, J.X., & Wolfe, A.M. 1999, ApJS, 121, 369
Prochaska, J.X., & Wolfe, A.M. 1996, ApJ, 470, 403
Quashnock, J. M., vanden Berk, D. E. & York, D. G., 1996,
ApJ 472, 69
Rigopoulou, D., Franceschini, A., Aussel, H., Genzel, R.,
Thatte, H., Cesarsky, C., 2002, ApJ, 580, 789
Savage, B. D. & Sembach, K. R., 1991, ApJ 379, 245
Savage, B. D. & Sembach, K. R., 1996, ARA&A, 34, 279
Schaye, J., 2001, ApJL, 559, L1
Steidel, C.C., Adelberger, K.L., Dickinson, M., Giavalisco, M.,
Yahata, N., & Quirrenbach, A. 2001, ApJ, 562, 628
Pettini, M. & Kellogg, M. 1998, ApJ, 492, 428.
Goswami, A., & Prantzos, N., 2000, A&A, 359, 191
Grevesse, N., & Sauval, A.J. 1998, Space Sci Rev, 85, 161
Haardt, F., & Madau, P. 1996, ApJ, 461, 20
Haines, C., Campusano, L. E. & Clowes, R., 2002,
Steidel, C., Giavalisco, M., Pettini, M., Dickinson, M.,
Adelburger, K., 1996, ApJL, 462, L17
Viegas, S. M., 1995 MNRAS, 276, 268
Vladilo, G., Centuri´on, M., Bonifacio, P., & Howk, J. C., 2001,
astro-ph/0301473
ApJ 557, 1007
Henry, R. B. C., Edmunds, M. G., & Koppen, J. 2000, ApJ,
541, 660
Vladilo, G., 2002, ApJ, 569, 295
Williger, G. M., Campusano, L. E., Clowes, R. G. & Graham,
Holweger, H., 2001, in Solar and Galactic Composition, ed. R.
M. J., 2002, ApJ, 578, 708
Wimmer-Schweingruber, (Berlin: Springer), 23
Hou, J. L., Boissier, S. & Prantzos, N. 2001, A&A, 370, 23
Le Brun, V., Bergeron, J., Boisse, P. & Deharveng, J. M., 1997,
Wolfe, A. M., Turnshek, D. A., Smith, H. E. & Cohen, R. D.
1986, ApJS 61, 249
Zaritsky, D., Smith, R., Frenk, C., White, S. D. M., 1997 ApJ,
478, 39
A&A, 321, 733
Ledoux, C., Bergeron J., & Petitjean, P., 2002, A&A, 385, 802
Levshakov, S. A., Dessauges-Zavadsky, M., D'Odorico, S. &
Molaro, P., 2002, ApJ, 565, 696
Lewis I., Balogh, M., De Propris, R., et al. 2002, MNRAS, 334,
673
Lopez S., Reimers D., Rauch M., Sargent W. L. W. & Smette
A., 1999, ApJ 513, 598
Lopez, S., Maza, J., Masegosa, J., & Marquez, I., 2001, A&A
366, 387
Lopez, S., Reimers, D. D'Odorico, S. & Prochaska, J. X., 2002,
A&A 385, 778
Lu, L., Sargent, W. L. W., Barlow, T.A. 1998, AJ, 115, 55
Lu, L., Sargent, W. L. W., Barlow, T. A., Churchill, C. W.,
Vogt, S. 1996, ApJS 107, 475
Molaro, P., Levshakov, S. A., D'Odorico, S., Bonifacio, P., &
Centurion, M., 2001, ApJ 549, 90
Nulsen, P. E. J., Barcons, X., Fabian, A. C., 1998, MNRAS,
301, 168
|
astro-ph/0307420 | 1 | 0307 | 2003-07-24T11:06:41 | On characterising the variability properties of X-ray light curves from active galaxies | [
"astro-ph"
] | We review some practical aspects of measuring the amplitude of variability in `red noise' light curves typical of those from Active Galactic Nuclei (AGN). The quantities commonly used to estimate the variability amplitude in AGN light curves, such as the fractional rms variability amplitude, F_var, and excess variance, sigma_XS^2, are examined. Their statistical properties, relationship to the power spectrum, and uses for investigating the nature of the variability processes are discussed. We demonstrate that sigma_XS^2 (or similarly F_var) shows large changes from one part of the light curve to the next, even when the variability is produced by a stationary process. This limits the usefulness of these estimators for quantifying differences in variability amplitude between different sources or from epoch to epoch in one source. Some examples of the expected scatter in the variance are tabulated for various typical power spectral shapes, based on Monte Carlo simulations. The excess variance can be useful for comparing the variability amplitudes of light curves in different energy bands from the same observation. Monte Carlo simulations are used to derive a description of the uncertainty in the amplitude expected between different energy bands (due to measurement errors). Finally, these estimators are used to demonstrate some variability properties of the bright Seyfert 1 galaxy Markarian 766. The source is found to show a strong, linear correlation between rms amplitude and flux, and to show significant spectral variability. | astro-ph | astro-ph | Mon. Not. R. Astron. Soc. 000, 1–14 (2003)
Printed 19 November 2011
(MN L ATEX style file v2.2)
On characterising the variability properties of X-ray light curves
from active galaxies
S. Vaughan,1,2 R. Edelson,3 R. S. Warwick2 and P. Uttley4
1 Institute of Astronomy, Madingley Road, Cambridge CB3 0HA
2X-Ray and Observational Astronomy Group, Department of Physics and Astronomy, University of Leicester, Leicester LE1 7RH
3Astronomy Department, University of California, Los Angeles, CA 90095-1562; USA
4Department of Physics and Astronomy, University of Southampton, Southampton SO17 1BJ
Accepted: 24/7/2003; Submitted: 23/7/2003; in original form: 3/3/2003
ABSTRACT
We review some practical aspects of measuring the amplitude of variability in ‘red noise’ light
curves typical of those from Active Galactic Nuclei (AGN). The quantities commonly used to
estimate the variability amplitude in AGN light curves, such as the fractional rms variability
amplitude, Fvar , and excess variance, σ2
XS , are examined. Their statistical properties, relation-
ship to the power spectrum, and uses for investigating the nature of the variability processes
are discussed. We demonstrate that σ2
XS (or similarly Fvar ) shows large changes from one part
of the light curve to the next, even when the variability is produced by a stationary process.
This limits the usefulness of these estimators for quantifying differences in variability ampli-
tude between different sources or from epoch to epoch in one source. Some examples of the
expected scatter in the variance are tabulated for various typical power spectral shapes, based
on Monte Carlo simulations. The excess variance can be useful for comparing the variability
amplitudes of light curves in different energy bands from the same observation. Monte Carlo
simulations are used to derive a description of the uncertainty in the amplitude expected be-
tween different energy bands (due to measurement errors). Finally, these estimators are used
to demonstrate some variability properties of the bright Seyfert 1 galaxy Markarian 766. The
source is found to show a strong, linear correlation between rms amplitude and flux, and to
show significant spectral variability.
Key words: galaxies: active — galaxies: Seyfert — galaxies: individua
galaxies — methods: data analysis
l (Mrk 766) — X-rays:
3
0
0
2
l
u
J
4
2
1
v
0
2
4
7
0
3
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
1 INTRODUCTION
One of the defining characteristics of Active Galactic Nucle i
(AGN) is that their X-ray emission is variable. X-ray light curves
from Seyfert 1 galaxies show unpredictable and seemingly aperi-
odic variability (Lawrence et al. 1987; McHardy 1989). Such ran-
dom variability is often referred to as noise, meaning that it is the
result of a stochastic, as opposed to deterministic, process. In this
context the ‘noise’ is intrinsic to the source and not a result of mea-
surement errors (such as Poisson noise), i.e. the signal itself is the
output of a noise process.
One of the most common tools for examining AGN variability
(and noise processes in general) is the fluctuation Power Spe ctral
Density (PSD) which represents the amount of variability power
(mean of the squared amplitude) as a function of temporal fre-
quency (timescale−1 ). The high frequency PSDs of Seyferts are
usually well-represented by power-laws over a broad range of fre-
quencies (P (f ) ∝ f −α , where P (f ) is the power at frequency f )
with slopes α = 1 − 2 (Green, McHardy & Lehto, 1993; Lawrence
& Papadakis, 1993; Edelson & Nandra, 1999; Uttley, McHardy &
Papadakis, 2002; Vaughan, Fabian & Nandra 2003; Markowitz et
al. 2003). Such a spectrum, with a slope α ∼> 1 is usually called
‘red noise’ (for an introduction to red noise see Press 1978).
If Seyfert 1 light curves are viewed as the product of a stochas-
tic (in this case red noise) process then the speci fic details of each
individual light curve provide little physical insight. Each light
curve is only one realisation of the underlying stochastic process,
i.e. it is one of the ensemble of random light curves that might be
generated by the process. Each new realisation will look different
and these changes are simply statistical fluctuations inher ent in any
stochastic process (as opposed to changes in the nature of the pro-
cess itself). Therefore one should expect two light curves to have
different characteristics (such as mean and variance) even if they
are realisations of the same process. On the other hand, data from
deterministic processes, for example the energy spectrum of a non-
varying source or the light curve of a strictly periodic source (such
as a pulsar), should be repeatable within the limits set by the mea-
surement errors.
It is the average properties of the variability (such as the PSD)
that often provide most insight into the driving process. For exam-
ple, the PSD of any real red noise process cannot continue as a
2
Vaughan et al.
steep power-law indefinitely to longer timescales or the int egrated
variability power would diverge. Therefore the PSDs of AGN vari-
ability must break to a flatter index at low frequencies; the p o-
sition of such a break would represent a characteristic variability
timescale and may yield information about the underlying driving
process. Recent timing studies have indeed found evidence that the
steep power-law PSDs of Seyfert 1s show a flattening, or turno ver,
at low frequencies (Edelson & Nandra 1999; Uttley et al. 2002;
Markowitz et al. 2003).
In many cases however the data are not adequate for PSD anal-
ysis. In these situations the variability is usually described in terms
of the statistical moments (e.g. the sample mean and variance, etc.).
However, due to the stochastic nature of red noise variability there
is a large degree of randomness associated with these quantities. In
practice this means that it is difficult to assign meaningful errors to
the variance. This in turn makes it difficult to quantitative ly com-
pare variances, say from repeated observations of the same source
(and thereby test whether the variability is stationary). Such an
analysis might be desirable; it could in principle reveal changes
in the ‘state’ of the source if its variability properties were found
to evolve with time. This paper discusses this and related problems
that are encountered when examining the variability properties of
AGN. Particular emphasis is placed on the mathematical properties
and implications of the inherent randomness in the variability. The
mathematical details are well understood from the general theory
of stochastic processes (e.g. Priestley 1981 for spectral analysis)
but some of the practical consequences for AGN observations have
not been discussed in detail. On the basis of simulated data some
recipes are developed that may serve as a useful guide for observers
wishing make quantitative use of their variability analysis without
the recourse to e.g. extensive Monte Carlo simulations.
The paper is organised as follows. Section 2 defines the esti-
mators to be discussed (namely the periodogram and the variance).
Simulated data are used to illustrate various aspects of these es-
timators; section 3 describes methods for producing arti fic ial red
noise time series. Section 4 discusses the stationarity of time se-
ries. Sections 5 and 6 discuss two sources of uncertainty associated
with measuring variability amplitudes, the first due to the s tochas-
tic nature of the variability and the second due to flux measur ement
errors. Section 7 gives an example using a real XMM-Newton ob-
servation of Mrk 766. Finally, a brief discussion of these results is
given in section 8 and the conclusions are summarised in section 9.
2 ESTIMATING THE POWER SPECTRAL DENSITY
The PSD defines the amount of variability ‘power’ as a functio n of
temporal frequency. It is estimated by calculating the periodogram1
(Priestley 1981; Bloomfield 2000).
For an evenly sampled light curve (with a sampling period
∆T ) the periodogram is the modulus-squared of the Discrete
Fourier Transform (DFT) of the data (Press et al. 1996). For a light
curve comprising a series of fluxes xi measured at discrete times
ti (i = 1, 2, . . . , N ):
xi e2π ifj ti (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
DF T (fj )2 = (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
N
Xi=1
1 Following Priestley (1981) the term “periodogram ” is used f or the dis-
crete function P (fj ), which is an estimator of the continuous PSD P (f ).
The periodogram is therefore specific to each realisation of
the process,
whereas the PSD is representative of the true, underlying process.
=
2
,
(1)
xi cos(2πfj ti ))2
xi sin(2πfj ti ))2
( N
+ ( N
Xi=1
Xi=1
at N/2 evenly spaced frequencies fj = j/N ∆T (where j =
1, 2, . . . , N/2), fN/2 = 1/2∆T is the Nyquist frequency, fNyq .
Note that it is customary to subtract the mean flux from the
light curve before calculating the DFT. This eliminates the zero-
frequency power. The periodogram, P (fj ), is then calculated by
choosing an appropriate normalisation A (see Appendix A for more
on periodogram normalisations). For example
2∆T
P (fj ) = ADF T (fj )2 =
N DF T (fj )2 .
If the time series is a photon counting signal such as normally
encountered in X-ray astronomy, and is binned into intervals of
∆T , the effect of Poisson noise is to add an approximately con-
stant amount of power to the periodogram at all frequencies. With
the above normalisation this constant Poisson noise level is 2 ¯x (as-
suming the light curve is not background subtracted).
(2)
2.1 Statistical properties of the periodogram
(3)
The periodogram of a noise process, if measured from a single time
series, shows a great deal of scatter around the underlying PSD. In
particular, the periodogram at a given frequency [P (f )] is scat-
tered around the PSD [P (f )] following a χ2 distribution with two
degrees of freedom (van der Klis 1989):
P (f ) = P (f )χ2
2/2,
where χ2
2 is a random variable distributed as χ2 with two degrees of
freedom, i.e. an exponential distribution with a mean and variance
of two and four, respectively. The periodogram is distributed in this
way because the real and imaginary parts of the DFT are normally
distributed for a stochastic process2 (section 6.2 of Priestley, 1981;
Jenkins & Watts 1968). The expectation value of the periodogram
is equal to the PSD but its standard deviation is 100 per cent, lead-
ing to the larger scatter in the periodogram (see Fig. 1). See Leahy
et al. (1983), van der Klis (1989), Papadakis & Lawrence (1993),
Timmer & K onig (1995) and Stella et al. (1997) for further discus-
sion of this point.
When applied to real data the periodogram is an inconsistent
estimator of the PSD, meaning that the scatter in the periodogram
does not decrease as the number of data points in the light curve
increases (Jenkins & Watts 1968). In order to reduce this scatter
the periodogram must be smoothed (averaged) in some fashion. As
the number of data points per bin increases (either by binning over
frequencies or averaging over many data segments) the scatter in
the binned periodogram decreases, i.e. the averaged periodogram
is a consistent estimator of the PSD (see Papadakis & Lawrence
1993 and van der Klis 1997 for more on binned periodogram esti-
mates). A further point is that periodograms measured from fi nite
data tend to be biased by windowing effects which further compli-
cate their interpretation (van der Klis 1989; Papadakis & Lawrence
1993; Uttley et al. 2002 and see below).
2 The DFT at the Nyquist frequency is always real when N is even so the
periodogram at this frequency is distributed as χ2
1 , i.e. with one degree of
freedom.
Properties of X-ray light curves from AGN
3
3 SIMULATING RED NOISE LIGHT CURVES
3.1 Algorithms
In order to elucidate the properties of the variance of red noise data,
random light curves were generated from power-law PSDs similar
to those of AGN. Fig. 1 shows two arti ficial time series and the ir
periodograms. It is worth reiterating that the large scatter in the
periodograms is an intrinsic property of stochastic processes – it
does not depend on the number of data points and is not related to
Poisson noise in the data.
These arti ficial time series were produced using the algorit hm
of Timmer & K onig (1995). This generates random time serieswith
arbitrary broad-band PSD, correctly accounting for the intrinsic
scatter in the powers (i.e. equation 3). Other methods of generating
random light curves include the related ‘summing of sines’ method
(Done et al. 1992). Note that it is not correct to randomise only the
phases of the component sine functions, their amplitudes must also
be randomised. Otherwise this method does not account for this
intrinsic scatter in the powers. Shot-noise models can produce red
noise time series with certain PSD shapes (see Lehto 1989). There
also exist various mathematical tricks for producing data with spe-
ci fic power-law PSD slopes. Data with a α = 1 PSD (often called
‘ flicker noise’) can be generated using the half-integral me thod out-
lined in Press (1978), while α = 2 (‘random walk’) data can be
generated using a first-order autoregressive process (AR[1 ]), essen-
tially a running sum of Gaussian deviates (see Deeming 1970 and
Scargle 1981 for more on such methods). The method of Timmer
& K onig (1995) is used below as this can generate time series from
an arbitrary PSD and is computationally efficient.
3.2 Simulating ‘realistic’ data
Some caution should be applied when using these routines to pro-
duce arti ficial time series. As mentioned briefly in the previ
ous sec-
tion, periodograms measured from real data tend to be biased by
windowing effects. For uninterrupted but finite observatio ns data
of red noise processes the most important of these effects is ‘red
noise leak’ – the transfer of power from low to high frequenci es
by the lobes of the window function (see e.g. Deeter & Boynton
1982; van der Klis 1997). If there is signi ficant power at freq uen-
cies below the lowest frequency probed by the periodogram (i.e. on
timescales longer than the length of the observation) this can give
rise to slow rising or falling trends across the light curve. These
trends contribute to the variance of the light curve. Thus variability
power ‘leaks’ into the data from frequencies below the observed
frequency band-pass. The degree to which this occurs, and the re-
sultant bias on the measured periodogram, depend on the shape of
the underlying PSD and the length of the observation (Papadakis &
Lawrence 1995; Uttley et al. 2002). For flat PSD slopes ( α < 1.5)
the amount of leakage from low frequencies is usually negligible.
Since AGN light curves usually contain signi ficant power on
timescales longer than those probed (see section 4.1) the effects
of red noise leak must be included in simulations of AGN light
curves. This can be achieved by using the Timmer & K onig (1995)
algorithm to generate a long light curve from a PSD that extends to
very low frequencies and then using a light curve segment of the re-
quired length. Data simulated in such a fashion will include power
on timescales much longer than covered in the short segment. The
effects of measurement errors (e.g. Poisson noise) can be included
in the simulations using standard techniques (e.g. Press et al. 1996).
Figure 1. Simulated time series (left) and their periodograms (right). The
upper panel shows a ‘flicker noise’ time series which has a f −1 PSD. The
lower panel shows a ‘random walk’ time series with a f −2 PSD. Note the
large scatter in the periodogram (dots) around the underlying PSD (solid
line). It is clear that the time series with the steeper PSD shows more power
in long-term variability while the time series with the flatt er PSD shows
relatively more power in short term variability (flickering ). The two series
were generated using the same random number sequence.
2.2 Integrated power
The integral of the PSD between two frequencies (f1 and f2 ) yields
the contribution to the expectation value of the (‘true’) variance
due to variations between the corresponding timescales (1/f1 and
1/f2 ). This result follows from Parseval’s theorem (see e.g. van der
Klis 1989)
hS 2 i = Z f2
f1
Correspondingly, for a discrete time series the integrated peri-
odogram yields the observed variance for that particular realisation
P (f )df .
(4)
S 2 =
P (fj )∆f ,
N/2
Xj=1
where ∆f is the frequency resolution of the DFT (∆f =
1/N ∆T ). The total variance of a real light curve is equal to its
periodogram integrated over the frequency range f1 = 1/N ∆T to
fNyq = 1/2∆T .
The sample variance (which will differ from observation to
observation) is given by:
(5)
S 2 =
N
1
Xi=1
(xi − ¯x)2 ,
N − 1
where ¯x is the arithmetic mean of xi . In the limit of large N these
two variance estimates are identical. The normalised variance3 is
simply S 2 / ¯x2 .
(6)
3 In AGN studies normalised quantities are often used in preference to ab-
solute quantities as they are independent of the flux of a spec ific source.
This means that, in principle, normalised amplitudes can be used to com-
pare sources with different fluxes.
4
Vaughan et al.
4 STATIONARITY
A stationary process is one for which the statistical properties (such
as mean, variance, etc.) do not depend on time. Fig. 2 shows an ar-
ti ficial red noise time series together with its mean and vari ance
measured every 20 data points. The simulation was produced from
a single, well-defined PSD (which did not vary). The process i s
therefore stationary. It would have been reasonable to expect the
resulting time series (a realisation of the process) to appear sta-
tionary. This is not the case however; both the mean and variance
change with time (panels 2 and 3). This has nothing whatsoever to
do with measurement errors - the simulation has zero errors. This
simulation demonstrates that, when dealing with red noise, fluctu-
ations in variance are not sufficient to claim the variabilit y process
is non-stationary.
As the purpose of time series analysis is to gain insight into
the process, not the details of any speci fic realisation, a mo re robust
approach is needed to determine whether these data were produced
by a time-stationary process or a non-stationary process. It would
be more insightful to consider whether the expectation values of
the characteristics (such as the variance) are time-variable. The ex-
pectation values should be representative of the properties of the
underlying process, not just any one realisation. See section 1.3 of
Bendat & Piersol (1986) for a discussion of this point.
4.1 Weak non-stationarity
For a process with a steep red noise PSD (α > 1), the integrated
periodogram will diverge as f → 0. This means that (following
equation 4) the variance of a red noise time series with a steep PSD
will diverge with time. In this case there is no well-defined m ean;
Press & Rybicki (1997) describe this form of variability as ‘weakly
non-stationary.’ For time series with power spectra flatter
than this
the variance converges as f → 0, thus for a white noise process
with a flat PSD ( α = 0), the variance will converge as the obser-
vation length increases, and there will be a well-defined mea n on
long timescales.
Of course, for any real process the PSD must eventually flat-
ten such that the power does not diverge (i.e. α < 1 on sufficiently
long timescales). Thus, weak non-stationarity is entirely due to ob-
servations sampling only the steep part of the PSD of a source.
But in AGN this flattening occurs on timescales much longer th an
those probed by typical observations. For instance, XMM-Newton
observations of AGN typically last for ∼ few × 104 s whereas in
many objects the PSD is steep until > 105 s and in some cases
probably much longer (Edelson & Nandra 1999; Uttley et al. 2002;
Markowitz et al. 2003). Therefore on the timescales relevant for
most X-ray observations, AGN light curves should be considered
weakly non-stationary.
4.2 Stochasticity
Fluctuations in the statistical moments (such as mean and variance)
of a light curve are intrinsic to red noise processes. Therefore, even
in the absence of measurement errors (e.g. no Poisson noise) the
means and variances of two light curves produced by exactly the
same process can be signi ficantly different. This can be seen in
Fig. 2 (panels 2 and 3), where each 20 point segment of the light
curve shows a different mean and variance. These random fluct u-
ations in variance are however governed by the normal statistical
rules of noise processes and can thus be understood in a statistical
sense.
Figure 2. Panel 1: Simulated red noise time series (with a f −2 PSD) with
N = 2800 points. Panel 2 and 3: mean and variance measured from seg-
ments of 20 points (calculated using equation 6). The variances follow a
distribution of the form shown in the bottom panel of Fig. 3. (Note the vari-
ance is plotted on a logarithmic scale.) Panel 4: averaged variance measured
by binning the individual variances into groups of 20 consecutive estimates.
The errors are the standard error on the mean (equation 4.14 of Bevington
& Robinson 1992). These averaged variances are consistent with staying
constant. In other words, although the instantaneous value of the variance
fluctuates, its expectation value is consistent with consta nt (i.e. a station-
ary process). Panel 5: fractional rms amplitude (pS 2 / ¯x2 ) measured from
segments of 20 points. Panel 6: averaged fractional rms amplitude measured
by binning the individual amplitudes into groups of 20. The fractional am-
plitude is anti-correlated with the light curve because hS 2 i is constant but
Fvar is normalised by the light curve flux.
Any given series is only one realisation of the processes and
its periodogram will show the scatter predicted by equation 3. The
integrated periodogram (which gives the variance; equation 5) will
therefore be randomly scattered around the true value for the PSD
of the process. The variance in a speci fic time series is given by
S 2 =
1
N ∆T
P (fi )χ2
2 /2,
N/2
Xi=1
i.e. the variance of a given realisation is a sum of χ2
2 distributions
weighted by the PSD4 . (This assumes biases such as red noise
leak are not signi ficant. If this is not true then these biases will
(7)
4 As noted earlier, the periodogram at the Nyquist frequency is actually
Properties of X-ray light curves from AGN
5
These points are illustrated by Fig. 3, which shows the distri-
butions of variances in random time series with three different PSD
slopes (α = 0, 1, 2). This plot was produced by generating 50,000
random time series (each 100 points long) for each PSD slope and
measuring the variance of each one. The scatter in the variance is
entirely due to random fluctuations between different reali sations
because the PSD normalisation was kept fixed and no instrumen tal
noise was added. The shape of the distribution of variances can be
seen to depend on the PSD slope.
Consider a white noise process (α = 0, i.e. P (f ) = const).
The periodogram of each realisation is randomly scattered around
its flat PSD. Following equation 7 the variance is simply the s um
2 -distributed powers in the periodogram, and these
of the N/2 χ2
2 distri-
are evenly weighted (PSD is constant). The sum of N/2 χ2
butions follows a χ2
N distribution. This tends to a normal distribu-
tion as N increases. Thus the variance of a white noise process is
approximately normally distributed (as can be seen in Fig. 3) and
converges as N increases. The fractional standard deviation of χ2
N
is given by p2/N , so for the 100 point light curves used the (1σ )
fractional width of the variance distribution is ≈ 14.1 per cent, in
agreement with the simulations (Fig. 3, top panel).
For time series with steeper PSDs the lower frequency peri-
odogram points contribute more strongly to the sum than the higher
frequency points. The variance of such a time series is therefore
dominated by a few low frequency powers5 and thus resembles a
ν distribution with low ‘effective degrees of freedom.’ The distri-
χ2
bution of variances in red noise data is dependent on the underlying
PSD and is, in general, non-Gaussian (Fig. 3). The fractional stan-
ν is p2/ν , which tends to unity as the PSD
dard deviation of χ2
gets steeper (i.e. as effective ν → 2). Thus the largest fluctuations
in variance (up to a limit of ∼ 100 per cent rms) are expected to
result from very steep PSD slopes.
5 INTRINSIC SCATTER IN VARIANCE
As discussed above, when examining AGN light curves one should
expect random changes in the mean and variance with time (be-
tween segments of a long observation or between observations
taken at different epochs). This is true even if the measurement
errors are zero and does not depend on the number of data points
used (due to the weak non-stationarity). However, it is also possi-
ble that the underlying process responsible for the variability itself
changes with time (e.g. the PSD changes), in which case the vari-
ability is non-stationary in a more meaningful sense – ‘stro ngly
non-stationary.’ Such changes in the variability process could pro-
vide insight into the changing physical conditions in the nuclear
regions. On the other hand the random changes expected for a red
noise process yield no such physical insight. The question thus
arises: how does one tell, from a set of time series of the same
source, whether they were produced by a strongly non-stationary
process? In other words, is it possible to differentiate between dif-
ferences in variance caused by real changes in the variability pro-
cess (physical changes in the system), and random fluctuatio ns ex-
pected from red noise (random nature of the process)?
If the process responsible for the variability observed in a
Figure 3. Distribution of variances in time series with three different PSD
shapes: f 0 (top), f −1 (middle) and f −2 (bottom). Each distribution is de-
rived from 50,000 realisations. As the PSD gets steeper the distribution of
variances becomes less Gaussian and more like a χ2 distribution with a low
effective degrees of freedom.
further distort the distribution of variances.) It is the expectation
value of the variance that is representative of the integrated power
in the PSD, and thus the average amplitude of the variability pro-
cess (eqn. 4). Thus, while the expectation value of the variance is
equal to the integrated PSD, each realisation (time series) of the
same process will show a different variance even if the parent vari-
ability process is stationary. This is particularly important for steep
PSD time series (i.e. weakly non-stationary data) since the vari-
ance does not converge as more data are collected. Only if the time
series spans timescales on which the integrated power converges
at low frequencies (i.e. α < 1) will the variance converge as the
length of time series increases.
distributed as χ2
1 for even N . However, for large N this will make a negli-
gible difference to the sum (cf. equation 2.9 of van der Klis, 1989).
5 The windowing effects mentioned above mean that fluctuation s at powers
above and below the frequency range of the periodogram may also affect the
variance.
6
Vaughan et al.
given source is stationary then its PSD is constant in time. The ex-
pectation value of the absolute (un-normalised) variance will there-
fore be the same from epoch to epoch, but the individual variance
estimates will fluctuate as discussed in section 4.2. This ma kes it
difficult to judge, from just the variances of two light curve s taken
at two epochs, whether they were produced by a stationary process.
Given sufficient data it is, however, possible to test whethe r the ex-
pectation values of the variance (estimated from an ensemble of
light curves) at two epochs are consistent with a stationary process.
5.1 Comparing PSDs
The methods most frequently employed involve comparing the
PSDs (estimated from the binned periodogram) at different epochs.
If the PSDs show signi ficant differences (at a given confidenc
e
level) the variability process can be said to be strongly non-
stationary. As an example of this, the PSDs of X-ray binaries
evolve with time, and the way in which the variability properties
evolve provides a great deal of information on the detailed work-
ings of these systems (see e.g. Belloni & Hasinger 1990; Uttley &
McHardy 2001; Belloni, Psaltis & van der Klis 2002; Pottschmidt
et al. 2002).
Papadakis & Lawrence (1995) suggested a method suitable for
testing whether large AGN datasets display evidence for strongly
non-stationary variability. Again this method works by comparing
the PSDs from different time intervals, in this case by determining
whether the differences between two periodograms are consistent
with the scatter expected based on equation 3. In particular, they de-
fine a statistic s based on the ratio of two normalised periodograms.
If s deviates signi ficantly from its expected value for stationa ry data
(if hsi = 0) then the hypothesis that the data are stationary can be
rejected (at some confidence level).
5.2 Comparing variances
A different approach is compare variances Si derived from M ob-
servations of the same source (either segments of one long obser-
vation or separate short observations). In order to test whether the
Si differ signi ficantly (i.e. more than expected for a red noise pro-
cess) a measure of the expected scatter is required. This error could
be obtained directly from the data (by measuring the standard de-
viation of multiple estimates) or through simulations (based on an
assumed PSD shape)6 .
5.2.1 Empirical error on variance
An empirical estimate of the mean and standard deviation of the
variance can be made given M non-overlapping data segments. The
M segments each yield an estimate of the variance7 , Si . Each of
these is an independent variable of (in general) unknown but identi-
cal distribution (unless the process is strongly non-stationary). The
central limit theorem dictates that the sum of these will become
6 It is assumed that the data segments being compared have identical sam-
pling (same bin size and observation length). Every effort should be made
to ensure this is the case, e.g. by clipping the segments to the same length.
The variances will then be calculated over the same range of timescale (fre-
quencies). As the variance can increase rapidly with timescale in red noise
data this is most important for steep PSD data such as AGN light curves.
7 Ideally each segment should contain at least N ∼> 20 data points in order
to yield a meaningful variance.
Figure 4. The average rms amplitude (σ = √S 2 ) as a function of flux for
the simulated light curve shown in Fig. 2. The individual rms estimates were
sorted by flux and binned to M = 20 estimates per bin. Errors correspond
to the error on the mean value. The amplitude is constant with flux.
normally distributed as M increases. Therefore by averaging the
M variance estimates it is possible to produce an averaged vari-
ance (hS 2 i) and assign an error bar in the usual fashion (e.g. equa-
tion 4.14 of Bevington & Robinson 1992). This gives a robust esti-
mate of the variance and the standard deviation of the M variances
around the mean gives an estimate of the uncertainty on the mean
variance.
If several sets of data segments are acquired it is therefore pos-
sible to compare the mean variance of each set statistically (since
each has an associated uncertainty). For example, with two long
XMM-Newton observations of the same source, taken a year apart,
one could measure the variance for each observation (by breaking
each into short segments and taking the mean variance of the seg-
ments). Thus it would be possible to test whether the variability
was stationary between the two observations. This method of esti-
mating the mean and standard deviation of the variance requires a
large amount of data. Of order N × M = 20 × 20 = 400 data
points are needed to produce a single well-determined estimate of
the mean variance and its error. A typical XMM-Newton observa-
tion of a bright Seyfert 1 galaxy (∼ 40 ks duration) is only likely
to yield enough data for one estimate of the mean variance. Thus
this method is suitable for testing whether the mean variance has
changed from observation to observation.
Fig. 2 (panel 4) demonstrates this empirically derived mean
variance and its error bar on a long, simulated time series. These
data were produced by calculating the variances Si in bins of
N = 20 data points (panel 3) and then averaging M = 20 vari-
ances to produce a mean variance with error bar (panel 4). These
averaged variances are consistent with constant, as expected; fit-
ting these data with a constant gave χ2
ν = 0.84. Figure 4 shows
the rms amplitude is constant with flux. These tests indicate
that
the integrated PSD is consistent with being constant with time; the
variance does not change signi ficantly from epoch to epoch (o r as
function of flux), as expected for a stationary process.
5.2.2 Estimating the error on the variance through simulations
The advantage of the above method is that it requires no assump-
tion about the shape of the PSD, The drawback is that it requires a
substantial amount of data to produce a single, robust variance esti-
mate. An alternative approach is to estimate the standard deviation
of the variances Si based on simulations.
Given an assumed shape for the PSD it is possible to calcu-
late the distribution of variances expected for a stationary process
(see section 4.2). Some example distributions are shown in Fig. 3,
which clearly demonstrates how the distribution depends on the
slope of the PSD. The distribution becomes more normal at flat -
ter slopes and more asymmetric at steep slopes. For a given PSD
shape these distributions are well defined (by eqn. 7) and can be
computed through Monte Carlo simulations. This makes it possible
to estimate limits within which one would expect the variance to be
distributed if the process is stationary.
The two primary factors that affect the distribution of variance
are the PSD shape and the sampling of the light curve (the length of
the data segments in the case of contiguously binned light curves).
Table 1 gives the expected confidence limits for four differe nt PSD
shapes and five different lengths for the data segments. Thes e val-
ues were computed by simulating one very long light curve with
the assumed PSD shape and breaking it into 1000 separate seg-
ments (of speci fied length). The variance within each segmen t was
measured and the distribution of the 1000 variances was calculated.
The 90 per cent confidence interval was calculated by finding the
5th and 95th percentiles of the variance distribution (in general
these upper and lower bounds will differ because the distribution
is asymmetric). The numbers given in the table are the boundaries
of the 90 and 99 per cent confidence regions estimated by aver-
aging the results from 50 runs. The limits are given in terms of
±∆ log(S 2 ) because they are multiplicative. That is, from a partic-
ular realisation the variance is expected to be scattered within some
factor of the true variance (for which the absolute normalisation
is irrelevant). The factors are tabulated in terms of their logarithms
(since multiplicative factors in linear-space become additive offsets
in log-space).
The PSD used for the simulations was chosen to match that ex-
pected for AGN, i.e. a steep power-law at high frequencies (with a
slope of α = 1.0, 1.5, 2.0, 2.5) breaking to a flatter slope ( α = 1.0)
at low frequencies. The frequency of the break was fixed to be
10−3 , in other words the break timescale was 1000 times the bin
size. (The absolute size of the time bins is arbitrary in the simula-
tions. When comparing the simulated results to real data sampled
with e.g. 25 s time resolution, the break timescale in the simulated
PSD is thus 25 ks.)
The numbers given in the table provide an approximate pre-
scription for the expected scatter in the variance of a stationary
process with a red noise PSD similar to that of AGN. The simu-
lated light curve shown in Fig. 2 was used to demonstrate the use
of this table. In this case the PSD is know to have a slope α = 2,
and the variances (shown in panel 3 of Fig. 2) were calculated ev-
ery 20 points. Therefore, the 90 interval for the expected variance
is given by log(S 2 )+0.45
−0.71 . Taking the mean variance as the expec-
tation value for S 2 , this translates to S 2 = 59.9 (11.7 − 168.8).
The interval boundaries were calculated by converting the logarith-
mic value into a linear factor and multiplying by the sample mean
(assumed to represent the true variance). This interval is shown on
Fig. 5 by the dotted lines. The corresponding 99 per cent confidence
interval is also marked.
As expected the individual variances fall within the expected
region. However, the 90 per cent region spans an order of magni-
tude in variance. Thus even order of magnitude differences in vari-
ance between short sections of a light curve are to be expected and
do not necessarily indicate that the underlying process is not sta-
tionary. Subtle changes in the PSD will thus be difficult to de tect
Properties of X-ray light curves from AGN
7
Table 1. Expected scatter in variance estimates. The 90 and 99 per cent
intervals are presented in terms of ±∆ log(S 2 ). (The 99 per cent interval
is given in bold.) The boundaries were calculated from Monte Carlo simu-
lations of light curves. The PSD was chosen to be a broken power-law with
a slope of α = 1 below the break (at a frequency 10−3 ) and a slope above
the break of α = 1.0, 1.5, 2.0, 2.5. The simulated data segments were cho-
sen to be 10, 20, 50, 100, 1000 points long (the timescale of the break in
the PSD being at 1000, in arbitrary units).
α
1.0
1.5
2.0
2.5
10
-0.84
-0.50
+0.33
+0.53
-0.96
-0.61
+0.39
+0.65
-1.16
-0.78
+0.46
+0.75
-1.49
-1.03
+0.52
+0.83
Number of data points
20
50
100
-0.58
-0.36
+0.28
+0.46
-0.75
-0.50
+0.36
+0.61
-1.01
-0.71
+0.45
+0.73
-1.37
-0.98
+0.52
+0.83
-0.40
-0.26
+0.23
+0.39
-0.62
-0.43
+0.34
+0.58
-0.93
-0.67
+0.44
+0.72
-1.31
-0.95
+0.51
+0.82
-0.32
-0.22
+0.20
+0.35
-0.57
-0.40
+0.33
+0.57
-0.90
-0.66
+0.43
+0.72
-1.28
-0.92
+0.50
+0.80
1000
-0.19
-0.13
+0.15
+0.27
-0.45
-0.32
+0.28
+0.49
-0.72
-0.50
+0.36
+0.59
-0.92
-0.63
+0.40
+0.66
Figure 5. Variance of the simulated data shown in Fig. 2 (panel 3) with the
90 (dotted line) and 99 per cent (dashed line) confidence intervals marked
(as calculated in section 5.2.2). Clearly the variances fall within these limits,
as expected for a stationary process. The solid line marks the mean variance.
by examining the raw variances as the intrinsic scatter is so large.
Such changes could be revealed by comparing averaged variances
or comparing the PSDs as described above.
6 EFFECT OF MEASUREMENT ERRORS
6.1 Excess variance and Fvar
The datasets considered thus far have been ideal, in the sense that
they are free from flux uncertainties. In real life, however,
a light
8
Vaughan et al.
(8)
(9)
(10)
.
1
N
σ 2
err =
σ 2
err,i .
curve xi will have finite uncertainties σerr,i due to measurement er-
rors (such as Poisson noise in the case of an X-ray photon counting
signal). These uncertainties on the individual flux measure ments
will contribute an additional variance. This leads to the use of the
‘excess variance’ (Nandra et al. 1997; Edelson et al. 2002) as an
estimator of the intrinsic source variance. This is the variance after
subtracting the contribution expected from measurement errors
XS = S 2 − σ 2
σ 2
err ,
err is the mean square error
where σ 2
N
Xi=1
The normalised excess variance is given by σ 2
XS/ ¯x2 and
NXS = σ 2
the fractional root mean square (rms) variability amplitude (Fvar ;
Edelson, Pike & Krolik 1990; Rodriguez-Pascual et al. 1997) is the
square root of this, i.e.
Fvar = r S 2 − σ 2
err
¯x2
NXS al-
The statistic Fvar is often chosen in preference to σ 2
though the two convey exactly the same information. Fvar is a
linear statistics and can therefore give the rms variability ampli-
tude in per centage terms. The choice of whether to quote Fvar
or σ 2
NXS is usually purely one of presentation. It is worth not-
ing that the Monte Carlo results given in section 5.2.2, to esti-
mate the expected scatter on the variance, can also be applied
to its square root. The expected boundaries of the confidence
re-
gion of the logarithm of the rms is approximately half those of
the variance. Speci fically, ∆ log(σ ) ≈ ∆ log(S 2 )/2 and similarly
NXS )/2.
∆ log(Fvar ) ≈ ∆ log(σ 2
6.2 Spectral variability
An X-ray light curve of an AGN can be split into different energy
bands. The light curves in each band will be strictly simultaneous
and can be used to test whether the X-ray variability is a function
of energy. For example, one might examine the ratio of a soft band
light curve to a hard band light curve. The statistical signi ficance
of any variations in the ratio can be quanti fied by propagatin g the
measurement errors and applying an appropriate test, such as the
χ2 test of the constant ratio hypothesis8 . If the ratio shows varia-
tions greater than those expected from the errors then the two light
curves are intrinsically different and the source does indeed show
spectral variability. Such changes in the energy spectrum with time
can in principle provide valuable clues to the nature of the X-ray
source. This test does not provide any quantitative description of
the spectral variability.
Another tool for investigating spectral variability is the rms
spectrum, i.e. the rms variability amplitude (or Fvar ) as a func-
tion of energy. See e.g. Inoue & Matsumoto (2001), Edelson et al.
(2002), Fabian et al. (2002) and Schurch & Warwick (2002) for
some examples of rms spectra from AGN. However, when exam-
ining rms spectra it is often not clear whether changes in the am-
plitude with energy reflect real energy-dependence of the in trinsic
variability amplitude or are caused by random errors in the light
s (e.g.
curves. The finite measurement errors on the individual fluxe
8 This does, of course, assume the light curves have been binned suffi-
ciently for the error bars to be approximately Gaussian.
due to Poisson noise) will introduce some uncertainty in the esti-
mated rms amplitudes. An estimate of this uncertainty would help
answer the question posed above, namely whether features in rms
spectra are the result of random errors in the data or represent spec-
tral variations intrinsic to the source.
The problem of how to assess the uncertainty on the excess
variance (or Fvar ) is a long-standing one (e.g. Nandra et al. 1997;
Turner et al. 1999; Edelson et al. 2002). The standard error formu-
lae presented in the literature (e.g. Turner et al. 1999; Edelson et al.
2002) are formally valid in the case of un-correlated Gaussian pro-
cesses. Typically AGN light curves at different X-ray energies are
strongly correlated and are not Gaussian. However, when searching
for subtle differences in amplitude between simultaneous and cor-
related light curves it may be more useful to have an indication of
the uncertainty resulting from the finite flux errors.
6.2.1 Uncertainty on excess variance due to measurement errors
. (11)
A Monte Carlo approach was used to develop a prescription of the
effect of measurement errors on estimates of Fvar (and σ 2
NXS ). A
short red noise light curve was generated. Poisson noise was added
(i.e. the individual flux measurements were randomised foll owing
the Poisson distribution) and the excess variance was recorded. The
fluxes of the original light curve were randomised again and t he ex-
cess variance recorded, this was repeated many times. The distribu-
tion of excess variances was then used to determine the uncertainty
in the variance estimate caused by Poisson noise. Full details of the
procedure are given in Appendix B
For these simulations it was found that the error on σ 2
NXS de-
creases as the S/N in the light curve is increased according to:
NXS) = vuut(r 2
¯x )2
¯x2 )2
+ (r σ 2
σ 2
2Fvar
err
err
err(σ 2
N ·
N ·
See appendix B for details of this equation and its equivalent in
terms of Fvar .
As this only accounts for the effect of flux measurement er-
rors (such as Poisson noise) in a given light curve it can be used
to test whether two simultaneously observed light curves of the
same source, but in different bands, show consistent amplitudes.
A demonstration of this using real data is given in the following
section. This uncertainty does not account for the random scatter
intrinsic to the red noise process, therefore the absolute value of
the rms spectrum will change between realisations (i.e. from epoch
to epoch). But if a source shows achromatic variability then the
values of Fvar calculated in each energy band (at a given epoch)
should match to within the limits set by the Poisson noise (i.e. the
fractional rms spectrum should be constant to within the uncertain-
ties given by the above equation). Differences in Fvar signi ficantly
larger than these would indicate that the source variability ampli-
tude is a function of energy. This would then mean the PSD ampli-
tude/shape is different in different energy bands, or there are multi-
ple spectral components that vary independently.
The above uncertainty estimates can be used to test the hy-
pothesis that the source variability is achromatic. If signi ficant dif-
ferences between energy bands are detected (as in the case of Mrk
766 presented below) then these errors should not be used to fi
t
the rms spectrum. The assumption that the differences are due only
to measurement errors is no longer the case. In such situations the
light curves in adjacent energy bands are likely to be partially cor-
related and so χ2 - fitting of the rms spectrum is not appropriate. The
Properties of X-ray light curves from AGN
9
Figure 7. Excess variance of the Mrk 766 data shown in Fig. 6 (panel 3)
with the 90 (dotted line) and 99 per cent (dashed line) confidence inter-
vals marked (as calculated in section 5.2.2). The variances fall within these
limits, as expected for a stationary process. The solid line marks the mean
variance.
7.1 Observation details
Mrk 766 was observed by XMM-Newton (Jansen et al. 2001) over
the period 2001 May 20 – 2001 May 21 (rev. 265). The present
analysis is restricted to the data from the pn European Photon Imag-
ing Camera (EPIC), which was operated in small window mode.
Extraction of science products from the Observation Data Files
(ODFs) followed standard procedures using the XMM-Newton Sci-
ence Analysis System (SAS) v5.3.3. Source data were extracted
from a circular region of radius 35 arcsec from the processed image
and only events corresponding to patterns 0–4 (single and do uble
pixel events) were used. Background events were extracted from
regions in the small window least effected by source photons, these
showed that the background rate increased dramatically during the
final ∼ 1.5 × 103 s of the observation. This section of the data
was excluded, leaving 1.05 × 105 s of uninterrupted data. The light
curves were corrected for telemetry drop outs (less than 1 per cent
of the total time) and background subtracted. The errors on the light
curves were calculated by propagating the Poisson noise.
7.2 Stationarity of the data
The broad band (0.2–10 keV) light curve extracted from the pn is
shown in Fig. 6 (panel 1). As was the case for the simulated data
shown in Fig 2, the mean and variance (calculated every 20 data
points)10 show changes during the length of the observation (pan-
els 2 and 3). The expected range for the excess variance, calculated
using the results of section 5.2.2 (and assuming a PSD slope of
α = 2.0), is marked in figure 7. Fig. 8 shows the same data in
terms of normalised excess variances. Neither of these show fluc-
tuations larger than expected for a stationary process. But given the
large expected scatter this is a rather insensitive test. In the case of
the Mrk 766 light curve however, there are sufficient data to e xam-
ine variations of the average variance with time, allowing a more
sensitive test for non-stationarity.
Figure 6. Top panel: 0.2–10.0 keV pn light curve of Mrk 766 (with 25 s
bins, in units of ct s−1 ). Panel 2 and 3: mean count rate and excess vari-
ance measured from segments of 20 points. Panel 4: averaged normalised
excess variance measured by binning the individual variance estimates into
groups of 20. This average variance is inconsistent with constant. Panel 5:
fractional rms amplitude measured from segments of 20 points. Panel 6:
averaged fractional rms amplitude measured by binning the individual am-
plitudes into groups of 20. This average fractional amplitude is consistent
with constant. This contrasts with the situation shown in Fig. 2
differences in excess variance will be a combination of intrinsic dif-
ferences and measurement errors. Their uncertainty will therefore
be more difficult to quantify.
7 CASE STUDY: AN XMM-Newton OBSERVATION OF
Mrk 766
In this section a long (∼ 105 s) XMM-Newton observation of the
bright, variable Seyfert 1 galaxy Markarian 766 is used to illus-
trate the points discussed above. The data were obtained from the
XMM-Newton Data Archive9 . Details of the observation are dis-
cussed in Mason et al. (2003) and an analysis of the PSD is pre-
sented in Vaughan & Fabian (2003).
9 http://xmm.vilspa.esa.es
10 These correspond to ‘instantaneous’ estimates of the source variance on
timescales of 50 − 500 s.
10
Vaughan et al.
Figure 8. As for Fig. 7 but using the normalised excess variance of the Mrk
766 data.
By averaging the excess variance estimates (in time bins con-
taining 20 excess variance estimates) signi ficant changes i n the
variance with time are revealed (panel 4). This contrasts with the
simulated data shown in Fig. 2. The binned excess variance is in-
consistent with a constant hypothesis: fitting with a consta nt gave
χ2 = 23.1 for 9 degrees of freedom (dof ), rejected at 99 per cent
confidence. The average variance is therefore changing with time,
indicating the variability is strongly non-stationary.
A careful inspection of Fig. 6 (panels 3 and 4) shows the indi-
vidual variance estimates have a tendency to track the source count
rate. This is difficult to discern from the individual varian ces (panel
3), due to the larger intrinsic scatter, but much clearer in the av-
eraged variances (panel 4). This can be seen clearly in Fig. 9 (top
panel) where the rms amplitude (pσ 2
XS ) is shown as a function of
count rate. To produce this plot the individual rms estimates (Fig. 6,
panel 3) were sorted by count rate and binned by flux (such that
there were 20 estimates per bin). The error on the mean rms was
calculated in the standard fashion (see above). This indicates that
the source does show a form of genuine non-stationarity: the abso-
lute rms amplitude of the variations increases, on average, as the
source flux increases. This effect has been noted in other Sey ferts
(Uttley & McHardy 2001; Edelson et al. 2002; Vaughan et al. 2003)
and is due to a linear correlation between rms and flux (see Utt
ley
et al. in prep. for further discussion of this effect). Non-stationarity
of this form can be ‘factored out’ by using the normalised ampli-
NXS ) instead of the absolute values. Normalising
tude (Fvar or σ 2
each variance (or rms) estimate by its local flux removes this
trend.
The bottom panel of Fig. 9 shows that Fvar is indeed constant with
flux ( fitting a constant gave
χ2 = 7.8 for 9 dof ). Fig. 6 (panels 5
and 6) shows Fvar and its average as a function of time; the aver-
age is consistent with staying constant (χ2 = 5.8 for 9 dof ). The
variability of Mrk 766 does show genuine (strong) non-stationarity,
in the sense that the absolute rms increases linearly with flu x, but
this trend can be removed by using normalised units – Fvar (and
therefore the normalised excess variance) is consistent with being
constant (with time and flux).
The above analysis suggests that, after accounting for the ef-
fect of the rms–
flux correlation, there is no other evidence f
or
strong non-stationarity in the rapid variability of Mrk 766. This
was confirmed using the s-test of Papadakis & Lawrence (1995; see
their Appendix A). A periodogram was calculated for three consec-
utive segments of 3.4 × 104 s duration, and normalised to fractional
Figure 9. Top panel: The average absolute rms amplitude (pσ2
XS ) as a
function of flux for the Mrk 766 light curve (compare with Fig. 4). Bottom
panel: The average it fractional rms amplitude (pσ2
NXS ) as a function
of flux. Clearly the absolute rms amplitude is a function of flu
x, but this
dependence is removed in the fractional rms.
units (see Appendix A). The s value was computed by comparing
periodograms at frequencies below 2 × 10−3 Hz (above which the
Poisson noise power becomes comparable to the source variabil-
ity). For each pair of periodograms the value of s was within the
range expected for stationary data (speci fically s < 1, within one
standard deviation of the expected value).
7.3 rms spectrum
The variability amplitude as a function of energy was calculated
by measuring Fvar from light curves extracted in various energy
ranges. The results are shown in Fig. 10 and the errors were cal-
culated using equation B2 to account for the effect of Poisson
noise. The variability amplitude is clearly a function of energy,
i.e. Mrk 766 shows signi ficant spectral variability. This wa s con-
firmed by a Fourier analysis of the light curves in different e nergy
bands (Vaughan & Fabian 2003) which revealed complex energy-
dependent variability.
The rms spectrum was re-calculated for light curves contain-
ing only single pixel (pattern 0) events and again for double pixel
(patterns 1–4) events. These two sets of data were extracted from
the same detector and using identical extraction regions etc. Af-
Properties of X-ray light curves from AGN
11
curves show large fluctuations in variance. These changes pr ovide
little insight as they are expected even when the underlying physical
process responsible for the variability is constant. Rather, they may
simply be statistical fluctuations intrinsic to the stochas tic process.
All red noise processes show random fluctuations in both mean
and variance and the variance will be distributed in a non-Gaussian
fashion with a large scatter (see sections 4.2 and 5).
Previous claims of non-stationary variability based on changes
in variance (e.g. Nandra et al. 1997; Dewangan et al. 2002; Gliozzi,
Sambruna & Eracleous 2003) should therefore be treated with cau-
tion since they did not account for this intrinsic scatter (see also
section 3.3.1 of Leighly 1999 for a discussion of this point). Real
changes in the PSD would indicate genuine non-stationarity and re-
flect real changes in the physical conditions of the variabil
ity pro-
cess. Such changes can be measured from the average properties of
the light curve, such as the averaged periodogram or the averaged
variance (see section 5).
A different issue is that differences between the variance of si-
multaneous light curves obtained in different energy bands can be
examined using the excess variance (or Fvar ) statistic. It is possible
to estimate the uncertainty in the excess variance due to errors in
the flux measurements. This uncertainty, accounting only fo r mea-
surement (e.g. Poisson) errors, can be used when testing for spectral
variability, as demonstrated in section 7.
Estimators such as the excess variance provide a useful, if
crude, means of quantifying the variability of AGN. Even though
the stochastic nature of AGN light curves makes it difficult t o es-
timate variability amplitudes robustly from short observations, the
excess variance can provide useful information. For example an
analysis of the excess variances measured from short observations
of Seyfert 1 galaxies demonstrated that the variability amplitude (at
a given range of timescales) is inversely correlated with the lumi-
nosity of the source (Nandra et al. 1997; Leighly 1999; Markowitz
& Edelson 2001). Although random fluctuations in variance ar e ex-
pected for AGN light curves the range of variances observed is far
larger than could be accounted for by this effect alone. Another
example is given in section 7.2 when it is demonstrated that the av-
erage variance of Mrk 766 is a function of the flux of the source .
A similar effect has been observed in X-ray binaries (Uttley &
McHardy 2001). A discussion of the implications of this result will
be given in Uttley et al. in prep.
9 CONCLUSIONS
This paper discusses some aspects of quantifying the variability of
AGN using simple statistics such as the variance. Various possible
uses of these are presented and some possible problems with their
signi ficance and interpretation are brought to light. The pr imary
issues are as follows:
(i) In order to search for non-stationary variability in an ensem-
ble of short light curves (or short light curve segments) one can test
whether the individual variances are consistent with their mean.
Two practical methods are presented (sections 5.2.1 and 5.2.2).
(ii) In the first method the mean variance and its error are cal -
culated at various epochs by binning the individual variance esti-
mates. This is most useful when searching for subtle changes in
variability amplitude but requires large datasets (in order that the
variance can be sufficiently averaged).
(iii) In the second method the individual variance estimates are
compared with the expected scatter around the mean. The expected
scatter is calculated using Monte Carlo simulations of stationary
Figure 10. rms spectrum of Mrk 766 measured using EPIC pn light curves
with 1000 s bins.
ter accounting for the difference in count rate between single and
double pixel events the two sets of light curves should be identi-
cal except for the effects of Poisson noise. The two rms spectra
should be the same except for Poisson errors. Comparing the ra-
tio of the two rms spectra using a χ2 -test (against the hypothesis
of unity ratio) gave χ2 = 25.3/20 degrees of freedom. Compar-
ing the difference of the two rms spectra to the hypothesis of zero
difference gave identical results and shows the two rms spectra are
indeed fairly consistent. This test indicates that for real data the er-
ror formula given above does provide a reasonable description of
the uncertainty induced by photon noise.
8 DISCUSSION
The analysis of stochastic processes, such as X-ray variability of
AGN, is conceptually different from the analysis of deterministic
data such as time-averaged spectra (see discussions in e.g. Jenkins
& Watts 1968; Priestley 1981; Bendat & Piersol 1986; Bloomfie ld
2000). For example, when observing the spectrum of a constant
source one expects repeatability of the data to within the limits set
by measurement errors, i.e. each new realisation of the spectrum
should be consistent within the errors. In AGN variability analy-
sis it is the signal itself that is randomly variable; one does not
expected repeatability of quantities such as the mean or variance.
These statistical moments will change (randomly) with each new
light curve even if there are no measurement errors.
The stochastic nature of red noise processes means that it is
usually only their average properties that can provide physical in-
sight. Non-deterministic data should be handled statistically. For
example, it is customary to examine the timing properties of X-ray
binaries using PSDs estimated from the average periodogram of an
ensemble of light curves (e.g. van der Klis 1995). Averaging over
many independent realisations reduces the random fluctuati ons in-
herent in the noise process.
In most AGN timing studies however there are rarely enough
data to construct averages in this way (but see Papadakis &
Lawrence 1993 and Uttley et al. 2002 for more on PSD estima-
tion for AGN). As a result of this relative lack of data, AGN tim-
ing studies often emphasise the properties of a single light curve.
But emphasis on the detailed properties of any single realisation
of a stochastic process can be misleading. For example, AGN light
12
Vaughan et al.
processes. The table gives some examples of the scatter expected
for various PSD shapes typical of AGN. This table can therefore
be used to provide a ‘quick look’ at whether the observed fluct u-
ations in the variance are larger than expected. One drawback is
that, because the intrinsic scatter in the variance is rather large for
red noise data, this method is only sensitive to very large changes
in the variability amplitude. Another drawback is that one has to
assume a shape for the PSD.
(iv) The excess variance can also be used to quantify how the
variance changes as a function of energy (section 6.2). An approx-
imate formula is presented (based on the results of Monte Carlo
simulations) that gives the expected error in the excess variance
resulting from only observation uncertainties ( flux errors
such as
Poisson noise). This can be used to test for signi ficant diffe rences
in variance between energy bands. If the normalised excess vari-
ances (or Fvar s) are found to differ signi ficantly between energy
bands this implies the PSD is energy dependent and/or there are
independently varying spectral components.
(v) Possibly the most robust yet practical approach to variabil-
ity analysis from AGN data is to test the validity of hypotheses
using Monte Carlo simulations. This approach has yielded reliable
PSD estimates for Seyfert galaxies (Green et al. 1999; Uttley et al.
2002; Vaughan et al. 2003; Markowitz et al. 2003) and has been
used to test the reliability of cross-correlation results (e.g. Welsh
1999) amongst other things. Section 3 discusses some methods for
simulating red noise data.
ACKNOWLEDGEMENTS
We are very grateful
to the referee, Andy Lawrence, for a
thoughtful report that prompted signi ficant improvements t o the
manuscript. SV and PU acknowledge support from PPARC. This
paper made use of observations obtained with XMM-Newton, an
ESA science mission with instruments and contributions directly
funded by ESA Member States and the USA (NASA).
REFERENCES
Belloni T., Hasinger G., 1990, A&A, 227, L33
Belloni T., Psaltis D., van der Klis M., 2002, ApJ, 572, 392
Bendat J. S., Piersol A. G., 1986, Random Data: Analysis and Measurement
Procedures, Wiley (New York)
Bevington P. R., Robinson D. K,. 1992, Data Reduction and Error Analysis
for the Physical Sciences, McGraw-Hill (New York)
Bloom field P, 2000, Fourier Analysis of Time Series, Wiley (New York)
Deeming T. J., 1970, AJ, 75, 1027
Deeming T. J., 1975, Ap&SS, 36, 137
Deeter J. E., Boynton P. E., 1982, ApJ, 261, 337
Dewangan G. C., Boller Th., Singh K. P., Leighly K. M., 2002, A&A, 390,
65
Done C., Madejski G. M., Mushotzky R. F., Turner T. J., Koyama K., Ku-
nieda H., 1992, ApJ, 400, 138
Edelson R., Pike G. F., Krolik J. H., 1990, ApJ, 359, 86
Edelson R., Nandra K. 1999, ApJ, 514, 682
Edelson R. et al. 2002, ApJ, 568, 610
Fabian A. C. et al. 2002, MNRAS, 335, L1
Green A. R., McHardy I. M., Lehto H. J. 1993, MNRAS, 265, 664
Green A. R., McHardy I. M., Done C., 1999, MNRAS, 305, 309
Gliozzi M., Sambruna R. M., Eracleous M., 2003, ApJ, 584, 176
Inoue, H., Matsumoto, C., 2001, AdSpR, 28, 445
Jenkins G. M., Watts, D. G., 1968, Spectral Analysis and its Applications,
Holden-Day (San Fancisco)
Lawrence A., Watson M. G., Pounds K. A., Elvis M., 1987, Nature, 325,
694
Lawrence A., Papadakis I., 1993, ApJ, 414, L85
Leahy D. A., Darbro W., Elsner R. F., Weisskopf M. C., Kahn S., Sutherland
P. G., Grindlay J. E., 1983, ApJ, 266, 160
Lehto H. J., 1989, in J. Hunt, B. Battrick, eds, Two Topics in X Ray Astron-
omy, (ESA SP-296; Noordwijk: ESA), p499
Leighly K., 1999, ApJS, 125, 297
Markowitz A., Edelson R., 2001, ApJ, 547, 684
Markowitz A. et al. 2003, ApJ, in press (astro-ph/0303273 )
Mason K. O., et al. 2003, ApJ, 582, 95
McHardy I. M., 1989, in J. Hunt, B. Battrick, eds, Two Topics in X Ray
Astronomy, (ESA SP-296; Noordwijk: ESA), p1111
Nandra K., George I. M., Mushotzky R. F., Turner T. J., Yaqoob T., 1997,
ApJ, 476, 70
Papadakis I. E., Lawrence A., 1993, MNRAS, 261, 612
Papadakis I. E., Lawrence A., 1995, MNRAS, 272, 161
Pottschmidt K., 2002, A&A, submitted (astro-ph/0202258 )
Press W. H. 1978, Comments on Astrophysics, 7, 103
Press W. H., Teukolsky S. A., Vetterling W. T., Flannery B. P., 1996, Nu-
merical Recipes, Cambridge Univ. Press. (Cambridge)
Press W. H., Rybicki G. B., 1997, in Astronomical Time Series, D. Maoz,
A. Sternberg, E. M. Leibowitz (eds.), Kluwer (Dordrecht), p61
Priestley M. B., 1981, Spectral Analysis and Time Series, Academic Press
(London)
Rodriguez-Pascual P. M. et al. , 1997, ApJS, 110, 9
Scargle J. D., 1981, ApJS, 45, 1
Schurch N. J., Warwick R. S., 2002, MNRAS, 334, 811
Stella L., Arlandi E., Tagliaferri G., Israel G. L., 1997, in ‘Applications
of Time Series Analysis in Astronomy and Meteorology’, T. Subba
Rao, M. B. Priestley, O. Lessi (eds.), Chapman & Hall (London)
(astro-ph/9411050 )
Timmer J., K onig M. 1995, A&A, 300, 707
Turner T. J., George I. M., Nandra K., Turcan D., 1999, ApJ, 524, 667
Uttley P., McHardy I. M., 2001, MNRAS, 323, 26
Uttley P., McHardy I. M., Papadakis, I. 2002, MNRAS, 332, 231
van der Klis M., 1989, in Timing Neutron Stars, H. Ogelman, E. P. J. van
den Heuvel (eds.), Kluwer (Dordrecht), NATO ASI Series C 262, p27
van der Klis M., 1995, in X-ray Binaries, W. H. G. Lewin, J. van Paradijs, E.
P. J. van den Heuvel (eds.), Cambridge Univ. Press (Cambridge), p252
van der Klis M., 1997, in Statistical Challenges in Modern Astronomy II,
ed. G.J. Babu, E.D. Feigelson, Springer-Verlag (New York), p321
Vaughan S., Fabian A. C., Nandra K. 2003, 339, 1237
Vaughan S., Fabian A. C., 2003, MNRAS, 341, 496
Welsh W. F., 1999, PASP, 111, 1347
APPENDIX A: PERIODOGRAM NORMALISATION
(A1)
The periodogram is calculated by normalising the modulus-squared
of the DFT (see equation 1):
P (fj ) = ADF T (fj )2 .
There are a variety of options for the normalisation A used
in the literature, each has desirable properties. In the mathematical
literature on time series analysis a normalisation of the form A =
2/N is standard (e.g. Priestley 1981; Bloomfield 2000). However ,
this normalisation is generally not used for time series analysis in
astronomy because the periodogram then depends on flux of the
source and the binning of the time series. Below are listed three
of the most commonly used normalisations, which only differ by
factors of ¯x, the mean count rate in cts/s (Aabs = ¯xALeahy =
¯x2Arms2 ). The factor of two is present in all these normalisations
to make the periodogram ‘one sided,’ meaning that integrating over
positive frequencies only yields the correct variance.
(i) Arms2 = 2∆Tsamp / ¯x2N – defined by van der Klis (1997)
(see also Miyamoto et al. 1991). This is the normalisation most
often used in analysis of AGN and X-ray binaries because the inte-
grated periodogram yields the fractional variance of the data. The
units for the periodogram ordinate are (rms/mean)2 Hz−1 (where
rms/mean is the dimensionless quantity Fvar ), or simply Hz−1 .
If a light curve consists of a binned photon counting signal (and
in the absence of other effects such as detector dead-time) the ex-
pected Poisson noise ‘background’ level in its periodogram is given
by
,
(A2)
Pnoise =
2( ¯x + B )
∆Tsamp
¯x2
∆Tbin
where ¯x is the mean source count rate, B is the mean background
count rate, ∆Tsamp is the sampling interval and ∆Tbin is the time
bin width. The factor of ∆Tsamp /∆Tbin accounts for aliasing of
the Poisson noise level if the original photon counting signal con-
tained gaps. If the light curve is a series of contiguous time bins
(i.e. ∆Tbin = ∆Tsamp ) and has zero background (which is approx-
imately true for many XMM-Newton light curves of AGN) then this
reduces to Pnoise = 2/ ¯x.
For a light curve with Gaussian errors σerr,i the noise level in the
periodogram is
.
(A3)
Pnoise =
2∆Tbinσ 2
∆Tsamp
err
¯x2
∆Tbin
(ii) ALeahy = 2∆Tsamp / ¯xN – originally due to Leahy et al.
(1983). This has the property that the expected Poisson noise level
is simply 2 (for continuous, binned photon counting data). If the
light curve consists only of Poisson fluctuations then the pe ri-
odogram should be distributed exactly as χ2
2 . It is this property that
makes this normalisation the standard for searching for periodic
signals in the presence of Poisson noise (see Leahy et al. 1983).
If the input light curve is in units of ct s−1 then the periodogram
ordinate is in units of ct s−1 Hz−1 .
(iii) Aabs = 2∆Tsamp /N – this is the normalisation used in
equation 2. This gives the periodogram in absolute units [e.g. (ct
s−1 )2 Hz−1 ] and so the integrated periodogram gives the total vari-
ance in absolute units [e.g. (ct s−1 )2 ] For a contiguously binned
light curve with Poisson errors the noise level is Pnoise = 2 ¯x, and
for Gaussian errors the noise level is Pnoise = 2∆Tbinσ 2
err .
APPENDIX B: MONTE CARLO DEMONSTRATION OF
POISSON NOISE INDUCED UNCERTAINTY ON EXCESS
VARIANCE
To estimate the effect on σ 2
NXS due only to Poisson noise the basic
strategy was as follows.
(i) Generate a random red noise light curve. This acts as the
‘true’ light curve of the source.
(ii) Add Poisson noise, i.e. draw fluxes from the light curve a c-
cording to the Poisson distribution. This simulates ‘observing’ the
true light curve. Error bars were assigned based on the ‘observed’
counts in each bin (√counts).
(iii) Measure the normalised excess variance σ 2
NXS of the ob-
served light curve. This will be different from the variance of the
true light curve because of the Poisson noise.
Steps 2 and 3 were repeated, using the same true light curve, to
obtain the distribution of σ 2
NXS . Fig. B1 shows some results. In this
example the ‘true’ light curve was generated with a f −2 PSD and
Properties of X-ray light curves from AGN
13
normalised to a pre-defined mean and variance, e.g. S 2 / ¯x2 = 0.04
(Fvar = 20%). This light curve was then observed (i.e. steps 2
and 3 were repeated) 104 times11 . The three panels correspond to
different mean count rates for the true light curve (i.e. different S/N
of the observation). The (1σ ) widths of the σ 2
NXS distributions are
Monte-Carlo estimates of the size of the error bars on σ 2
NXS due to
Poisson noise.
As is clear from Fig. B1 the distribution of σ 2
NXS becomes
narrower, i.e. the error on σ 2
NXS gets smaller, as the S/N of the
data increases. Obviously in the limit of very high S/N data the
measured value of σ 2
NXS will tend to the ‘true’ value (in this case
0.04), i.e. err(σ 2
NXS) → 0 as counts → ∞. It should also be
noted that the distributions are quite symmetrically centred on the
correct value, indicating that σ 2
NXS is an unbiased estimator of the
intrinsic variance in the light curve, even in relatively low S/N data.
In order to assess how the error on σ 2
NXS changes with S/N,
the width of its distribution was measured from simulated data at
various different settings of S/N and intrinsic variance (i.e. S 2 / ¯x2 ).
Width of the distribution at each setting was calculated from only
500 ‘observations’ of each light curve. In order that no particular
realisation adversely affect the outcome, and to increase the statis-
tics, this was repeated for 20 different random light curves (of the
same fractional variance) and the width of the σ 2
NXS distributions
were averaged (i.e. the whole cycle of steps 1–3 was repeated 20
times). Thus for each speci fied value of S/N and fractional va riance,
the error on σ 2
NXS is estimated from 104 simulated ‘observations.’
These Monte Carlo estimated errors on the normalised excess vari-
ance are shown in Fig. B2.
The solid lines show the functions defined by equation 11
(which was obtained by fitting various trial functions to the Monte
Carlo results). Clearly this equation gives a very good match to the
Monte Carlo results.
If the variability is not well detected, either because the S/N
is low or the intrinsic amplitude is weak, then S 2 ≈ σ 2
err . It is the
first term on the right hand side of equation 11 that dominates . If
the variability is well detected, i.e. S 2 ≫ σ 2
err , then it is the second
term that dominates.
NXS) ≈
N · σ2
p 2
: S 2 ≈ σ 2
err
err
¯x2
err(σ 2
q σ2
N · 2Fvar
: S 2 ≫ σ 2
err .
err
¯x
In the former case the deviations from the mean are dominated
by the errors and the fluxes are approximately normally distr ibuted.
In this regime the error equation becomes the same as that given
in equation A9 of Edelson et al. (2002). In the latter case the de-
viations in the light curve are enhanced by the intrinsic variance.
The second term is similar to the first except multiplied by a f actor
q2σ 2
err to account for this.
XS /σ 2
Equation 11 can be used to give the uncertainty on Fvar thusly
1
err(σ 2
err(Fvar ) =
NXS) =
2Fvar
vuut(r 1
¯x2Fvar )2
+ (r σ 2
σ 2
err
err
N ·
2N ·
11 As this measures only the effect due to Poisson noise, the results are
largely independent of the details of the light curve, including the PSD, as
long as the flux is non-zero throughout the light curve. This w as confirmed
by repeating the above experiment using data produced from PSD slopes in
the range α = 0 − 2.
¯x )2
1
(B1)
,
(B2)
Figure B2. Width of the distribution of σ2
NXS (resulting from Poisson
noise) as a function of the number of counts per bin. Compare with Fig. B1.
The solid curve shows the function described in the text (equation 11).
errors on the fluxes. It does not account for the intrinsic sca tter in
the fluxes inherent in any red noise process.
This paper has been typeset from a TEX/ LATEX file prepared by the
author.
14
Vaughan et al.
Figure B1. Distribution of measured σ2
NXS from 10,000 ‘observations’ of
the same light curve. In each case the ‘true’ σ2
NXS is 0.04 (dotted line). The
top panel used the lowest S/N data, the bottom panel used the highest S/N
data. The mean number of counts per bin in the simulated light curves was
15 (top), 30 (middle) and 100 (bottom). As the S/N increases (count rate
increases) the distribution of σ2
NXS becomes narrower. (Note: this is differ-
ent from Fig. 3, which shows how the variance changes between different
realisations of the same stochastic process.)
and this is the equation used to derive the errors shown in Fig. 10.
In the two regimes this becomes:
err(Fvar ) ≈
2N · σ2
p 1
: S 2 ≈ σ 2
err
err
¯x2 Fvar
q σ2
: S 2 ≫ σ 2
N · 1
err .
err
¯x
In the first instance, when the variability is not well detect ed, σ 2
NXS
should be preferred over Fvar as negative values of σ 2
NXS are possi-
ble. Additional Monte Carlo simulations confirmed the above equa-
tions are valid for both Gaussian and Poisson distributed flu x errors.
It is worth reiterating that this error accounts only for measurement
(B3)
|
astro-ph/0404556 | 1 | 0404 | 2004-04-28T13:32:53 | The structure of the ICM from High Resolution SPH simulations | [
"astro-ph"
] | We present results from a set of high (512^3 effective resolution), and ultra-high (1024^3) SPH adiabatic cosmological simulations of cluster formation aimed at studying the internal structure of the intracluster medium (ICM). We derive a self-consistent analytical model of the structure of the intracluster medium (ICM). We discuss the radial structure and scaling relations expected from purely gravitational collapse, and show that the choice of a particular halo model can have important consequences on the interpretation of observational data. The validity of the approximations of hydrostatic equilibrium and a polytropic equation of state are checked against results of our simulations. The properties of the ICM are fully specified when a 'universal' profile is assumed for either the dark or the baryonic component. We also show the first results from an unprecedented large-scale simulation of 500 Mpc/h and 2 times 512^3 gas and dark matter particles. This experiment will make possible a detailed study of the large-scale distribution of clusters as a function of their X-ray properties. | astro-ph | astro-ph |
Outskirts of Galaxy Clusters: intense life in the suburbs
Proceedings IAU Colloquium No. 195, 2004
A. Diaferio, ed.
c(cid:13) 2004 International Astronomical Union
DOI: 00.0000/X000000000000000X
The structure of the ICM from High
Resolution SPH simulations
G. Yepes1, Y. Ascasibar2, R. Sevilla1, S. Gottlober3 and V. Muller3
1Grupo de Astrof´ısica, Universidad Aut´onoma de Madrid, Madrid E-280049, Spain
2Theoretical Physics, University of Oxford, 1 Keble Road, Oxford OX1 3NP, United Kingdom
3Astrophysikalisches Institut Potsdam, An der Sternwarte 16, Potsdam D-14482, Germany
Abstract. We present results from a set of high (5123 effective resolution), and ultra-high
(10243) SPH adiabatic cosmological simulations of cluster formation aimed at studying the in-
ternal structure of the intracluster medium (ICM). We discuss the radial structure and scaling
relations expected from purely gravitational collapse, and show that the choice of a particular
halo model can have important consequences on the interpretation of observational data. The
validity of the approximations of hydrostatic equilibrium and a polytropic equation of state are
checked against results of our simulations. We also show the first results from an unprecedented
large-scale simulation of 500 h−1 Mpc and 2 × 5123 gas and dark matter particles. This experi-
ment will make possible a detailed study of the large-scale distribution of clusters as a function
of their X-ray properties.
1. Introduction
Galaxy clusters are a unique laboratory to test the hierarchical paradigm of structure
formation. They are the best probes of the large scale structure of the Universe and have
often been used as a diagnostic of the cosmological parameters. The intrinsic non-linear
nature of gravitational collapse and gas dynamics makes numerical simulations the most
useful tool to study in detail the process of cluster formation and evolution.
Simple analytical models for the structure of the ICM can be derived from the hy-
potheses of hydrostatic equilibrium and polytropic equation of state, P ∝ ργ
g . But real
clusters might not be well described by these two hypothesis. For instance, kinetic energy
makes a significant contribution to the energy budget of merging systems, and therefore
thermally-supported hydrostatic equilibrium ceases to be a valid approximation. Even in
relaxed systems, this assumption is not very accurate in the outermost parts, where gas
motions become more important. Departures from spherical symmetry can also play a
role in the final structure of the ICM (Lee & Suto 2003) and, last but not least, there is
no obvious physical reason for the gas to follow a polytropic relation. From our numeri-
cal experiments, we showed (see Ascasibar et al. (2003)) that hydrostatic equilibrium is
fulfilled within ∼ 20 per cent accuracy by all simulated clusters, as long as they are not
heavily disturbed. A polytropic equation of state seems to be a good approximation as
well, although its reliability near the centre is still a matter of debate (see e.g. Rasia et al.
2003). From our data we derive a polytropic index of γ ∼ 1.18.
We compared four different analytical halo models with our simulations. The first two
models assume that haloes are well described by Navarro et al. (1997) and Moore et al.
(1999) fitting formulae, while the other two assume that the gas follows a β-model
(Cavaliere & Fusco-Femiano 1976). We consider a 'canonical' version of the β-model, in
which the gas is isothermal (γ = 1) and β = 2/3, and a polytropic version with γ = 1.18
and β = 1. The same value of the polytropic index has been used for the first two models
1
2
G. Yepes et al.
as well. Hereafter we will use the abbreviations NFW, MQGSL, BM and PBM to refer to
these models. For a detailed description, the reader is referred to Ascasibar et al. (2003).
2. Numerical experiments
We have carried out a series of high-resolution gasdynamical simulations of cluster for-
mation in a flat ΛCDM universe (Ωm = 0.3; ΩΛ = 0.7; h = 0.7; σ8 = 0.9; Ωb = 0.02 h−2).
Simulations have been done with the parallel GADGET code (Springel et al. 2001), with
a novel version of SPH in which the entropy is explicitly conserved (Springel & Hernquist
2002). In a cubic volume of 80 h−1 Mpc on a side, an unconstrained realization of the
power spectrum of density fluctuations corresponding to the ΛCDM model was gener-
ated for a total of 10243 Fourier modes. The density field was then resampled to a grid
of 1283 particles, which were displaced from their Lagrangian positions according to the
Zeldovich approximation up to z = 49. Their evolution until the present epoch is traced
by means of a pure N-body simulation with 1283 dark matter particles. A sample of
clusters selected from this preliminary low-resolution experiment were re-simulated with
higher resolution by means of the multiple mass technique (see Klypin et al. 2001, for
details). Mass resolution is then increased by using smaller masses in the Lagrangian
volume depicted by these particles, including the additional small-scale waves from the
ΛCDM power spectrum in the new initial conditions. We use 3 levels of mass refinement,
reaching an effective resolution of 5123 CDM particles (2.96 × 108 h−1 M⊙). Gas was
added in the highest resolved area only. The total number of particles (dark+SPH) in
this area is greater than 1 − 2 × 106 for all clusters. The gravitational softening length
was set to ǫ = 2 − 5 h−1 kpc, depending on number of particles within the virial radius
(Power et al. 2003). The minimum smoothing length for SPH was fixed to the same value
as ǫ.
In order to study effects of resolution in the determination of X-ray properties of our
clusters, we have resimulated one of the objects with 8 times more mass resolution,
reaching an effective resolution of 10243 particles (i.e. mdark ∼ 3 × 107 h−1M⊙; msph ∼
5 × 106M⊙). The total number of particles within the virial radius was 11, 106, 465.
The list of our simulated clusters extracted from the 80 h−1 Mpc volume span a
relative small range in X-ray emission temperature (from 0.6 to 3 keV). In order to extend
our numerical sample of clusters to wider temperature (mass) range, we have simulated
a considerable much bigger volume (500 h−1 Mpc) in which a random realization of
the ΛCDM power spectrum was generated with 20483 particles. In this way, we will
have a similar resolution for our clusters than in the previous experiments. We have
resimulated the whole 500 h−1 Mpc box with different mass resolutions: 2×1283, 2×2563,
and 2 × 5123 dark and SPH particles. We identified all halos in the lowest resolution
run (1283). Using the same technique as before, we selected clusters in this run and
resimulated them at full resolution (20483), using 5 different species of dark particles
and SPH with msph ∼ 1.6 × 108h−1M⊙. The clusters selected cover a range of masses
from 2 × 1014 − 2 × 1015 M⊙. In this regard, we could extend the temperature range of
our cluster sample up to 11 keV. The total number of particles within virial radius in
these clusters is comparable to the number of particles of previous simulations (∼ 106).
We have recently finished the run with 2 × 5123 gas and dark particles that, to our
knowledge, is one of the largest adiabatic SPH simulations of large scale structure done
so far. The mass resolution is mdark ∼ 6 × 1010h−1 M⊙, which means that we can resolve
from galactic halos (100+100 particles) to the biggest galaxy clusters (4 × 106 particles).
We identified a total of 4 × 105 dark matter halos, with 10 or more particles, in this
simulation. Well resolved halos (7000 dark particles or more) correspond to clusters with
Outskirts of Galaxy Clusters: intense life in the suburbs
3
Figure 1. Results from 5003 simulated volume. Left, the X-ray Temperature function and its
observational estimates (Ikebe et al. 2002). Middle, the Mvir − Tx relation. Right, the LX − TX
relation. Stars represent the clusters resimulated with high resolution. The dotted line is a fit
to the hottest halos ( > 6 keV). The slope is ∼ 2, as in the self-similar scaling behaviour
emission weighted temperatures > 3 keV. If we go down to 1000 particles, the halos
resolved have temperatures > 1 keV, and temperatures go down to ∼ 0.6 keV for halos
with 500 or more particles. Due to the large simulated volume, we have a statistical
significant sample of clusters and groups. The total number of objects with TX > 0.6
keV, excluding substructure, is ∼ 30, 000. A total of 126 hot clusters have been found
with TX > 5 keV. The X-ray temperature function for clusters in this simulation is
shown in Figure 1, together with recent observational estimates. As can be seen, the
number density of clusters for the highest temperatures (> 6 keV) found in simulations
is compatible with observations (see Borgani et al. (2004) for results from non-adiabatic
simulations).
The resulting M − TX and LX − TX relations are depicted in Figure 1, in which
we also show 6 of the hotter clusters resimulated with high resolution. They show a
convergence of results at least for clusters with TX > 3 keV. The fit to the M − TX
relation for clusters with TX > 1 keV is (Mvir/M0) = (Tx/keV)α, with α = 1.56 ± 0.05
and M0 = 4.4 × 1013h−1M⊙. The slope is compatible with simulations which include
non adiabatic effects, although the zero point, M0, is a factor of 2 higher. Note however
that we are using virial mass instead of mass for overdensity 500. This implies that the
M-T relation is rather insensitive to the energy transfers due to non adiabatic processes
associated to star-gas interactions.
3. Radial structure of the ICM
We compare the universal halo models described earlier, with results from our simu-
lations. For NFW and MQGSL, we obtain the characteristic density and radius of each
cluster by fitting the dark matter distribution. Gas density and temperature profiles are
genuine predictions of these models. For BM and PBM, we fit the gas distribution and
predict the gas temperature and the dark matter density. In Figure 2 we plot the average
dark matter and gas density profiles. Not surprisingly, both NFW and MQGSL provide
good fits to the dark matter density. These models are able to accurately predict the
gas density, but they are too steep at large radii due to a systematic departure from
hydrostatic equilibrium.
On the other hand, β-models (BM and PBM) show a core in the dark matter density
that is not seen in the numerical data. Moreover, they do not give an accurate description
4
G. Yepes et al.
Figure 2. Universal radial profiles: Upper panel dark matter and gas density. Lower panel:
projected temperature and gas entropy. Black squares represent the numerical data, averaged
over all clusters except major mergers. Error bars denote one-sigma scatter. Lines are used
to plot the analytical models, and top panels quantify deviations from the simulated profile.
Dashed lines in the projected emission-weighted temperature profile, shows the 'universal' profile
proposed by Loken et al. (2002), while solid line represents the best fit to our data. Observations
by De Grandi & Molendi (2002, circles) and Markevitch et al. (1998, boxed region) are ploted
for comparison.
of the gas density profile. The inner regions are better described with low values of β,
while the outer parts require higher values for this parameter.
One of the most striking results from our simulations is that they favour the hy-
pothesis of a 'universal' temperature profile. As can be seen in Figure 2, the ICM is
not isothermal, but the temperature decreases by a factor of three or four from the
centre to the virial radius. The projected X-ray emission-weighted temperature profile
is plotted on the right panel of Figure 2. We compare our simulations with previous
work by Loken et al. (2002) based on Eulerian simulations. These authors propose a
'universal' form Tp(r) = T0(1 + r/ax)−δ Our results are well described by this rela-
tion. Although real clusters seem to be consistent with a polytropic equation of state
(Markevitch et al. 1998), recent observations indicate the presence of a large isothermal
core (De Grandi & Molendi 2002). Apart from this feature, which is not observed in our
objects, adiabatic simulations are in good agreement with observational data beyond
0.2R200.
Entropy profiles are plotted on the lower panel of Figure 2. Contrary to the com-
mon view, neither the analytical models nor the simulation data yield a pure power-
law profile, despite the fact that purely adiabatic gasdynamics has been considered. As
shown in Ascasibar et al. (2003), the standard implementation of the SPH algorithm can
lead to misleading results in the inner regions due to artificial entropy losses (see e.g.
Springel & Hernquist 2002). We find that the shape of the entropy profile does not depend
on the cluster mass or temperature, in agreement with recent observations (Ponman et al.
2003). NFW or MQGSL models provide a better estimate of the entropy profile than the
β-model, but the low gas densities predicted at large radii yield a very steep slope at
r ∼ R200.
Outskirts of Galaxy Clusters: intense life in the suburbs
5
Figure 3. Radial profiles for a cluster run with two different resolutions.
4. Convergence of results
To check for convergence of results in terms of numerical resolution, we compare, in
Figure 3 radial profiles for one of our clusters from the ultra-high resolution simulation
described in previous section. As it can be appreciated, the structure of the ICM is quite
similar, at least from the virial radius down to 1% of that. The features at the most
internal parts of the cluster are related with the different dynamical stage of the cluster.
The positive results of this comparison is that the overall properties of the halos are well
described with the resolution adopted in our simulations in which 500,000-1,000,000 SPH
particles are used.
Acknowledgements
This work has been partially supported by the MCyT (Spain, AYA-0973), by the Ac-
ciones Integradas Hispano-Alemanas HA2000-0026 and by DAAD (Germany). We thank
the Forschungszentrum Julich for allowing us to use the IBM p690+ supercomputer.
References
Ascasibar, Y., Yepes, G., Muller, V., & Gottlober, S. 2003, MNRAS, 346, 731
Borgani, S., Murante, G., Springel, V., et al. 2004, MNRAS, 348, 1078
Cavaliere, A. & Fusco-Femiano, R. 1976, A&A, 49, 137
De Grandi, S. & Molendi, S. 2002, ApJ, 567, 163
Ikebe, Y., Reiprich, T. H., Bohringer, H., Tanaka, Y., & Kitayama, T. 2002, A&A, 383, 773
Klypin, A., Kravtsov, A. V., Bullock, J. S., & Primack, J. R. 2001, ApJ, 554, 903
Lee, J. & Suto, Y. 2003, ApJ, 585, 151
Loken, C., Norman, M. L., Nelson, E., et al. 2002, ApJ, 579, 571
Markevitch, M., Forman, W. R., Sarazin, C. L., & Vikhlinin, A. 1998, ApJ, 503, 77
Moore, B., Quinn, T., Governato, F., Stadel, J., & Lake, G. 1999, MNRAS, 310, 1147
Navarro, J. F., Frenk, C. S., & White, S. D. M. 1997, ApJ, 490, 493
Ponman, T. J., Sanderson, A. J. R., & Finoguenov, A. 2003, MNRAS, 343, 331
Power, C., Navarro, J. F., Jenkins, A., et al. 2003, MNRAS, 338, 14
Rasia, E., Tormen, G., & Moscardini, L. 2003, astroph/0309405
Springel, V. & Hernquist, L. 2002, MNRAS, 333, 649
Springel, V., Yoshida, N., & White, S. D. M. 2001, New Astronomy, 6, 79
|
astro-ph/9810357 | 1 | 9810 | 1998-10-22T09:21:41 | The eclipsing Cataclysmic Variable GS Pavonis: Evidence for disk radius changes | [
"astro-ph"
] | We have obtained differential time series photometry of the cataclysmic variable GS Pavonis over a timespan of 2 years. These show that this system is deeply eclipsing (~2-3.5 mag) with an orbital period of 3.72 hr. The eclipse depth and out-of-eclipse light levels are correlated. From this correlation we deduce that the disk radius is changing and that the eclipses in the low state are total. The derived distance to GS Pav is 790+/-90 pc, with a height above the galactic plane of 420+/-60 pc. We classify GS Pav as a novalike system. | astro-ph | astro-ph |
A&A manuscript no.
(will be inserted by hand later)
Your thesaurus codes are:
06(08.02.2;08.09.2; 08.14.2)
ASTRONOMY
AND
ASTROPHYSICS
The eclipsing Cataclysmic Variable GS Pavonis: Evidence
for disk radius changes ⋆
P.J. Groot1, T. Augusteijn2, O. Barziv1,3, and J. van Paradijs1,4
1 Astronomical Institute 'Anton Pannekoek'/ CHEAF, Kruislaan 403, 1098 SJ, Amsterdam, The Netherlands
2 European Southern Observatory, Casilla 19001, Santiago 19, Chile
3 European Southern Observatory, Karl-Schwarzschildstr. 2, D-85748, Garching-bei-Munchen, Germany
4 Physics Department, University of Alabama in Huntsville, Huntsville, USA
Received date, accepted date
Abstract. We have obtained differential time series pho-
tometry of the cataclysmic variable GS Pavonis over a
timespan of 2 years. These show that this system is deeply
eclipsing (∼2 -- 3.5 mag) with an orbital period of 3.72 hr.
The eclipse depth and out-of-eclipse light levels are corre-
lated. From this correlation we deduce that the disk radius
is changing and that the eclipses in the low state are to-
tal. The derived distance to GS Pav is 790±90 pc, with a
height above the galactic plane of 420±60 pc. We classify
GS Pav as a novalike system.
Key words: Binaries:eclipsing -- stars: GS Pav -- cata-
clysmic variables
Table 1. Log of V-band observations of GS Pav.
Date
Start UT Integr. Time (s) No. Obs.
Sept 5, 1993
Sept 6, 1993
Sept 7, 1993
Sept 8, 1993
March 22, 1995
March 23, 1995
June 13, 1995
July 11, 1995
July 25, 1995
Aug 16, 1995
Sept 15, 1995
Sept 19, 1995
01h50m
23h21m
23h11m
03h57m
08h06m
07h15m
03h20m
05h40m
04h23m
04h52m
03h23m
02h50m
120
120
120
120
240
240
240
240
120
120
240
240
82
66
48
25
25
30
61
65
24
54
21
20
1. Introduction
Eclipsing non-magnetic Cataclysmic Variables (CVs) (for
a review see Warner 1995, hereafter W95) are of particular
interest not only because the masses of both stars can be
determined, but especially because studying their eclipses,
e.g., by the eclipse mapping method (Horne 1985), gives
the opportunity to learn more about the physics of ac-
cretion disks. In this Letter we report that GS Pav is an
eclipsing CV, that shows substantial disk radius changes.
GS Pav was first discovered by Hoffmeister (1963)
who denoted it as star S7040 and gave the comment
'raschwechselnd' (rapidly varying). It was classified as a
dwarf nova type CV in the GCVS and in the catalogue
of Downes, Webbink and Shara (1997), who also give a
finding chart for the object. It was selected for our ob-
servations as a possible member of the halo population
(Augusteijn, 1994). Zwitter and Munari (1995) show that
it has a normal CV spectrum. We determined its location
at RA= 20h08m07.s58, Dec= -- 69◦48′58.′′1 (J2000), almost
identical to that of Downes, Webbink and Shara (1997).
Send offprint requests to: Paul Groot ([email protected])
⋆ Based on observations with the 0.9m Dutch telescope at
ESO, Chile
2. Observations
Photometric observations were obtained with the Dutch
0.9m telescope at ESO La Silla, Chile. A log of the ob-
servations is given in Table 1. All observations were made
with a 512x512 TEK CCD detector, using a Bessel V fil-
ter. Standard flatfielding and debiasing were applied to all
observations. A photometric calibration was obtained on
September 6, 1993 using the standard star EG 21 (Lan-
dolt 1992). Table 2 gives the coordinates and magnitudes
of the reference stars we have used.
3. Photometric ephemeris
Arrival times of mid-eclipse were determined by fitting a
Gaussian profile to the eclipses, and by determining the
mid-point between the points of steepest ascent and de-
scent. The final arrival times listed in Table 3 were taken
as the average of the results from these two methods. We
estimate the accuracy of these arrival times to be 5·10−4
days, which corresponds to the typical difference between
the results from the two methods.
2
P.J. Groot et al.: The Eclipsing Cataclysmic Variable GS Pavonis
Table 2. Reference stars used for the differential photom-
etry of GS Pav.
No. Name
RA (J2000) Dec (J2000) Va
1
2
3
4
5
6
7
8
GAB J200816 -- 6949
GAB J200810 -- 6949
GAB J200804 -- 6949
GAB J200755 -- 6949
GAB J200802 -- 6948
GAB J200800 -- 6948
GAB J200757 -- 6948
GAB J200810 -- 6947
20h08m16.
20h08m10.
20h08m04.
20h07m55.
20h08m02.
20h08m00.
20h07m57.
20h08m10.
s47
s49
s83
s45
s49
s64
s63
s41
-- 69◦49.
-- 69◦49.
-- 69◦49.
-- 69◦49.
-- 69◦48.
-- 69◦48.
-- 69◦48.
-- 69◦47.
′36.
′34.
′40.
′14.
′57.
′49.
′19.
′55.
′′0
′′0
′′4
′′4
′′6
′′5
′′6
′′4
17.31(4)
17.41(4)
17.42(4)
17.09(4)
18.54(7)
16.33(3)
17.87(5)
15.45(2)
Figure 1 shows the phase folded eclipse light curves for
the 12 epochs listed in Table 1. The light curves of June
13 and July 11, 1995 contain two eclipses each.
The width of the eclipse (∆ϕ) can be estimated by
the phases of steepest descent and ascent. For GS Pav we
measure a mean ∆ϕ = 0.064 ± 0.005, where the error
is the scatter on the average of all measurements. From
this value and the range in mass ratio's we deduce (Horne
1985) a range in orbital inclination of 74◦< i < 83◦.
5. Correlation between eclipse depth and
out-of-eclipse light
a The quoted errors reflect the internal errors in the bright-
ness measurements of the stars which do not include a 0.1 mag
uncertainty in the transformation to standard magnitudes.
Table 3. Times of arrival, deduced cycle numbers and the
observed minus computed (O -- C) residuals for the obser-
vations of GS Pav
Cycle No. HJDmin -- 244 0000 O -- C (days)
-- 3062
-- 3050
-- 3044
-- 3043
565
571
1098
1099
1279
1280
1369
1511
1704
1729
9235.7370
9237.6014
9238.5330
9238.6876
9798.9010
9799.8340
9881.6603
9881.8147
9909.7641
9909.9191
9923.7376
9945.7857
9975.7542
9979.6362
-- 0.70·10−3
0.46·10−3
0.44·10−3
-- 0.23·10−3
-- 0.32·10−3
1.05·10−3
0.16·10−3
-- 0.71·10−3
0.13·10−3
-- 0.14·10−3
-- 0.66·10−3
-- 0.87·10−3
0.56·10−3
0.82·10−3
A linear fit to the arrival times listed in Table 3 yields
the following ephemeris:
HJDmin = 244 9711.17388(17) + 0.155269817(87) · N, (1)
with N the cycle number. The error estimates for the pa-
rameters are scaled to give a χ2
red = 1.0. The rms value of
the arrival times around the fit is 6.4·10−4 days, which is
in reasonable agreement with our error estimate.
4. Mass ratio, width of the eclipse and inclination
If the secondary follows the lower main-sequence standard
mass-period relation (W95):
M2 = 0.065 P
5/4
orb (h)
1.3 ≤ Porb(h) ≤ 9,
(2)
with M2 the mass of the secondary, M2 is ∼ 0.34 M⊙.
Since the mass ratio, q=M2/M1, has to be smaller than
2/3 for stable mass transfer (W95) and the mass of the pri-
mary can at most be the Chandrasekhar mass (1.4 M⊙),
the mass ratio is limited to the range 0.67 ≥ q ≥ 0.24.
Figure 1 shows that GS Pav does not have a constant
out-of-eclipse magnitude. In our observations the source
varied between V∼14.9 and V∼ 17.1, being mostly at the
bright end. Also the eclipse depth is not constant. We have
investigated if these two variations are correlated. Since
the eclipses are well represented by Gaussian functions,
we have taken the zero-level and depth of the Gaussians,
used to determine the times of mid-eclipse, as estimates
of the depth in magnitudes of mid-eclipse (∆m) and the
out-of-eclipse light level (mV ) (Fig. 2). The numbers in
Fig. 2 refer to the eclipses as shown in Fig. 1.
In the following we will make a distinction between
the observed radius of the accretion disk, which is the size
we infer from our observations, and the physical radius
of the accretion disk. The difference between these two is
determined by the fractions of the optically thick and thin
parts of the accretion disk and the brightness distribution
across the disk which determines what parts are visible
in the chosen passband (here in the V-band). A changing
brightness distribution can mimic a change in the physical
disk radius, when observed in only one band.
In Fig. 2 the straight line labeled 'Line of Totality'
shows what the correlation between ∆m and mV looks
like for a system in which the eclipse is total. A total
eclipse means that the observed size of the accretion disk
is smaller than the size of the secondary and that at mid-
eclipse the amount of observed light from the disk is neg-
ligible. In a total eclipse the mid-eclipse light level will
be constant and equal to the brightness of the secondary.
Every magnitude of brightening of the disk will cause a
magnitude of deepening of the eclipse: the system will fol-
low a straight line, with an angle of 45◦, in the ∆m-mV
diagram. The position of this line in the diagram will be
different for each individual system, but can be fixed by
determining the brightness of the secondary. If at any time
during our observation GS Pas was totally eclipsing, then
its mid-eclipse light level will be the brightness of the sec-
ondary. The minimum level occured on June 13 and July
11, 1995: V =19.9±0.1. We use this point to fix the position
of the 'Line of Totality' in Fig. 2. We see that points '7'
and '8' (which are from June 13 and July 11, 1995) lie on
this line. From the fact that the eclipse depths in point '7'
and '8' are different, but the brightness at mid-eclipse, we
P.J. Groot et al.: The Eclipsing Cataclysmic Variable GS Pavonis
3
Fig. 1. The phasefolded V-band eclipse light curves for the 12 epochs. The heliocentric corrected time of mid-eclipse
of the observations is given in UT.
conclude that at these epochs the eclipse is indeed total. It
follows that the observed minimum of V =19.9±0.1 is the
brightness of the secondary. The lack of a flat-bottom in
the light curves of point '7' and '8' shows that the eclipse is
only just total, although the integration time of 4 minutes
may be too long to resolve a flat bottom.
If we now look at the other points in Fig. 2 we see
that they do not fall on the 'Line of Totality'. With in-
creasing out-of-eclipse light levels, the depth of the eclipse
does not increase anymore (as it would have on the Line
of Totality), but decreases. Apparently, between point '7'
and '1' on the track the observed size of the disk increases
to the extent that the eclipse is no longer total, but that
part of the accretion disk remains visible in mid-eclipse.
With a further increase of the observed accretion disk ra-
dius, more and more of the disk is visible at mid-eclipse
and the eclipse becomes less and less deep. This behaviour
was first found by Walker (1963) in his study of RW Tri,
which shows no total eclipses, but does move back and
forth on this part of the track. We have therefore labeled
this the 'Walker Branch'. Another transition, which to our
knowledge has never been noted before, is the one that
happens near point '10' in Fig. 2. The out-of-eclipse light
level reaches a maximum, after which it declines again
(the curve bends to the right), but the eclipse depth con-
tinues to decrease. We have labeled this part the 'Shallow
Branch' because of its decreasing eclipse depth.
If the change in the observed radius is caused by a
change in the physical size of the disk (rather than a
change in the brightness distribution), then the manifesta-
tion of the Shallow Branch may be explained by the effect
of self-eclipses. Because the height of the disk is corre-
lated with its physical size (Frank, King and Raine 1985),
self-eclipses of the hot, and therefore luminous inner parts
of the concave disk by its outer parts, will occur when
the physical size of the disk exceeds a critical value and
therefore a critical height. The out-of-eclipse light level
decreases because the more luminous parts of the accre-
tion disk are self-eclipsed, and at the same time the eclipse
depth can continue to decrease if the disk radius continues
to increase. To eclipse the inner parts the disk flaring an-
gle must be (90-i)◦ or higher. In our case this would mean
a flaring angle of 7◦ -- 16◦, similar to what has been found
in other CVs (e.g. Robinson et al, 1995).
The variations appear to trace out a unique track over
a substantial period of time. Changes from one part of the
track to another can occur quite rapidly, e.g. the transition
from point '1' to '4' took place within 4 days. We are
P.J. Groot et al.: The Eclipsing Cataclysmic Variable GS Pavonis
4
∆ m
5
4
3
2
1
Line of Totality
9
6
1
7
8
Walker Branch
10
5
2
11
4
12
3
Shallow Branch
systems having low states in their long term light curves,
and showing no DN outbursts during these low states, we
should also classify it as a VY Scl star. This is supported
by its orbital period, since almost all known VY Scl stars
have periods between 3 and 4 hrs. However, the physical
interpretation as outlined in W95 may not apply to GS
Pav. In this description a VY Scl system in its low state
M , that is lower than the crit-
has a mass-transfer rate
Mcrit, below which DN outbursts are expected
ical rate,
to occur, but it does not show these outbursts. This dis-
tinguishes VY Scl systems from Z Cam systems, that do
show these outbursts in their low state. In our observations
GS Pav at all times seems to have an absolute magnitude
which was brighter or equal than that of Z Cam systems
M is thought to be larger than
during standstill, where
Mcrit. We therefore cannot conclude if GS Pav is a VY
Scl system or not. However, it could be that all NL sys-
tems with periods between 3 and 4 hours turn out to have
low and high states if sufficiently long observed (W95).
An interesting example of a system that may be anal-
ogous to GS Pav is VZ Scl for which O'Donoghue, Fairall
and Warner (1987) concluded that the size of the accre-
tion disk has changed from one observation to the other.
Unfortunately they only observed two eclipses. Since this
system, in the low state, is also totally eclipsing, a com-
parison with GS Pav would be of interest.
15
16
17
m V
18
19
20
Fig. 2. The depth of the eclipse in GS Pav as function
of the out-of-eclipse light. The numbers correspond to the
eclipses in Fig. 1.
currently modelling these changes in detail to constrain
the geometry of the system and the changing disk (Groot
et al., 1999).
6. Distance to the system
8. Conclusions
With the orbital period of 3.72 hr we can use the MV−Porb
relation of W95 to derive a MV = 10.4 with an esti-
mated error of 0.2 magnitudes for the secondary (from Fig.
2.46 in W95). Combined with the apparent magnitude of
V =19.9±0.1, this gives a distance to GS Pav of 790±90pc.
Given its position on the sky, the implied height above the
galactic plane is 420±60 pc. From the distance we can de-
duce the absolute magnitude of the system out of eclipse,
which is dominated by the accretion disk. To obtain an
absolute magnitude of the accretion disk, we correct for
the inclination of the disk, according to Eq. 2.63 of W95.
This correction varies from 1.0 mag (for i=74◦) to 2.1 mag
(for i=83◦). The absolute magnitude (MV) lies therefore
between 6.6 ≥ MV ≥ 4.4 (for i=74◦) and 5.5 ≥ MV ≥ 3.3
(for i=83◦). Comparison with the mean absolute magni-
tudes for NLs and DNe (Fig. 4.16 and 3.9 from W95)
shows that the derived range is in the normal regime for
NLs, but too bright for DN in quiescence.
7. Classification as a novalike system
From the shape of the light curve and its absolute magni-
tude, we conclude that GS Pav is a novalike system, and
considering its emission line spectrum (Zwitter and Mu-
nari, 1995), that it is of the RW Tri subclass, which is
defined as having emission line spectra (W95). According
to the definition in W95 of the VY Scl subclass, as NL
We have shown that GS Pav is a deeply eclipsing cata-
clysmic variable with a 3.72 hr period. Based on its photo-
metric behaviour, orbital period and absolute magnitude
we classify the system as a novalike variable. The depth of
the eclipse and the brightness of the system out-of-eclipse
are correlated. From this relation we infer that the disk
radius changes and that in two of our observations the
eclipse is total. Therefore, the apparent visual magnitude
of the secondary is V =19.9±0.1, giving a distance to the
system of 790±90 pc.
References
Augusteijn, T., 1994, PhD thesis, University of Amsterdam
Downes, R, Webbink, R.F and Shara, M.M., 1997, PASP 109,
345
Frank, J., King, A.R. and Raine, D.J., 1985, Accretion Power
in Astrophysics, Cambridge Astrophysics Series Vol. 21,
CUP
Groot, P.J., Heemskerk, M.H.M., Augusteijn, T. and Van
Paradijs, J., 1999, in preparation
Hoffmeister, C., 1963, Veroff. Sternwarte Sonneberg, 6, 1
Horne, K., 1985, MNRAS 213, 129
de Jong, J.A., Van Paradijs, J. and Augusteijn, T., A&A 314,
484
Landolt, A., 1992, AJ 104, 372
O'Donoghue, D., A.P. Fairall and Warner, B., 1987, MNRAS
225, 43
Ritter, H. and Kolb, U., 1998, A&AS 129, 83
P.J. Groot et al.: The Eclipsing Cataclysmic Variable GS Pavonis
5
Robinson, E.L., Wood, J.H, Bless, R.C. et al., 1995, ApJ 443,
295
Warner, B., 1995, Cataclysmic Variables, Cambridge Astro-
physics Series Vol. 28, CUP.
Walker, M.F., 1963, ApJ 138, 146
Zwitter, T. and Munari, U., 1995, A&AS 114, 575
|
astro-ph/0402422 | 1 | 0402 | 2004-02-18T03:38:36 | RR Lyrae Stars in the Andromeda Halo from Deep Imaging with the Advanced Camera for Surveys | [
"astro-ph"
] | We present a complete census of RR Lyrae stars in a halo field of the Andromeda galaxy. These deep observations, taken as part of a program to measure the star formation history in the halo, spanned a period of 41 days with sampling on a variety of time scales, enabling the identification of short and long period variables. Although the long period variables cannot be fully characterized within the time span of this program, the enormous advance in sensitivity provided by the Advanced Camera for Surveys on the Hubble Space Telescope allows accurate characterization of the RR Lyrae population in this field. We find 29 RRab stars with a mean period of 0.594 days, 25 RRc stars with a mean period of 0.316 days, and 1 RRd star with a fundamental period of 0.473 days and a first overtone period of 0.353 days. These 55 RR Lyrae stars imply a specific frequency S_RR=5.6, which is large given the high mean metallicity of the halo, but not surprising given that these stars arise from the old, metal-poor tail of the distribution. This old population in the Andromeda halo cannot be clearly placed into one of the Oosterhoff types: the ratio of RRc/RRabc stars is within the range seen in Oosterhoff II globular clusters, the mean RRab period is in the gap between Oosterhoff types, and the mean RRc period is in the range seen in Oosterhoff I globular clusters. The periods of these RR Lyraes suggest a mean metallicity of [Fe/H]=-1.6, while their brightness implies a distance modulus to Andromeda of 24.5+/-0.1, in good agreement with the Cepheid distance. | astro-ph | astro-ph | TO APPEAR IN THE ASTRONOMICAL JOURNAL
Preprint typeset using LATEX style emulateapj v. 04/03/99
4
0
0
2
b
e
F
8
1
1
v
2
2
4
2
0
4
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
RR LYRAE STARS IN THE ANDROMEDA HALO FROM DEEP IMAGING WITH THE ADVANCED
CAMERA FOR SURVEYS1
Space Telescope Science Institute, 3700 San Martin Drive, Baltimore, MD 21218; [email protected], [email protected],
[email protected]
THOMAS M. BROWN, HENRY C. FERGUSON, ED SMITH
RANDY A. KIMBLE, ALLEN V. SWEIGART
Code 681, NASA Goddard Space Flight Center, Greenbelt, MD 20771; [email protected], [email protected]
European Southern Observatory, Karl-Schwarzschild-Strasse 2, Garching bei Munchen, Germany; [email protected]
Division of Astronomy, Dpt. of Physics & Astronomy, UCLA, Los Angeles, CA 90095; [email protected]
R. MICHAEL RICH
ALVIO RENZINI
To appear in The Astronomical Journal
ABSTRACT
We present a complete census of RR Lyrae stars in a halo field of the Andromeda galaxy. These deep observa-
tions, taken as part of a program to measure the star formation history in the halo, spanned a period of 41 days with
sampling on a variety of time scales, enabling the identification of short and long period variables. Although the
long period variables cannot be fully characterized within the time span of this program, the enormous advance in
sensitivity provided by the Advanced Camera for Surveys on the Hubble Space Telescope allows accurate char-
acterization of the RR Lyrae population in this field. We find 29 RRab stars with a mean period of 0.594 days, 25
RRc stars with a mean period of 0.316 days, and 1 RRd star with a fundamental period of 0.473 days and a first
overtone period of 0.353 days. These 55 RR Lyrae stars imply a specific frequency SRR ≈ 5.6, which is large given
the high mean metallicity of the halo, but not surprising given that these stars arise from the old, metal-poor tail
of the distribution. This old population in the Andromeda halo cannot be clearly placed into one of the Oosterhoff
types: the ratio of RRc/RRabc stars is within the range seen in Oosterhoff II globular clusters, the mean RRab
period is in the gap between Oosterhoff types, and the mean RRc period is in the range seen in Oosterhoff I glob-
ular clusters. The periods of these RR Lyraes suggest a mean metallicity of [Fe/H]≈ −1.6, while their brightness
implies a distance modulus to Andromeda of 24.5±0.1, in good agreement with the Cepheid distance.
Subject headings: galaxies: evolution -- galaxies: stellar content -- galaxies: halos -- galaxies: individual (M31) --
stars: variables
1. INTRODUCTION
The textbook picture of a spiral galaxy halo comes from that
of our own Milky Way, which is old and metal-poor (Vanden-
Berg 2000; Ryan & Norris 1991). However, the stellar popu-
lation of the Andromeda (M31; NGC224) halo offers a strik-
ing contrast to this picture, with its wide range in metallicity
(Durrell, Harris, & Pritchet 2001) and age (Brown et al. 2003;
Brown 2003). These spreads in metallicity and age can sig-
nificantly affect the variable star population. For example, the
characteristics of RR Lyraes in Galactic globular clusters place
these clusters into two distinct Oosterhoff types (Oosterhoff
1939), while the RR Lyraes in Local Group dwarf spheroidals
(dSphs) place these galaxies in the gap between the two Ooster-
hoff types (Siegel & Majewski 2000; Dall'Ora et al. 2003; Pritzl
et al. 2002, 2004). Compared to the M31 halo, dSphs have an
even broader age range (van den Bergh 1999) and are generally
more metal-poor (Mateo 1998), but the old (> 10 Gyr) compo-
nent capable of producing RR Lyrae stars might be similar in
each case.
The specific frequency of RR Lyraes in the M31 halo has
been the subject of some debate. In a field 40 arcmin from the
1Based on observations made with the NASA/ESA Hubble Space Tele-
scope, obtained at the Space Telescope Science Institute, which is op-
erated by AURA, Inc., under NASA contract NAS 5-26555. These
observations are associated with proposal 9453.
nucleus on the southeast minor axis, Pritchet & van den Bergh
(1987) found 30 RR Lyraes, and with an estimated complete-
ness of 25%, determined that the frequency per unit luminosity
was very high (about half of that in variable-rich M3). More
recently, Dolphin et al. (2003) found only 24 RR Lyraes in a
larger field that included the Pritchet & van den Bergh (1987)
field, and with their estimated completeness of 24%, claimed
that the frequency of RR Lyrae was ∼15 times smaller. To sup-
port their claim, they estimated that the deep color magnitude
diagram (CMD) of Brown et al. (2003) contained only 10 RR
Lyraes, but as we shall show here, this was a severe underesti-
mate.
We have observed a field along the southeast minor axis of
the M31 halo, 51 arcmin from the nucleus, using the Advanced
Camera for Surveys (ACS; Ford et al. 1998) on the Hubble
Space Telescope (HST). The primary goal of this program was
to investigate the halo star formation history, by constructing
a deep CMD in the F606W (broad V ) and F814W (I) band-
passes, reaching V ≈ 30.7 mag on the main sequence (Brown
et al. 2003). However, the 250 individual exposures are scat-
tered with variable time sampling over a 41 day period, and
thus provide excellent time series photometry for the variable
star population in the M31 halo; in particular, the completeness
for RR Lyraes in our field is approximately 100%. In this pa-
per, we present a survey of the RR Lyraes and the other bright
variables in our field.
1
2
2. OBSERVATIONS AND DATA REDUCTION
Using the Wide Field Camera on the ACS, we obtained deep
optical images of a 3.5′ × 3.7′ field along the southeast minor
axis of the M31 halo, at α2000 = 00h46m07s, δ2000 = 40o42′34′′.
The surface brightness in this region is µV ≈ 26.3 mag arcsec−2
(Brown et al. 2003). The area is not associated with the tidal
streams and substructure found by Ferguson et al. (2002), and
lies just outside the "flattened inner halo" in their maps. We
placed a metal-rich M31 globular cluster, GC312 (Sargent et
al. 1977), near the edge of the field. From 2 Dec 2002 to 11
Jan 2003, we obtained 39.1 hours of images in the F606W fil-
ter (broad V ) and 45.4 hours in the F814W filter (I), with each
of the 250 exposures dithered to allow for hot pixel removal,
optimal point spread function sampling, smoothing of spatial
variations in detector response, and filling in the gap between
the two halves of the 4096 × 4096 pixel detector. Of the 250
exposures, 16 were short (< 600 s), to allow the correction of
bright saturated objects, leaving 234 long exposures (> 1200 s)
suitable for deep time-series photometry.
In order to create a deep catalog of the objects in this field,
we first co-added these images using the IRAF DRIZZLE pack-
age. This coaddition included masks for the cosmic rays and
hot pixels, and produced geometrically-correct images with
a plate scale of 0.03′′ pixel−1 and an area of approximately
210′′ × 220′′. The mask for each exposure was created in an
iterative process, comparing the value in every pixel to the dis-
tribution through the entire stack at that location on the sky;
pixels are masked on the bright end of the distribution (cosmic
rays and hot pixels) and the faint end of the distribution (dead
pixels), but this technique also masks bright, large-amplitude,
variable stars when they are near maximum or minimum. The
final coadded image in each bandpass thus approximates the av-
erage flux observed for the variable stars, but these masks can-
not be used in the time series photometry (discussed below).
We then performed both aperture and PSF-fitting photom-
etry using the DAOPHOT-II package (Stetson 1987), assum-
ing a variable PSF constructed from the most isolated stars.
The aperture photometry on isolated stars was corrected to true
apparent magnitudes using TinyTim models of the HST PSF
(Krist 1995) and observations of the standard star EGGR 102
(a V = 12.8 mag DA white dwarf) in the same filters, with
agreement at the 1% level. The PSF-fitting photometry was
then compared to the corrected aperture photometry, in order to
derive the offset between the PSF-fitting photometry and true
apparent magnitudes. Our photometry is in the STMAG sys-
tem: m = −2.5× log10 fλ − 21.1. For readers more familiar
with the Johnson V and Cousins I bandpasses, a star in the
middle of the RR Lyrae strip has V − mF606W = −0.17 mag
and I − mF814W = −1.29 mag. Light curve amplitudes are typ-
ically 6 -- 10% smaller in mF606W than in V , and 1 -- 2% smaller
in mF814W than in I. The transformation between the ACS and
ground-based bandpasses is still being characterized indepen-
dently by several groups, so for most of this paper we will refer
to magnitudes in the unambiguous STMAG system.
The CMD for the co-added images was shown by Brown et
al. (2003). In that analysis, we discarded ≈20% of the exposed
area (around bright foreground stars, near GC312, and in re-
gions with less than the full exposure time due to the dither
pattern). In the current analysis of the brighter stars, we include
the entire image area. We will only consider stars brighter than
mF814W = 28.25 mag, which have a signal-to-noise ratio of ∼ 5
in individual exposures. Extensive artificial star tests demon-
strate that our catalog, created from the deep co-added data, is
≈100% complete above this limit.
To obtain time-series photometry, we re-drizzled the entire
dataset into a stack of individual registered exposures, with
masked pixels set to an invalid data value. Because our orig-
inal masks sometimes discard data points near the maxima and
minima of variable stars, we unmasked those pixels that were
masked more than once in a two-hour window; this correction
will occasionally restore a true cosmic ray or dead pixel when
it should have been masked, but these occasional events will be
significant outliers in the time series photometry, which can be
discarded after the fact. We then performed aperture photome-
try on each individual frame, with positions fixed by the catalog
of the coadded exposures.
3. VARIABLE DETECTION AND CHARACTERIZATION
As explained above, we restricted our time-series analysis
to stars brighter than mF814W = 28.25 mag; out of the nearly
300,000 stars in the full image catalog, these 19,450 stars have
a signal-to-noise of at least 5 in individual exposures. Stars in
the RR Lyrae gap, near mF814W ≈ 26 mag, have typical photo-
metric errors of 0.03 mag in F606W exposures and 0.04 mag
in F814W exposures. Most of the stars in our search list have
∼ 100 photometric measurements in each bandpass over our 41
day observing program, but a small fraction have significantly
less because they fell near the edges of the two halves of the
detector; 424 stars, representing only 2% of the stars in the rel-
evant brightness range, were discarded because they had less
than 30 valid measurements in each bandpass. For the remain-
ing stars, we looked for variability using two methods, chosen
to suit the high signal-to-noise and large number of time sam-
ples in these data.
The first variability search was tuned to provide a complete
sample of RR Lyraes, but it also recovered a large number of
brighter, long period variables and fainter, short period vari-
ables. We used a fast algorithm (Press & Rybicki 1989) of
the Lomb-Scargle periodogram (Lomb 1976; Scargle 1982),
which looks for weak periodic signals in irregularly sampled
data. Besides providing a good initial estimate of the period,
this algorithm also quantifies the statistical significance of the
periodic signal. Our threshold was a non-random signal at 0.01
significance, independently found in each bandpass; a score
of 0.01 implies that the chance this signal arose from random
fluctuations is less than 1 percent. This stringent threshold
should only produce one or two false detections in the entire
search list. However, this threshold is not so stringent that
it misses RR Lyraes; all of the RR Lyraes recovered by this
method, regardless of amplitude or period, were detected at a
significance orders of magnitude beyond this threshold (if our
choice of threshold were in fact discarding RR Lyraes, then the
distribution of RR Lyrae scores would approach the threshold).
The initial period estimate returned by this method was then
refined by a search of the Lafler & Kinman (1965) statistic, for
periods within 5% of the initial period. Our method recovered
169 variables, two of which were false detections (each a faint
star sitting in the wings of a bright variable), and 55 of which
are clearly RR Lyraes. Of the remaining variables, 17 are short
period variables below the horizontal branch (HB), 3 have peri-
ods of 0.6 -- 7 days and lie above the HB, 1 is an eclipsing binary,
82 are long period variables (LPVs) and semiregulars near the
tip of the red giant branch (RGB), 5 are LPVs fainter than the
HB, and 4 are variables with periods of ∼1 week that lie on
the RGB. Thirteen of the detected variables, all with periods
V130 P=0.263
V89 P=0.267
V76 P=0.274
V100 P=0.275
3
mF606W
25.0
25.5
26.0
mF814W
26.5
mF606W
25.0
25.5
26.0
mF814W
26.5
mF606W
25.0
25.5
26.0
mF814W
26.5
mF606W
25.0
25.5
26.0
mF814W
26.5
mF606W
25.0
25.5
26.0
mF814W
26.5
mF606W
25.0
25.5
26.0
mF814W
26.5
mF606W
25.0
25.5
26.0
mF814W
26.5
V137 P=0.280
V37 P=0.301
V80 P=0.313
V11 P=0.331
V83 P=0.351
V54 P=0.366
V120 P=0.283
V102 P=0.301
V131 P=0.327
V43 P=0.338
V90 P=0.353
V1 P=0.382
V27 P=0.287
V8 P=0.306
V157 P=0.329
V59 P=0.339
V95 P=0.361
V147 P=0.442
V40 P=0.289
V163 P=0.313
V161 P=0.330
V5 P=0.339
V50 P=0.366
V44 P=0.464
0.5
1.0
0.5
1.0
Phase
0.5
1.0
0.5
1.0
FIG. 1 -- The light curves for RR Lyraes in our field, arranged in order of increasing period (labeled), shown as photometric error
bars (not data points). The first 26 stars are RRc and RRd stars (V90 being the RRd, shown phased with its first overtone period);
the remaining 29 are RRab stars.
longer than 20 days, fall within the tidal radius of GC312 (10′′;
Holland et al. 1997); because the area within the tidal radius of
GC312 comprises less than 1 percent of our total field, these 13
LPVs and semiregulars are clearly associated with GC312.
The second variability search simply looked for photome-
try that exceeded the expected scatter (given the photometric
errors) by 50%. This method was motivated by a search for
halo-on-halo microlensing events. We found no microlensing
events or new pulsating variables, but we did find 7 additional
eclipsing binaries. As might be expected, the Lomb-Scargle
periodogram is better suited to finding variables with roughly
sinusoidal light curves. We show the light curves for the RR
4
mF606W
25.0
25.5
26.0
mF814W
26.5
mF606W
25.0
25.5
26.0
mF814W
26.5
mF606W
25.0
25.5
26.0
mF814W
26.5
mF606W
25.0
25.5
26.0
mF814W
26.5
mF606W
25.0
25.5
26.0
mF814W
26.5
mF606W
25.0
25.5
26.0
mF814W
26.5
mF606W
25.0
25.5
26.0
mF814W
26.5
V47 P=0.496
V88 P=0.506
V79 P=0.529
V28 P=0.532
V162 P=0.533
V142 P=0.554
V164 P=0.580
V167 P=0.621
V136 P=0.634
V114 P=0.726
V126 P=0.534
V140 P=0.559
V166 P=0.583
V124 P=0.627
V77 P=0.678
V82 P=0.735
V42 P=0.552
V133 P=0.572
V122 P=0.589
V36 P=0.627
V10 P=0.687
V57 P=0.554
V112 P=0.573
V160 P=0.611
V123 P=0.631
V66 P=0.714
V78 P=0.774
0.5
1.0
0.5
1.0
Phase
0.5
1.0
FIG. 1 -- Continued.
Lyraes in Figure 1, light curves for the variables above and be-
low the HB in Figure 2, light curves for the eclipsing binaries in
Figure 3, and light curves for a subset of the LPVs and semireg-
ulars in Figure 4 (only those that varied by more than ∼ 0.1 mag
within the 41 days of observations). Note that only photometric
error bars are visible at the plotted scale of these light curves.
Five of the RR Lyrae light curves show significant scatter
(V1, V90, V95, V122, and V163). We reevaluated the peri-
odicity searches on these five stars, and found that one of the
stars, V90, showed clear evidence for double-mode pulsations,
making it a definite RRd star, with a fundamental period (P0)
of 0.4735 days and a first overtone period (P1) of 0.3534 days.
V163 shows a hint of an additional period at 0.4243 days,
but the signal is weak, and we found no additional periods
22.0
V84 P=0.611
21.0
V118 P=2.010
22.0
V60 P=7.135
5
mF606W
mF814W
mF606W
mF814W
mF606W
mF814W
mF606W
mF814W
mF606W
mF814W
mF606W
mF814W
mF606W
mF814W
24.5
25.0
27.5
26.5
29.0
26.0
28.5
26.0
28.5
26.0
28.5
24.5
27.0
23.5
25.5
28.0
26.5
29.0
26.0
28.5
26.0
28.5
25.0
27.5
26.5
29.0
V48 P=0.112
V105 P=0.115
V119 P=0.226
24.5
26.5
29.0
24.5
27.0
26.0
28.5
25.5
28.0
26.0
28.5
V113 P=0.230
V94 P=0.252
V6 P=0.267
V65 P=0.309
V12 P=0.317
V93 P=0.338
V63 P=0.351
V128 P=0.353
V91 P=0.367
V99 P=0.390
V85 P=0.407
V74 P=0.439
V29 P=0.573
V117 P=0.680
0.5
1.0
0.5
1.0
Phase
FIG. 2 -- The light curves for other variables bluer than the RGB (excluding eclipsing binaries), shown as error bars. The stars in
the top 3 panels are brighter than the HB; they are arranged in order of increasing period and increasing mF606W − mF814W color,
making them easy to identify in the CMD (Figure 5). The stars in the lower 17 panels are likely dwarf Cepheids fainter than the
HB, but V6 and V29, which lie immediately below the HB, might be RR Lyraes.
in the light curves of V1 and V95. Although V1, V95, and
V163 might be RRd stars, we have classified them as RRc. The
scatter in the light curve of V122 might be due to the Blazhko
effect -- a secondary modulation in the variability of 20 -- 30% of
RRab stars, with a period of ∼10 -- 500 days (Smith et al. 2003).
To determine the mean magnitudes and amplitudes of the
variables with periods less than ∼ 10 days, we fit a Fourier se-
ries to the magnitudes in each bandpass as a function of phase,
with an order of 1 or 2 for the sinusoidal light curves (e.g.,
RRcd stars) and an order of 8 for the sawtooth light curves (e.g.,
6
26.5
V168 P=0.811
26
V172 P=0.913
26.5
V170 P=0.976
25
V173 P=1.245
mF606W
mF814W
29.5
25.5
mF606W
mF814W
28.5
V169 P=1.550
V171 P=1.806
29
25.5
28.5
0.5 1.0
0.5 1.0
29.5
25.5
28.5
Phase
V174 P=2.254
V64 P=3.037
28
25
28
0.5 1.0
0.5 1.0
FIG. 3 -- The light curves for candidate eclipsing binaries in our field. The most secure (V64) was found through the Lomb-Scargle
periodogram search, while the others were found because their photometric errors were larger than expected. Note that a secondary
minimum is sometimes visible at a phase opposite to the primary minimum.
RRab stars). Table 1 lists the properties of the RR Lyraes in
our field, and Table 2 gives their time-series photometry (with
the full dataset available only as a machine-readable table in the
electronic version of this paper). Table 3 gives the positions of
the other variables in our field (Figures 2 -- 4). Figure 5 shows
the CMD locations of the stars with light curves (Figures 1 -- 4).
The variables are shown at their mean mF606W − mF814W color
and mF814W magnitude, as determined from the fitting above,
instead of the average observed value in the catalog from the
deep co-added images, which can be systematically brighter or
fainter in a given bandpass by a few hundredths of a magnitude,
depending upon the random sampling of each light curve. Fig-
ure 6 shows the distribution of mF606W amplitude vs. period,
mean mF606W vs. period, and the period distribution for the RR
Lyrae stars. The RR Lyraes are cleanly separated from the re-
maining variables by their position in the mF606W vs. period
diagram, although two stars lying immediately below the HB
might also be RR Lyraes (V6 and V29 in Figure 2). The RR
Lyraes can be classified as either RRab or RRcd by the gap in
amplitude vs. period distribution; these classes do not overlap
in the CMD (Figure 5).
4. RR LYRAE PROPERTIES
4.1. Oosterhoff Type
The RR Lyrae population of the M31 halo cannot be classi-
fied into either of the Oosterhoff (1939) types, but this is not
because the various characteristics appear as an average of the
two types (Figure 7). Looking at the ensemble variable popu-
lation in Galactic globular clusters, the ratio of RRc to RRabc
stars is 0.22 in Oosterhoff I clusters and 0.48 in Oosterhoff II
clusters (Clement et al. 2001), while in the M31 halo it is 0.46
-- a value in the Oosterhoff II regime. In contrast, looking at
the mean period of the RRc stars in Galactic globular clusters,
it is 0.326 days in Oosterhoff I clusters and 0.368 days in Oost-
erhoff II clusters (Clement et al. 2001), while in the M31 halo
it is 0.316 days, within the Oosterhoff I regime. The mean pe-
riod for the known RRab stars in Oosterhoff I clusters is 0.559
days, and in Oosterhoff II clusters it is 0.659 days (Clement
et al. 2001), while in the M31 halo it is 0.594 days -- midway
between the two types. The RRab period distribution is more
sharply peaked at this intermediate value than one might ex-
pect from an equal mix of Oosterhoff types (see Figure 8), but
a Kolmogorov-Smirnov test shows the difference is not statis-
tically significant. The period-amplitude diagram for the RRab
stars is also suggestive of an intermediate Oosterhoff type (Fig-
ure 9). If we convert our mF606W amplitudes to V amplitudes,
and compare against the period-amplitude relations of Ooster-
hoff I and Oosterhoff II clusters (Clement 2000), the RRab stars
in M31 fall between the two types, although there is a tendency
toward Oosterhoff I at the larger amplitudes.
Although the Galactic halo field is more difficult to charac-
terize than the Galactic clusters, it is interesting to note that the
field RR Lyraes also show the Oosterhoff dichotomy, with a real
gap in the distribution of period shifts at fixed amplitude with
respect to the M3 RR Lyraes (see, e.g., Figure 8 of Suntzeff,
Kinman, & Kraft 1991). However, these same data show that
the Galactic field population is predominantly of Oosterhoff
type I, as has been known for some time. In their study of the
Palomar-Groningen Survey, Cacciari & Renzini (1976) found
<Pab>= 0.529 days, <Pc>= 0.329 days, and Nc/Nabc = 0.09.
Selection effects in a field survey might favor RR Lyraes with
higher amplitudes (thus reducing the relative number of RRc
stars and long-period RRab stars), but these characteristics are
all well within the Oosterhoff I regime. In contrast, the cor-
responding values for the M31 halo RR Lyraes are quite dif-
ferent. In particular, the mean RRab period is much longer in
M31 (<Pab>= 0.594 days), and the frequency of RRc variables
is much higher (Nc/Nabc = 0.46). Most importantly, the distri-
bution of period shifts of the M31 RR Lyraes with respect to
the M3 RR Lyraes (i.e., the Oosterhoff I line in Figure 9) shows
no such gap. All of these characteristics clearly distinguish the
M31 halo field from the Milky Way halo field.
The characteristics of the M31 halo variables do not track
those of the Local Group dSphs, either, except in the broad
sense that the M31 halo cannot be put cleanly into one of
the Oosterhoff types (Figure 7). For dSphs, the mean period
of RRab stars tends to fall in the gap between Oosterhoff types
7
FIG. 4 -- The light curves for a subset of the bright LPVs and semiregulars in our field (those that varied by more than 0.1 mag in
each bandpass), arranged roughly in order of increasing period (the 41 day span of our observations limits our ability to characterize
the periods).
(Siegel & Majewski 2000; Pritzl et al. 2002, 2004; Dall'Ora et
al. 2003 and references therein), as found in M31. However, the
fraction of RR Lyrae stars that are RRc is higher in M31 (typi-
cal of Oosterhoff II) than that in any of the dSphs (which tend
toward the Oosterhoff I values), while the mean period of RRc
stars is much lower in M31 (typical of Oosterhoff I) than in any
of the dSphs (which tend to fall in the gap between Oosterhoff
types). These differences might suggest that the M31 halo is
not comprised of dissolved globular clusters like those in the
Milky Way, or dissolved Local Group dSphs. Note that globular
clusters in the Large Magellanic Cloud are also predominantly
of intermediate Oosterhoff type, in contrast to the clusters of the
8
FIG. 4 -- Continued.
Milky Way (Bono, Caputo, & Stellingwerf 1994).
It has been argued that the Oosterhoff dichotomy among
Galactic globular clusters is due to a real gap in metallicity
between the two Oosterhoff types, with clusters at metallici-
ties between the two Oosterhoff types having very blue HBs
that do not populate the RR Lyrae instability strip (Renzini
1983; Sandage 1993). Local Group dSphs and LMC globular
clusters fill the gap in a plot of [Fe/H] vs. <Pab>, as illustrated
by Siegel & Majewski (2000) and Pritzl et al. (2004) (see Fig-
ure 6 in each paper). This indicates that the dSphs and Galactic
globular clusters occupy separate regions of the age vs. chem-
ical composition parameter space (Renzini 1980), which in the
simplest option reduces to the [Fe/H], age, and [α/Fe] three-
dimensional space. The M31 halo appears to bridge the gap
in a manner similar to the dSphs, signaling that the early star
formation history in the M31 halo proceeded at a different pace
Name
V1
V5
V8
V10
V11
V27
V28
V36
V37
V40
V42
V43
V44
V47
V50
V54
V57
V59
V66
V76
V77
V78
V79
V80
V82
V83
V88
V89
V90
V95
V100
V102
V112
V114
V120
V122
V123
V124
V126
V130
V131
V133
V136
V137
V140
V142
V147
V157
V160
V161
V162
V163
V164
V166
V167
R.A.
(J2000)
0h45m59.76s
0h46m 0.61s
0h46m 1.02s
0h46m 0.68s
0h45m56.88s
0h45m57.64s
0h46m 0.08s
0h46m 2.04s
0h46m 3.26s
0h46m 1.20s
0h46m 0.77s
0h46m 1.74s
0h45m59.93s
0h46m 1.55s
0h46m 2.68s
0h46m 4.33s
0h46m 2.60s
0h46m 3.80s
0h46m 4.71s
0h46m 5.86s
0h46m 4.22s
0h46m 8.34s
0h46m 4.60s
0h46m 5.78s
0h46m 5.29s
0h46m 8.84s
0h46m 9.86s
0h46m 6.68s
0h46m 7.71s
0h46m 9.60s
0h46m 6.94s
0h46m 8.57s
0h46m11.88s
0h46m10.11s
0h46m12.78s
0h46m13.10s
0h46m12.46s
0h46m 9.88s
0h46m10.38s
0h46m11.91s
0h46m13.72s
0h46m13.43s
0h46m10.23s
0h46m11.13s
0h46m14.51s
0h46m11.47s
0h46m11.66s
0h46m16.21s
0h46m14.30s
0h46m13.32s
0h46m16.08s
0h46m15.92s
0h46m14.21s
0h46m13.71s
0h46m10.38s
Dec.
(J2000)
40o41′18.2′′
40o40′52.4′′
40o40′44.1′′
40o41′ 2.9′′
40o43′44.1′′
40o43′33.5′′
40o42′ 2.9′′
40o41′29.0′′
40o40′39.9′′
40o42′19.8′′
40o43′ 0.9′′
40o42′20.6′′
40o43′38.9′′
40o42′44.8′′
40o42′31.0′′
40o41′35.2′′
40o43′10.6′′
40o42′36.9′′
40o42′32.6′′
40o42′52.9′′
40o44′ 5.8′′
40o41′24.6′′
40o44′ 9.8′′
40o43′28.8′′
40o44′ 2.4′′
40o41′35.5′′
40o41′ 3.5′′
40o43′21.9′′
40o42′51.1′′
40o41′39.8′′
40o43′58.7′′
40o43′10.5′′
40o41′21.5′′
40o42′54.5′′
40o41′24.6′′
40o41′14.2′′
40o41′42.2′′
40o43′34.7′′
40o43′32.5′′
40o42′45.1′′
40o41′30.6′′
40o41′54.5′′
40o44′23.2′′
40o43′48.8′′
40o41′37.0′′
40o44′ 4.1′′
40o44′23.3′′
40o41′47.6′′
40o43′22.7′′
40o44′18.7′′
40o42′25.4′′
40o42′37.9′′
40o43′55.8′′
40o44′24.4′′
40o43′44.0′′
TABLE 1
RR LYRAE PROPERTIES
Period <mF606W > <mF814W > AF606W
(mag)
(days)
0.42
0.382
0.339
0.40
0.37
0.306
0.97
0.687
0.40
0.331
0.287
0.49
0.85
0.532
1.06
0.627
0.46
0.301
0.289
0.15
1.05
0.552
0.56
0.338
1.07
0.464
1.09
0.496
0.366
0.45
0.38
0.366
1.03
0.554
0.43
0.339
0.714
0.51
0.15
0.274
0.63
0.678
0.49
0.774
0.529
1.12
0.42
0.313
0.47
0.735
0.52
0.351
0.506
1.07
0.41
0.267
0.50
0.353
0.36
0.361
0.48
0.275
0.301
0.46
0.95
0.573
0.33
0.726
0.52
0.283
0.589
1.09
0.67
0.631
0.75
0.627
0.96
0.534
0.263
0.48
0.38
0.327
0.88
0.572
0.47
0.634
0.34
0.280
0.559
1.12
0.91
0.554
1.12
0.442
0.40
0.329
0.611
0.68
0.37
0.330
1.01
0.533
0.37
0.313
0.580
0.85
0.65
0.583
0.621
0.67
(mag)
25.38
25.55
25.61
25.17
25.64
25.65
25.59
25.32
25.48
25.40
25.35
25.39
25.67
25.57
25.47
25.42
25.42
25.46
25.37
25.42
25.52
25.38
25.51
25.52
25.48
25.36
25.56
25.53
25.49
25.42
25.59
25.52
25.41
25.30
25.39
25.36
25.27
25.35
25.48
25.66
25.50
25.37
25.56
25.60
25.38
25.41
25.52
25.47
25.26
25.56
25.57
25.35
25.45
25.46
25.43
9
class
c
c
c
ab
c
c
ab
ab
c
c
ab
c
ab
ab
c
c
ab
c
ab
c
ab
ab
ab
c
ab
c
ab
c
d
c
c
c
ab
ab
c
ab
ab
ab
ab
c
c
ab
ab
c
ab
ab
ab
c
ab
c
ab
c
ab
ab
ab
AF814W
(mag)
0.30
0.29
0.29
0.66
0.28
0.32
0.70
0.73
0.33
0.10
0.71
0.36
0.79
0.80
0.31
0.26
0.68
0.30
0.39
0.11
0.43
0.42
0.77
0.28
0.37
0.35
0.80
0.28
0.26
0.26
0.35
0.31
0.62
0.24
0.35
0.52
0.50
0.52
0.68
0.33
0.22
0.64
0.32
0.22
0.84
0.65
0.68
0.29
0.53
0.27
0.78
0.25
0.64
0.59
0.47
(mag)
25.99
26.17
26.27
25.72
26.25
26.32
26.15
25.91
26.17
26.14
25.91
26.04
26.28
26.15
26.13
26.05
25.99
26.12
25.87
26.14
26.04
25.92
26.07
26.21
25.99
26.03
26.15
26.26
26.11
26.05
26.29
26.21
26.00
25.80
26.11
25.96
25.82
25.89
26.05
26.37
26.15
25.95
26.06
26.27
25.96
25.97
26.10
26.16
25.81
26.18
26.15
26.00
26.01
25.99
25.95
10
MJD
(days)
TABLE 2
RR LYRAE PHOTOMETRYa
Photometry (mag)b
Band
V1
V8
25.87
F606W 25.58
F606W 25.27
25.57
25.48
F606W 99.99
25.51
F606W 25.22
26.39
F814W 26.09
F814W 25.98
99.99
26.47
F814W 25.96
26.37
F814W 25.75
25.42
F606W 25.29
F606W 25.37
25.47
V5
25.75
25.57
25.42
25.44
26.14
26.07
26.09
26.19
25.74
25.83
52610.06377
52610.12537
52610.19249
52610.25919
52611.20557
52611.25942
52611.27610
52611.32653
52611.39623
52611.41250
aTable 2 is published in its entirety in the electronic edition of The
Astronomical Journal. A portion is shown here for guidance re-
garding its form and content.
bA magnitude of 99.99 signifies missing data.
with respect to the Galactic halo, therefore resulting in different
age vs. chemical composition patterns.
4.2. Specific Frequency
The specific frequency of RR Lyraes in the M31 halo has
been the subject of debate, due perhaps in part to the diffi-
culty of observing these stars from the ground. Pritchet & van
den Bergh (1987) observed a 2.2′ × 3.3′ field 40′ from the nu-
cleus on the southeast minor axis, and found 30 RR Lyraes
with amplitudes ∼>0.7 mag. Assuming a completeness of 25%,
they determined there should be 120 RR Lyraes with ampli-
tudes ∼>0.7 mag. They took this to be a very strict lower limit,
given the lack of low-amplitude RR Lyraes, and claimed their
value was 55% of the frequency for M3 (NGC5272), a Galac-
tic globular cluster with a high frequency of RR Lyraes. The
specific frequency of RR Lyraes in M3, as given by Harris
(1996) and normalized to a total cluster luminosity (MV t) of
−7.5 mag, is SRR = NRR10(MVt +7.5)/2.5 = 49. Thus, Pritchet
& van den Bergh (1987) determined SRR ≈ 27, which was a
surprisingly high number, given the mean metallicity [Fe/H]=
−0.6 of Mould & Kristian (1986).
Recently, Dolphin et al. (2003) observed a 9.6′ × 9.6′ field
that included the Pritchet & van den Bergh (1987) field, and
found a much lower frequency of RR Lyraes. After finding 24
RR Lyraes and estimating a completeness of 24%, they esti-
mated that there are 100 RR Lyraes in their field. Given the
difference in field size, the Dolphin et al. (2003) frequency is
15 times smaller than that of Pritchet & van den Bergh (1987),
i.e., SRR ≈ 1.8. Dolphin et al. (2003) claimed that their lower
frequency was supported by the CMD of Brown et al. (2003)
-- the same data we are using in this paper -- because they es-
timated that there were only 10 RR Lyraes in the Brown et al.
(2003) data. However, the CMD shown by Brown et al. (2003)
was displayed as a greyscale Hess diagram, in order to clearly
show the characteristics of the main sequence turnoff in a cat-
alog of ∼ 300,000 stars; individual stars were not shown, nor
was variability indicated in any way. In reality, 41 of the 55 RR
Lyraes reported in our current work also appeared in the CMD
of Brown et al. (2003); the remaining 14 were in the ∼ 20%
of the image area masked for that earlier analysis of the star
formation history (see §2).
These earlier estimates of the RR Lyrae frequency were
limited by several factors. The first was uncertainty in the
completeness. RR Lyraes were at the edge of detection in
both the Dolphin et al. (2003) and Pritchet & van den Bergh
(1987) studies, and neither group quantified the completeness
at that depth through artificial star tests. Furthermore, the total
luminosity in these fields could be estimated only from photo-
graphic plates or counts of the bright RGB stars. In contrast,
our artificial star tests indicate that the completeness in our data
POSITIONS OF OTHER VARIABLES
Name
V6
V7
V12
V14
V18
V19
V20
V21
V25
V29
V31
V38
V39
V41
V48
V51
V52
V60
V61
V63
R.A.
(J2000)
0h45m57.91s
0h45m58.14s
0h45m58.06s
0h45m58.78s
0h45m58.87s
0h45m58.90s
0h46m 1.02s
0h45m58.90s
0h45m58.97s
0h46m 1.76s
0h46m 2.46s
0h46m 1.79s
0h46m 0.08s
0h46m 1.07s
0h46m 2.27s
0h46m 1.59s
0h46m 2.51s
0h46m 4.91s
0h46m 5.72s
0h46m 5.43s
Dec.
(J2000)
Name
40o42′48.1′′
V64
40o42′38.6′′
V65
40o42′55.2′′
V67
40o42′29.7′′
V70
40o42′32.3′′
V72
40o42′30.9′′
V74
40o41′ 2.1′′
V84
40o42′32.7′′
V85
40o42′32.0′′
V86
40o40′53.1′′
V91
40o40′41.3′′
V93
40o41′44.0′′
V94
40o42′57.0′′
V99
40o42′44.9′′ V103
40o42′16.3′′ V105
40o43′18.1′′ V106
40o42′39.2′′ V111
40o41′55.9′′ V113
40o41′24.2′′ V116
40o41′55.9′′ V117
TABLE 3
R.A.
(J2000)
0h46m 3.27s
0h46m 4.06s
0h46m 4.42s
0h46m 3.88s
0h46m 6.25s
0h46m 4.44s
0h46m 5.30s
0h46m 9.12s
0h46m 9.55s
0h46m 7.74s
0h46m 9.40s
0h46m 9.77s
0h46m 6.54s
0h46m 8.70s
0h46m 9.32s
0h46m10.84s
0h46m11.28s
0h46m 8.95s
0h46m12.44s
0h46m11.67s
Dec.
(J2000)
Name
40o43′29.4′′ V118
40o42′59.2′′ V119
40o42′46.9′′ V127
40o43′28.5′′ V128
40o41′53.8′′ V129
40o43′25.7′′ V134
40o44′ 5.8′′ V139
40o41′24.7′′ V143
40o41′ 8.5′′ V145
40o42′50.6′′ V150
40o41′45.2′′ V151
40o41′29.9′′ V158
40o44′ 8.0′′ V159
40o43′10.2′′ V168
40o42′53.2′′ V169
40o41′51.5′′ V170
40o41′42.7′′ V171
40o43′29.5′′ V172
40o41′20.8′′ V173
40o41′54.6′′ V174
R.A.
(J2000)
0h46m10.05s
0h46m11.23s
0h46m11.02s
0h46m12.17s
0h46m 9.88s
0h46m13.50s
0h46m11.90s
0h46m14.94s
0h46m14.38s
0h46m12.63s
0h46m13.16s
0h46m13.67s
0h46m14.91s
0h46m 0.63s
0h46m 2.52s
0h46m 8.72s
0h46m 7.35s
0h46m 9.93s
0h46m10.45s
0h46m 8.98s
Dec.
(J2000)
40o43′ 6.0′′
40o42′19.0′′
40o43′ 5.9′′
40o42′18.4′′
40o43′57.4′′
40o41′53.2′′
40o43′17.9′′
40o41′42.8′′
40o42′13.8′′
40o43′48.2′′
40o43′36.0′′
40o43′37.6′′
40o42′51.6′′
40o41′13.8′′
40o40′37.7′′
40o41′53.0′′
40o44′17.1′′
40o43′18.3′′
40o43′ 1.6′′
40o43′ 9.2′′
11
0
1
2
3
4
22
24
26
)
G
A
M
T
S
(
W
4
1
8
F
m
28
-1
25.5
26.0
26.5
27.0
27.5
28.0
)
G
A
M
T
S
(
W
4
1
8
F
m
-1.0
-0.8
-0.6
mF606W-mF814W (STMAG)
-0.4
-0.2
FIG. 5 -- Top panel: Subset of the M31 halo CMD of Brown et al. (2003), highlighting the bright variables. From Figure 2, V84
is marked by a closed circle, while V118 and V60 are marked by closed diamonds. The variables of Figure 4 are shown by open
diamonds. Bottom panel: Expanded view of the faint variables: RRab stars (crosses), RRc stars (open triangles), an RRd star
(closed triangle), dwarf Cepheids (open circles), and eclipsing binaries (closed squares).
is ≈100% at the horizontal branch, our time sampling enables
a search for periodicity with high statistical significance, and
our deep catalog allows an accurate determination of the to-
tal luminosity in our field. Brown et al. (2003) determined
µV ≈ 26.3 mag arcsec−2 in this field. Assuming an extinc-
tion of E(B −V ) = 0.08 ± 0.03 mag (Schlegel et al. 1998) and
a distance modulus (m − M)0 = 24.44 ± 0.1 mag (Freedman
& Madore 1990), we estimate a total brightness in our field
of MV ≈ −10.0 mag, and SRR ≈ 5.6 (∼ 1 RR Lyrae star per
1.5 × 104L⊙). Our frequency is ∼3 times higher than that
found by Dolphin et al. (2003), with the number of RR Lyraes
compared either to the number of RGB stars (the metric of
Dolphin et al. 2003) or the total luminosity of the field (the
metric of Harris 1996). The specific frequency we find is much
higher than one would expect for a population with metallic-
ity [Fe/H]= −0.8 (Figure 10), but not surprising given that the
M31 halo has a wide metallicity distribution, with ∼40% of
the stars at −2.5 <[Fe/H]< −1 (Durrell et al. 2001). Because
about half of the M31 halo is metal-rich and of intermediate age
(Brown et al. 2003), and thus incapable of producing RR Lyrae
stars, one could reasonably shift the location of the M31 data
point in Figure 10 to a metallicity near [Fe/H]= −1.6 and to a
frequency twice as large as that found at [Fe/H] = −0.8.
4.3. Metallicity
Given a larger number of RR Lyraes than expected for a
mean metallicity of [Fe/H]= −0.8, it is reasonable to assume
that the RR Lyraes come from the old, metal-poor component
of the M31 halo. As done in studies of dSph populations (e.g.,
Pritzl et al. 2002; Siegel & Majewski 2000), we can use one
of several empirical relations to estimate the metallicity of the
RR Lyraes in our field. Looking at the mean properties of the
0.6
0.5
0.4
0.3
0.2
0.1
c
b
a
N
/
c
N
0.0
-0.40
-0.42
-0.44
-0.46
-0.48
-0.50
-0.52
-0.54
c
>
P
<
g
o
l
-0.28
-0.24
-0.20
log <Pab>
-0.16
FIG. 7 -- Top panel: The fractions of RRab+RRc stars that are RRc,
vs. the mean RRab periods, for different populations. Galactic glob-
ular clusters (Clement et al. 2001) of Oosterhoff I (open circles) and
Oosterhoff II (closed circles) types are well separated, while Local
Group dSphs (diamonds; Pritzl et al. 2002, 2004; Dall'Ora et al.
2003) and the M31 halo (asterisk) bridge the gap; the Milky Way
halo field is, on average, Oosterhoff I (cross; Cacciari & Renzini
1976). The fraction of RRc stars in the M31 halo is like that in Oost-
erhoff II clusters, while the fraction of RRc stars in dSphs is like that
in Oosterhoff I clusters. Bottom panel: The mean RRc period vs.
the mean RRab period. Again, the dSphs and the M31 halo bridge
the gap between Oosterhoff types, but the mean RRc period in M31
is like that in Oosterhoff I clusters, while the mean RRc period in
dSphs is like that in Oosterhoff II clusters.
12
)
G
A
M
T
S
(
W
6
0
6
F
A
)
G
A
M
T
S
(
W
6
0
6
F
m
s
r
a
t
S
e
a
r
y
L
R
R
1.2
1.0
0.8
0.6
0.4
0.2
0.0
23
24
25
26
27
28
10
8
6
4
2
Ncd / Nabcd = 0.47
Nc / Nabc = 0.46
6
1
3
.
0
=
>
P
<
c
4
9
5
.
0
=
>
b
a
P
<
0
-1.0
-0.8
-0.6
-0.4
-0.2
Log Period (days)
-2
-1
)
g
0
a
m
(
V
M
1
2
3
0.0
FIG. 6 -- Top panel: The amplitude vs. period diagram for short
period variables (excluding the eclipsing binaries). The RRab stars
(crosses) are well-separated from the RRc stars (open triangles), but
the dwarf Cepheids (open circles) are mixed in with the RR Lyraes.
The RRd star (closed triangle) is mixed in the main RRc clump. The
bright blue variable above the HB is also marked (closed circle).
Middle panel: The luminosity vs. period diagram, with the same
symbols. The RR Lyraes are well-separated from the other variables,
although two "dwarf Cepheids" might actually be RR Lyraes. The
left axis provides the observed F606W magnitudes, while the right
gives the transformation to MV appropriate for the color of the center
of the RR Lyrae gap, assuming the Cepheid distance to M31 (see §4.6
for details). Bottom panel: The period histogram for the RR Lyraes.
The parameters used to distinguish Oosterhoff type are labeled.
7 M31
6
5
4
3
2
1
0
7 M3+M15 (normalized)
6
5
4
3
2
1
0
40
M3 - OoI
(a)
(b)
(c)
)
g
a
m
(
V
A
1.4
1.2
1.0
0.8
0.6
0.4
0.2
0.0
-0.4
13
0.0
OoII
OoI
-0.3
-0.2
-0.1
Log Period (days)
FIG. 9 -- The period-amplitude diagram for the RRab stars of the M31
halo (crosses), compared to the period-amplitude relations for Ooster-
hoff I and Oosterhoff II clusters (lines; Clement 2000). Note that we
converted amplitudes in the mF606W bandpass to amplitudes in John-
son V using Lejeune synthetic spectra, but this correction is small (a
6 -- 10% increase in the F606W amplitude). The M31 halo stars appear
to be an intermediate case between Oosterhoff types, although there is
a tendency toward the Oosterhoff I type at large amplitudes.
4.4. Double-Mode Pulsators
As mentioned in §2, five RR Lyrae stars in our sample
show significant scatter in their light curves, as expected for
double-mode pulsators (RRd stars) and RRab stars exhibiting
the Blazhko effect. One of these stars, V90, shows clear pe-
riodicity at two different periods, with P0 = 0.4735 days and
P1 = 0.3534 days. The ratio P1/P0 is sensitive to the mass of the
RRd star, with RRd stars from the two Oosterhoff types clearly
separated in a P1/P0 vs P0 diagram (see Bono et al. 1996 and
references therein). Using these periods to place our confirmed
RRd star in the diagrams of Bono et al. (1996), we find that the
P0 for this star is similar to that of RRd stars found in Oosterhoff
I clusters, but the P1/P0 ratio is typical for RRd stars found in
Oosterhoff II clusters, with a mass of ≈ 0.75 M⊙. It is curious
that an individual star in the M31 halo cannot be placed into
either Oosterhoff type in such a diagram; again, this suggests
that the M31 halo population is not a simple mix of Oosterhoff
types.
4.5. Distance
With the metallicities derived above, we can estimate the
distance to the M31 halo population using the relation of Car-
retta et al. (2000): MV = (0.18 ± 0.09)([Fe/H]+ 1.5) + (0.57 ±
0.07). Note that this relation is consistent with the average of
the methods discussed by Cacciari & Clementini (2003), who
give an excellent review of distance determination using RR
Lyrae stars. Using the synthetic spectra of Lejeune et al. (1997),
an assumed extinction of E(B − V ) = 0.08 ± 0.03 (Schlegel
et al. 1998), and the extinction curve of Fitzpatrick (1999),
we can calculate the offset between mF606W and V as a func-
tion of mF606W − mF814W . The RRab stars have a color range
−0.61 ≤ (mF606W − mF814W ) ≤ −0.5 mag, with correspond-
ing corrections of −0.17 ≤ (V − mF606W ) ≤ −0.13 mag; the
b
a
R
R
f
o
r
e
b
m
u
N
30
20
10
0
10 M15 - OoII
8
6
4
2
0
-0.40 -0.35 -0.30 -0.25 -0.20 -0.15 -0.10
(d)
Log Period (Days)
FIG. 8 -- Panel (a): The period distribution of RRab stars in the M31
halo. Panel (b): The period distribution of RRab stars when the stars
are drawn equally from M3 and M15 -- two clusters rich in RR Lyraes
with distinct Oosterhoff types. Panel (c): The period distribution in
M3 (Clement et al. 2001). Panel (d): The period distribution in M15
(Clement et al. 2001).
RRab and RRc stars, the period-metallicity relations of Sandage
(1993) give [Fe/H] = ( -- log<Pab> −0.389)/0.092 = −1.77,
and [Fe/H] = ( -- log<Pc> −0.670)/0.119 = −1.43, on the Zinn
& West (1984) scale. Given the ensemble population in the
M31 halo and the scatter for these relations shown by Sandage
(1993), the difference in metallicity for the RRab and RRc stars
is not significant, but we can say that the RR Lyraes lie in the
metal-poor tail of the halo metallicity distribution. This finding
is further supported by the period-amplitude-metallicity rela-
tion of Alcock et al. (2000), which provides metallicities on the
Zinn & West (1984) scale for individual RRab stars: [Fe/H]= --
8.85(logPab + 0.15AV) -- 2.60. Converting our amplitudes in the
mF606W bandpass to those in Johnson V , we find a mean [Fe/H]
of -- 1.79, with a standard deviation of 0.32; the accuracy of this
method is σ[Fe/H] = 0.31 per star (Alcock et al. 2000), so the
dispersion in our individual metallicities is not statistically sig-
nificant.
14
y
c
n
e
u
q
e
r
F
c
i
f
i
c
e
p
S
e
a
r
y
L
R
R
140
120
100
80
60
40
20
0
-2.0
-1.5
-1.0
[Fe/H]
-0.5
0.0
FIG. 10 -- The specific frequency of RR Lyraes, normalized to a pop-
ulation at MVt = −7.5 mag, as a function of metallicity, for Galactic
globular clusters (points; Harris 1996) and the M31 halo (asterisk).
Although the frequency of RR Lyraes in the M31 halo is higher than
expected for the mean halo metallicity, the RR Lyraes have a metal-
licity near [Fe/H]= −1.6. Considered as part of the old metal-poor
population, the frequency in the M31 halo is not surprising.
RRc stars have a color range −0.74 ≤ (mF606W − mF814W ) ≤
−0.61 mag, with corresponding corrections of −0.21 ≤ (V −
mF606W ) ≤ −0.17 mag. After correcting the individual stars
with the offsets appropriate to their colors, the RRab stars
lie at <mF606W >= 25.43 mag and <V >= 25.28 ± 0.01 mag,
while the RRc stars lie at <mF606W >= 25.49 mag and <V >=
25.31 ± 0.01 mag (the statistical errors on <V > are much
smaller than 0.01 mag, but the throughputs of the ACS filters
are only known to ∼1%). Assuming [Fe/H]= −1.77, the Car-
retta et al. (2000) relation yields MV = 0.52 ± 0.07 mag for
the RRab stars; assuming [Fe/H]= −1.43 gives MV = 0.58 ±
0.07 mag for the RRc stars. With an extinction of AV = 0.25 ±
0.09, we have distance moduli of (m − M)0 = 24.51 ± 0.11 mag
for the RRab stars, and (m − M)0 = 24.48 ± 0.11 mag for the
RRc stars. Averaging these results gives an RR Lyrae distance
modulus of (m − M)0 = 24.5 ± 0.1 mag, which is in very good
agreement with the Cepheid distance modulus of (m − M)0 =
24.44 ± 0.1 mag (Freedman & Madore 1990), given the uncer-
tainties.
Dolphin et al. (2003) made similar estimates of the RR Lyrae
luminosity in their field, but found <V0>= 24.81 ± 0.11 mag
-- about 0.24 mag brighter than our own estimate. Given their
relatively low completeness, it is plausible that their RR Lyrae
sample is biased toward brighter stars (although they argue oth-
erwise). Dolphin inspected the Brown et al. (2003) CMD and
found support for their bright RR Lyrae estimate, but we stress
again that our previous presentation of the M31 halo data only
showed a greyscale Hess diagram, making such estimates from
the published figure difficult.
5. OTHER VARIABLES
Although the focus of this paper is the RR Lyrae population
in the Andromeda halo, we briefly note the other classes of vari-
ables suitable for followup studies. The short period variables
below the HB (see Figures 2 and 5) fall in the broad category
of dwarf Cepheids; given the wide age range in the M31 halo
(Brown et al. 2003), these are probably a mix of pulsating and
eclipsing stars on the main sequence (e.g., δ Scuti stars) and in
the blue straggler population (e.g., SX Phoenicis stars). Above
the HB, the bluest star (V84) is outside of the Cepheid instabil-
ity strip, but might be a pulsating post-asymptotic branch star.
The other two stars above the HB (V118 and V60) have periods
that are longer than typically found for Anomalous Cepheids,
but their luminosities and periods put them close to the Popula-
tion I and Population II Cepheids. In any case, if they are both
pulsating variables, they cannot belong to the same class, given
that the star with the longer period is fainter. The semiregulars
and long period variables (see Figure 4) include RV Tauri stars
and Mira stars. In globular clusters, Miras are found only at
[Fe/H] > −1 (Frogel & Whitelock 1998), where they have pe-
riods ∼< 310 days. Given the wide age spread in the M31 halo,
one might expect some of the Miras to have periods longer than
310 days, but characterization of these stars would require a
longer baseline than that currently available.
6. SUMMARY
We have presented a complete survey of the RR Lyrae stars
in an M31 halo field, 51 arcmin from the nucleus. We find 29
RRab stars with a mean period of 0.594 days, 25 RRc stars
with a mean period of 0.316 days, and 1 RRd star with a fun-
damental period of 0.473 days and a first overtone period of
0.353 days. The RR Lyrae population of the M31 halo can-
not be clearly placed into either Oosterhoff type, and is dis-
tinct from the Milky Way cluster and halo field populations. In
a broad sense, the Local Group dSphs share the intermediate
Oosterhoff status found the M31 halo, but the characteristics
of the M31 RR Lyraes (<Pab>, <Pc>, Nc/Nabc) are distinct
from the dSphs, suggesting that the M31 halo is not comprised
of dissolved globular clusters like those in the Milky Way or
dissolved Local Group dSphs. The specific frequency of RR
Lyraes (SRR = 5.6) is very high for a mean halo metallicity of
[Fe/H]= −0.8, but within the normal range when considered
as a component of the old, metal-poor halo population. The
mean metallicity of the RR Lyrae population is indeed much
lower than that of the halo, with a mean [Fe/H]= −1.77 for
the RRab stars and a mean [Fe/H]= −1.43 for the RRc stars.
The distance to M31 determined from the RR Lyrae luminos-
ity is (m − M)o = 24.5 ± 0.1 mag, in good agreement with the
Cepheid distance.
Support for proposal 9453 was provided by NASA through
a grant from STScI, which is operated by AURA, Inc., under
NASA contract NAS 5-26555. We are grateful to P. Stetson for
providing the latest version of DAOPHOT and assistance with
its use, and to R. Gilliland for useful discussions. H. Smith
kindly reviewed a draft of this manuscript and offered helpful
comments. We thank the members of the scheduling and op-
erations teams at STScI (especially P. Royle, D. Taylor, and D.
Soderblom) for their efforts in executing a large program during
a busy HST cycle. Our paper was improved by the suggestions
of an anonymous referee.
REFERENCES
15
Krist, J. 1995, ASP Conference Series 77, Astronomical Data Analysis
Software and Systems IV, ed. R.A. Shaw, H.E. Payne, & J.J.E. Hayes, 349
Lafler, J., & Kinman, T.D. 1965, ApJS, 11, 216
Lejeune, T., Cuisinier, F., & Buser, R. 1997, A&AS, 125, 229
Lomb, N.R. 1976, Ap&SS
Mateo, M. 1998, ARA&A, 36, 435
Mould, J., & Kristian, J. 1986, ApJ, 305, 591
Oosterhoff, P. Th. 1939, Observatory, 62, 104
Press, W.H., & Rybicki, G.B. 1989, ApJ, 338, 277
Pritchet, C.J., & van den Bergh, S. 1987, ApJ, 316, 517
Pritzl, B.J., Armandroff, T.E., Jacoby, G.H., & Da Costa, G.S. 2002, AJ, 124,
Pritzl, B.J., Armandroff, T.E., Jacoby, G.H., & Da Costa, G.S. 2002, AJ, 127,
Alcock, C., et al. 2000, AJ, 119, 2194
Bono, G., Caputo, F., Castellani, V., & Marconi, M. 1996, ApJ, 471, 33
Bono, G., Caputo, F., & Stellingwerf, R.F. 1994, ApJ, 423, 294
Brown, T.M. 2003, in "The Local Group as an Astrophysical Laboratory," ed.
M. Livio (Cambridge University Press), in press, astro-ph/0308298.
Brown, T.M., Ferguson, H.C., Smith, E., Kimble, R.A., Sweigart, A.V.,
Renzini, A., Rich, R.M., & VandenBerg, D.A. 2003, ApJ, 592, L17.
Cacciari, C., & Clementini, G. 2003, in Lecture Notes in Physics Volume 635,
Stellar Candles for the Extragalactic Distance Scale, ed. D.M. Alloin & W.
Gieren (New York: Springer), 635, 105
Cacciari, C., & Renzini, A. 1976, A&AS, 25, 303
Carretta, E., Gratton, R.G., Clementini, G., Fusi Pecci, F. 2000, 533, 215
Clement, C.M. 2000, in ASP Conf. Ser. 203, The Impact of Large-Scale
Surveys on Pulsating Star Research, ed. L. Szabados & D.W. Kurtz (IAU
Colloq. 176) (San Francisco: ASP), 266
Clement, C.M., et al. 2001, AJ, 122, 2587
Dall'Ora, M., et al. 2003, AJ, 126, 197
Dolphin, A.E., Saha, A., Olszewski, E.W., Thim, F. Skillman, E.D., Gallagher,
J.S., & Hoessel, J. 2003, AJ, in press, astro-ph/0311300
Durrell, P.R., Harris, W.E., & Pritchet, C.J. 2001, AJ, 121, 2557
Ferguson, A.M.N., Irwin, M.J., Ibata, R.A., Lewis, G.F., & Tanvir, N.R. 2002,
AJ, 124, 1452
Fitzpatrick, E.L. 1999, PASP, 111, 63
Ford, H.C., et al. 1998, Proc. SPIE, 3356, 234.
Freedman, W.L., & Madore, B.F. 1990, ApJ, 365, 186
Frogel, J.A., & Whitelock, P.A. 1998, 116, 754
Harris, W.E. 1996, AJ, 112, 1487
Holland, S., Fahlman, G.G., & Richer, H.B. 1997, AJ, 114, 1488
1464
318
82, 947
Renzini, A. 1980, Mem. S. A. It., 51, 749
Renzini, A. 1983, Mem. S. A. It., 54, 335
Ryan, S.G., & Norris, J.E. 1991, AJ, 101, 1865
Sandage, A. 1993, AJ, 106, 687
Sargent, W.L.W., Kowal, C.T., Hartwick, F.D.A., van den Bergh, S. 1977, AJ,
Scargle, J.D. 1982, ApJ, 263, 835
Schlegel, D.J., Finkbeiner, D.P., & Davis, M. 1998, ApJ, 500, 525
Siegel, M.H., & Majewski, S.R. 2000, AJ, 120, 284
Smith, H.A., et al. 2003, PASP, 115, 43
Stetson, P. 1987, PASP, 99, 191
Suntzeff, N.B., Kinman, T.D., & Kraft, R.B. 1991, ApJ, 367, 528
van den Bergh, S. 1999, A&A Rev., 9, 273
VandenBerg, D.A., 2000, ApJS, 129, 315
Zinn, R., & West, M.J. 1984, ApJS, 55, 45
|
astro-ph/0011094 | 1 | 0011 | 2000-11-03T20:36:28 | Stellar population synthesis models at low and high redshift | [
"astro-ph"
] | The basic assumptions behind Population Synthesis and Spectral Evolution models are reviewed. The numerical problems encountered by the standard population synthesis technique when applied to models with truncated star formation rates are described. The Isochrone Synthesis algorithm is introduced as a means to circumvent these problems. A summary of results from the application of this algorithm to model galaxy spectra by Bruzual and Charlot (1993, 2000) follows. I present a comparison of these population synthesis model predictions with observed spectra and color magnitude diagrams for stellar systems of various ages and metallicities. It is argued that models built using different ingredients differ in the resulting values of some basic quantities (e.g. $M/L_V$), without need to invoking violations of physical principles. The range of allowed colors in the observer frame is explored for several galaxy redshifts. | astro-ph | astro-ph |
Stellar population synthesis models at low
and high redshift†
By Gustavo Bruzual A.
Centro de Investigaciones de Astronom´ıa (CIDA), A.P. 264, M´erida, Venezuela
The basic assumptions behind Population Synthesis and Spectral Evolution models are reviewed.
The numerical problems encountered by the standard population synthesis technique when
applied to models with truncated star formation rates are described. The Isochrone Synthesis
algorithm is introduced as a means to circumvent these problems. A summary of results from
the application of this algorithm to model galaxy spectra by Bruzual and Charlot (1993, 2000)
follows. I present a comparison of these population synthesis model predictions with observed
spectra and color magnitude diagrams for stellar systems of various ages and metallicities. It is
argued that models built using different ingredients differ in the resulting values of some basic
quantities (e.g. M/LV ), without need to invoking violations of physical principles. The range
of allowed colors in the observer frame is explored for several galaxy redshifts. †
1. Introduction
The number distribution of the stellar populations present in a galaxy is a function of
time. Thus, the number of stars of a given spectral type, luminosity class, and metallicity
content changes as the galaxy ages.
In early-type galaxies (E/S0) most of the stars
were formed during, or very early after the initial collapse of the galaxy and the stellar
population ages as times goes by. The chemical abundance in these systems must have
reached the value measured in the stars very quickly during the formation process since
most E/S0 galaxies show little evidence of recent major events of star formation. In late-
type galaxies the stellar population also ages, but there is a significant number of new
stars being formed. Depending on the star formation rate, Ψ(t), the mean age of the stars
in a galaxy may even decrease as the galaxy gets older. In general, in late-type systems
the metal content of the stars and the interstellar medium is an increasing function of
time. Ψ(t) can also increase above its typical value due to interactions between two or
more galaxies or with the environment.
As a consequence of the aging of the stellar population, or its renewal in the case of
galaxies with high recent Ψ(t), the rest-frame spectral distribution of the light emitted by
a galaxy is a function of its proper time. Observational properties such as photometric
magnitude and colors, line strength indices, metal content of gas and young stars, depend
on the epoch at which we observe a galaxy on its reference frame. This intrinsic evolution
should not be mistaken with the apparent evolution produced by the cosmological redshift
z. For distant galaxies both effects may be equally significant.
In order to search for spectral evolution in galaxy samples we must (a) quantify the
amount of evolution expected between cosmological epoch t1 and t2, and (b) design
observational tests that will reveal this amount of evolution, if present. Evolutionary
population synthesis models predict the amount of evolution expected under different
scenarios and allow us to judge the feasibility of measuring it. In the ideal case (Arag´on-
Salamanca et al. 1993; Stanford et al. 1995, 1998; Bender et al. 1996) spectral evolution
is measured simply by comparing spectra of galaxies obtained in such a way that the same
† To appear in Proceedings of the XI Canary Islands Winter School of Astrophysics on Galax-
ies at High Redshift, eds. I. P´erez-Fournon, M. Balcells and F. S´anchez
1
2
Gustavo Bruzual A.: Population Synthesis at low and high z
rest frame wavelength region is sampled in all galaxies, irrespective of z. In this case there
is no need to apply uncertain K-corrections to transform all the spectra to a common
wavelength scale. Alternatively, if this approach cannot be applied, e.g. when studying
large samples of faint galaxies, one can use models which allow for different degrees of
evolution (including none) to derive indirectly the amount of evolution consistent with
the data (Pozzetti et al. 1996; Metcalfe et al. 1996). Clearly, the first approach is to be
preferred whenever possible. In all cases we must rule out possible deviations from the
natural or passive evolution of the stellar population in some of the galaxies under study
induced, for example, by interactions with other galaxies or cluster environment, etc.
The large amount of astrophysical data that has become available in the last few years
has made possible to build several complete sets of stellar population synthesis models.
The predictions of these models have been used to study many types of stellar systems,
from local normal galaxies to the most distant galaxies discovered so far (approaching
z of 4 to 5), from globular clusters in our galaxy to proto-globular clusters forming in
different environments in distant, interacting galaxies. In this paper I present an overview
of results from population synthesis models directly applicable to the interpretation of
galaxy spectra.
2. The Population Synthesis Problem
Stellar evolution theory provides us with the functions Tef f (m, Z, t) and L(m, Z, t)
which describe the behavior in time t of the effective temperature Tef f and luminosity
L of a star of mass m and metal abundance Z. For fixed m and Z, L(t) and Tef f (t)
describe parametrically in the H-R diagram the evolutionary track for stars of this mass
and metallicity. The initial mass function (IMF), φ(m), indicates the number of stars of
mass m born per unit mass when a stellar population is formed. The star formation rate
(SFR), Ψ(t), gives the amount of mass transformed into stars per unit time according
to φ(m). The metal enrichment function (MEF), Z(t), also follows from the theory of
stellar evolution. The population synthesis problem can then be stated as follows. Given
a complete set of evolutionary tracks and the functions φ(m) and Ψ(t), compute the
number of stars present at each evolutionary stage in the H-R diagram as a function
of time. To solve this problem exactly we need additional knowledge about the MEF,
Z(t), which gives the time evolution of the chemical abundance of the gas from which
the successive generations of stars are formed. For simplicity, it is commonly assumed
that φ(m) and Ψ(t) are decoupled from Z(t), even though it is recognized that in real
stellar systems these three quantities are most likely closely interrelated. The spectral
evolution problem can be solved trivially once the population synthesis problem is solved,
provided that we know the spectral energy distribution (SED) at each point in the H-R
diagram representing an evolutionary stage in our set of tracks. The discussion that
follows is based in the work of Charlot and Bruzual (1991, hereafter CB91) and Bruzual
and Charlot (1993, 2000, hereafter BC93 and BC2000). See also Bruzual (1998, 1999,
2000). Unless otherwise indicated, I will ignore chemical evolution and assume that at
all epochs stars form with a single metallicity, Z(t) = constant, and the same φ(m).
Let N o
i be the number of stars of mass Mi born when an instantaneous burst of star
formation occurs at t = 0. This kind of burst population has been called simple stellar
population (SSP, Renzini 1981). When we look at this population at later times, we will
see the stars traveling along the corresponding evolutionary track. If the stars live in
their kth evolutionary stage from time ti,k−1 to time ti,k, then at time t the number of
Gustavo Bruzual A.: Population Synthesis at low and high z
stars of this mass populating the kth stage is simply
Ni,k(t) = (cid:26) N o
0,
i ,
if ti,k−1 ≤ t < ti,k;
otherwise.
3
(1)
For an arbitrary SFR, Ψ(t), we compute the number of stars ηi,k(t) of mass Mi at the
kth evolutionary stage from the following convolution integral
ηi,k(t) = Z t
0
Ψ(t − t′)Ni,k(t′)dt′,
which in view of (1) can be written as
ηi,k(t) = N o
i Z min(t,ti,k)
ti,k−1
Ψ(t − t′)dt′.
(2)
(3)
From (3) we see that the commonly heard statement that the number of stars expected
in a stellar population at a given position in the H-R diagram is proportional to the time
spent by the stars at this position, i.e.
ηi,k ∝ N o
i (ti,k − ti,k−1),
(4)
is accurate only for a constant Ψ(t). For non-constant SFRs, the integral in (3) assigns
more weight to the epochs of higher star formation. For instance, for an exponentially
decaying SFR with e-folding time τ , Ψ(t) = exp(−t/τ ), we have
ηi,k(t) ∝ N o
i {exp[−(t − ti,k)/τ ] − exp[−(t − ti,k−1)/τ ]}.
(5)
Ψ(t) was stronger at time (t − ti,k) than at time (t − ti,k−1), which is clearly taken into
account in (5).
A prerequisite for building trustworthy population synthesis and spectral evolution
models is an adequate algorithm to follow the evolution of consecutive generations of
stars in the H-R diagram. This goal is accomplished by the standard technique described
above provided that the function Ψ(t) extends from t = 0 to t = ∞. Special caution is
required if Ψ(t) becomes 0 at a finite age. As an illustration, let us consider the case of
a burst of star formation which lasts for a finite length of time τ ,
or equivalently,
Ψ(t) = (cid:26) Ψo,
0,
if 0 ≤ t ≤ τ ;
otherwise,
Ψ(t − t′) = (cid:26) Ψo,
0,
if t − τ ≤ t′ ≤ t;
otherwise.
In this case, equation (2) reduces to
ηi,k(t) = (cid:26) > 0,
0,
if [ti,k−1, ti,k] ∩ [t − τ, t] 6= ∅;
otherwise.
(6)
(7)
(8)
From (8) we see that ηi,k(t) can be = 0 depending on the value of τ and on the value of t
chosen to sample the stellar population. The most extreme case is that of the SSP (τ = 0),
for which Ψ(t) = δ(t), and ηi,k(t) in (8) is identical to Ni,k(t) in (1). For a typical set of
evolutionary tracks, there is no grid of values tj of the time variable t for which all the
stellar evolutionary stages included in the tracks can be adequately sampled for arbitrarily
chosen values of τ . This is obviously an undesirable property of population synthesis
models. The models should be capable of representing galaxy properties in a continuous
and well behaved form, independent of the sampling time scale t. Consequently, standard
population synthesis models for truncated Ψ(t) as given by (6) will reflect the coarseness
of the set of evolutionary tracks, and may miss stellar evolutionary phases depending on
4
Gustavo Bruzual A.: Population Synthesis at low and high z
our choice of t. This results in unwanted and unrealistic numerical noise in the predicted
properties of the stellar systems studied with the synthesis code (see example in CB91).
A solution to this problem is to build a set of evolutionary tracks with a resolution in
mass which is high enough to guarantee that all evolutionary stages, i.e. all values of k in
(2), can be populated for any choice of t or τ in (6). In other words, there must be tracks
for so many stellar masses in this ideal library that for any model age, we have at hand
the position in the H-R diagram of the star which at that age is in the kth evolutionary
stage. The realization of this library is not possible with present day computers. This
limitation can be circumvented by careful interpolation in a relatively complete set of
evolutionary tracks.
3. The Isochrone Synthesis Algorithm
The isochrone synthesis algorithm described in this section allows us to compute con-
tinuous isochrones of any age from a carefully selected set of evolutionary tracks. The
isochrones are then used to build population synthesis models for arbitrary SFRs Ψ(t),
without encountering any of the problems mentioned above. This algorithm has been
used by CB91, BC93, and BC2000. The details of the algorithm follow.
A relationship is built between the main sequence (MS) mass of a star M and the age
t of the star during the kth evolutionary stage (CB91). Ignoring mass loss is justified
because M is used only to label the tracks. For the set of tracks used by BC93 (described
in the next section) 311 different relationships log M vs. log t are built, which can be
visualized as 311 different curves in the (log t,
log M ) plane. We derive by linear
interpolation the MS mass mk(t′) of the star which will be at the kth evolutionary stage
at age t′, given by
log mk(t′) = Ak,i log(Mi) + (1 − Ak,i) log(Mi+1),
where
Ak,i =
log ti+1,k − log t′
log ti+1,k − log ti,k
.
(9)
(10)
ti,k represents the age of the star of mass Mi at the kth evolutionary stage, and
and
ti,k ≤ t′ < ti+1,k,
Mi+1 ≤ mk(t′) < Mi.
The procedure is performed for all the curves that intersect the log t = log t′ line. We
thus obtain a series of values of mk which must now be assigned values of log L and
log Tef f in order to define the isochrone corresponding to age t′.
To compute the integrated properties of the stellar population, we must specify the
number of stars of mass mk. This number is determined from the IMF, which we can
write generically as
n(mk) = φ(m−
k , m+
k , 1 + x).
(11)
Of this number of stars,
(12)
stars are assigned the observational properties of the star of mass Mi at the kth evolu-
tionary stage, and
Ni,k = Ak,in(mk)
(13)
stars, the observational properties of the star of mass Mi+1 at the same kth stage. This
Ni+1,k = (1 − Ak,i)n(mk)
Gustavo Bruzual A.: Population Synthesis at low and high z
5
procedure is equivalent to interpolating the tracks for the stars of mass Mi and Mi+1 to
derive the values of log L and log Tef f to be assigned to the star of mass mk, but this
intermediate step is unnecessary. In (11) m−
k are computed from
k and m+
m−
k = (mk−1mk)1/2,
m+
k = (mkmk+1)1/2.
(14)
In this case mk−1, mk, and mk+1 represent the masses obtained by interpolation at the
given age in the segments representing the (k−1)th, kth, and (k+1)th stages, respectively.
The IMF has been assumed to be a power law of the form (Salpeter 1955)
φ(m1, m2, 1 + x) = cZ m2
m1
m−(1+x)dm.
(15)
For a given Ψ(t) equation (2) gives the number ηi,k(t) of stars of mass Mi at the kth
evolutionary stage. If fi,k(λ) represents the SED corresponding to the star of mass Mi
during the kth stage, then the contribution of the ηi,k(t) stars to the integrated evolving
SED of the stellar population is simply given by
The resulting SED for the stellar population is given by
Fi,k(λ, t) = ηi,k(t)fi,k(λ).
F (λ, t) = Xi,k
Fi,k(λ, t).
(16)
(17)
4. Evolutionary Population Synthesis Models
A number of groups has developed in recent years different population synthesis mod-
els which provide a sound framework to investigate the problem of spectral evolution
of galaxies. Some of the most commonly used models are Arimoto & Yoshii (1987),
Guiderdoni & Rocca-Volmerange (1987), Buzzoni (1989), Bressan, Chiosi & Fagotto
(1994), Fritze-v.Alvensleben & Gerhard (1993), Worthey (1994), Bruzual & Charlot
(1993, 2000). The basic astrophysical ingredients used in these models are: (1) Stellar
evolutionary tracks of one or more metallicities; (2) Spectral libraries, either empirical
or theoretical model atmospheres; (3) Sets of rules, or calibration tables, to transform
the theoretical HR diagram to observational quantities (e.g. B − V vs. Tef f , V −
K vs. Tef f , B.C. vs. Tef f , etc.). These rules are not necessary when theoretical model
atmosphere libraries are used which are already parameterized according to Tef f , log g,
and [Fe/H]; (4) Additional information, such as analytical fitting functions, required to
compute various line strength indices (Worthey et al. 1994). Regardless of the specific
computational algorithm used, all evolutionary synthesis models depend on three ad-
justable parametric functions: (1) the stellar initial mass function, f (m), or IMF; (2)
the star formation rate, Ψ(t); and (3) the chemical enrichment law, Z(t). For a given
choice of f (m), Ψ(t), and Z(t), a particular set of evolutionary synthesis models provides:
(1) Galaxy spectral energy distribution vs. time, Fλ(λ, Z(t), t); (2) Galaxy colors and
magnitude vs. time; (3) Line strength and other spectral indices vs. time. Some authors
(e.g. Bressan et al. 1994; Fritze-v.Alvensleben & Gerhard 1994) consider that Z(t) can
be derived self-consistently from their models. In other instances Z(t) is introduced as
an external piece of information. I discuss below some results from work still in progress
in collaboration with S. Charlot.
6
Gustavo Bruzual A.: Population Synthesis at low and high z
Figure 1. Evolving spectral energy distributions. (a) Evolution in time of the SED of a SSP
computed for the Salpeter IMF (mL = 0.1, mU = 125 M⊙). The age in Gyr is indicated next
to each spectrum. (b) Same as (a) but for a composite population in which stars form according
to Ψ(t) = exp(−t/τ ) for τ = 3 Gyr. The total mass of each model galaxy is 1 M⊙. Fλ in frame
(b) has been multiplied by 100 to use a common vertical scale.
5. Stellar Ingredients
BC2000 have extended the BC93 evolutionary population synthesis models to provide
the evolution in time of the spectrophotometric properties of SSPs for a wide range of
stellar metallicity. In an SSP all the stars form at t = 0 and evolve passively afterward.
The BC2000 models are based on the stellar evolutionary tracks computed by Alongi et
al. (1993), Bressan et al. (1993), Fagotto et al. (1994a, b, c), and Girardi et al. (1996),
which use the radiative opacities of Iglesias et al. (1992). This library includes tracks for
stars with initial chemical composition Z = 0.0001, 0.0004, 0.004, 0.008, 0.02, 0.05, and
0.10 (Table 1), with Y = 2.5Z + 0.23, and initial mass 0.6 ≤ m/M⊙ ≤ 120 for all
metallicities, except Z = 0.0001 (0.6 ≤ m/M⊙ ≤ 100) and Z = 0.1 (0.6 ≤ m/M⊙ ≤ 9).
This set of tracks will be referred to as the Padova or P tracks hereafter. A similar set
of tracks for slightly different values of Z has been published by Girardi et al. (2000).
A comparison of the predictions of models built with both sets of Padova tracks will be
shown elsewhere (but see Fig. 7 and Bruzual 2000).
The published tracks go through all phases of stellar evolution from the zero-age main
sequence to the beginning of the thermally pulsing regime of the asymptotic giant branch
(AGB, for low- and intermediate-mass stars) and core-carbon ignition (for massive stars),
and include mild overshooting in the convective core of stars more massive than 1 M⊙.
Gustavo Bruzual A.: Population Synthesis at low and high z
7
Z
X
Y
[Fe/H]
0.0001
0.0004
0.0040
0.0080
0.0200
0.0500
0.1000
0.7696
0.7686
0.7560
0.7420
0.7000
0.5980
0.4250
0.2303
0.2310
0.2400
0.2500
0.2800
0.3520
0.4750
-2.2490
-1.6464
-0.6392
-0.3300
0.0932
0.5595
1.0089
Table 1. Model chemical composition
The Post-AGB evolutionary phases for low- and intermediate-mass stars were added to
the tracks by BC2000 from different sources (see BC2000 for details).
BC2000 use as well a parallel set of tracks for solar metallicity computed by the Geneva
group (Geneva or G tracks hereafter), which provides a framework for comparing models
computed with two different sets of tracks.
The BC2000 models use the library of synthetic stellar spectra compiled by Lejeune
et al. (1997, 1998, LCB97 and LCB98 hereafter) for all the metallicities in Table 1. This
library consists of Kurucz (1995) spectra for the hotter stars (O-K), Bessell et al. (1989,
1991) and Fluks et al. (1994) spectra for M giants, and Allard & Hauschildt (1995)
spectra for M dwarfs. For Z = Z⊙, BC2000 also use the Pickles (1998) stellar atlas,
assembled from empirical stellar data.
6. Spectral evolution at fixed metallicity
Fig. 1a shows the evolution in time of the SED for the SSP model. In an SSP all
the stars form at t = 0 and evolve passively afterward. In all the examples shown in
this paper I assume that stars form according to the Salpeter (1955) IMF in the range
from mL = 0.1 to mU = 125 M⊙. The total mass of the model galaxy is 1 M⊙. The
evolution is fast and is dominated by massive stars during the first Gyr in the life of
the SSP (6 top SEDs). The flux seen around 2000 A at 4 and 7 Gyr is produced by
the turn-off stars. The UV-rising branch (Burstein et al. 1988, Greggio and Renzini
1990) seen after 10 Gyr is produced by the PAGB stars. These stars are also responsible
of the decrease in the amplitude of the 912 A discontinuity observed after 4 Gyr. The
SSP model is the basic ingredient which, together with the convolution integral (2), is
used to compute models with arbitrary SFRs and equal IMF. For illustration I show in
Fig. 1b the evolution of a model with Ψ(t) = exp(−t/τ ) for τ = 3 Gyr. The UV to
optical spectrum remains roughly constant during the main episode of star formation
because of the continuous input of young massive stars, but the near-infrared light rises
as evolved stars accumulate. When star formation drops, the spectral characteristics at
various wavelengths are determined by stars in advanced stages of stellar evolution.
7. Dependence of galaxy properties on stellar metallicity
Fig. 2 shows the predicted SEDs at t = 12 Gyr for chemically homogeneous SSPs of
the indicated metallicity. The SEDs shown in Fig. 2 have been normalized at λ = 5500A
to make the comparison more clear. Fig. 3. shows the evolution in time of the B − V
8
Gustavo Bruzual A.: Population Synthesis at low and high z
Figure 2. Chemically homogeneous BC2000 SSP model galaxy SEDs at age = 12 Gyr. Each
line pattern represents a different metallicity, as indicated inside the frame. All the models
shown were computed for the Salpeter IMF (mL = 0.1, mU = 125 M⊙). The total mass of each
model galaxy is 1 M⊙. The SEDs have been normalized at λ = 5500A.
and V − K colors, and the M/LV ratio predicted by BC2000 for the same SSPs shown
in Fig. 2.
From Figs. 2 and 3 it is apparent that there is a uniform tendency for galaxies to
become redder in B − V as the metallicity increases from Z = 0.0001 ( 1
200 Z⊙) to Z =
0.05 (2.5 × Z⊙). The V − K color and the M/LV ratio show the expected tendency with
metallicity, i.e. V − K becomes redder and M/LV becomes higher with increasing Z.
Fig. 4 shows the evolution in time of the Mgb, Hβ, and Ca spectral indices as defined
by Worthey (1994) for the same BC2000 SSP models shown in Figs. 2 and 3. Again, the
models show the expected tendency with Z and match the values computed by Worthey
(1994).
It should be remarked that the time behavior of the line strength indices at
constant Z is due to the change in the number of stars at different positions in the
HR diagram produced by stellar evolution and is not related to chemical evolution. The
indices change also in chemically homogeneous populations. The Hβ index is less sensitive
to the stellar metallicity than the Mgb and Ca index. Instead, the Hβ index is high when
there is a large fraction of MS A-type stars (t < 1) Gyr.
In Fig. 5 I compare the behavior of the SSP models in the (U − B) vs. (B − V ) color
plane with the LMC cluster data from Bica et al. (1996b), discriminated by SWB class
and model age. In each panel the models (lines) are shown in the range of age for which
the predicted colors overlap the observed colors for the class. It is apparent from the
figure that the models reproduce quite well the observed colors.
Gustavo Bruzual A.: Population Synthesis at low and high z
9
Figure 3. Evolution in time of the B − V , and V − K colors, and the M/LV ratio for the
BC2000 SSP models shown in Fig. 2. Each line represents a different metallicity, as indicated
in the bottom panel.
8. Calibration of the models in the C-M diagram
Population synthesis models, frequently used to study composite stellar populations in
distant galaxies, are rarely confronted with observations of local simple stellar popula-
tions, such as globular star clusters, whose age and metal content have been constrained
considerably in later years. If the models cannot reproduce the CMDs and SEDs of these
objects for the correct choice of parameters, their predictive power becomes weaker and
their usage to study complex galaxies may not be justified.
It is important to compare the properties of the population synthesis models to obser-
vations of stellar systems whose age and metallicity is well determined. This is a means
to test to what extent the adopted relationships between stellar color and magnitude
and effective temperature and luminosity (or surface gravity) introduce systematic shifts
between the predicted and observed isochrones in the C-M diagram.
For each value of the metallicity Z listed in Table 1, and a particular choice of the IMF
φ(m), a BC2000 SSP model consists of a set of 221 evolving integrated SEDs spanning
from 0 to 20 Gyr. The isochrone synthesis algorithm used to build these models renders
it straightforward to compute the loci described by the stellar population in the CMD
at any time step and in any photometric band. We can extract the model SED that best
reproduces a given observed SED and assign an age to the program object, and then
examine how well the isochrone computed at this age fits the most significant features in
the CMD of this object, if available.
10
Gustavo Bruzual A.: Population Synthesis at low and high z
Figure 4. Evolution in time of the Mgb, Hβ, and Ca spectral indices as defined by Worthey
(1994) for the BC2000 models shown in Fig. 2. The different symbols represent the values of
the indices computed by Worthey for the same range in Z. Each line represents a different
metallicity, as indicated in the middle panel.
Fig. 6 compares the (B, B − V ) observations of the clusters M67 and the Hyades with
the isochrones obtained from the BC2000 models. These two clusters have nearly solar
metallicity: M67 ([Fe/H]≈ 0.01), the Hyades ([Fe/H]≈ 0.15). For M67 we adopted a
distance modulus of 9.5 mag and a color excess E(B − V ) = 0.06 mag (Janes 1985). For
the Hyades a distance modulus of 3.4 mag (Peterson and Solensky 1988) and E(B −V ) ≈
0. Estimates of the ages of these clusters vary from 4 to 4.3 Gyr for M67 and from 0.5 to
0.8 Gyr for the Hyades. The isochrones are shown at 4 Gyr (M67) and 0.6 Gyr (Hyades).
The data points for M67 are from Eggen and Sandage (1964, open circles), Racine (1971,
filled circles), Janes and Smith (1984, triangles), and Gilliland et al. (1991; crosses). For
the Hyades the observations are from Upgren (1974, triangles), Upgren and Weis (1977,
filled circles), and Micela et al. (1988, open circles).
Fig. 7 shows a comparison of the excellent HST CMD diagram of NGC 6397 assembled
from various sources by D'Antona (1999) with isochrones computed from the Padova-
1994 tracks for Z = 0.0004 and the Padova-2000 tracks for Z = 0.001 at ages 10 to 16
Gyr. The original version of the model atmospheres in the LCB atlas was used to derive
the colors in Fig. 7. The corrected version of these models produces considerably worse
agreement with the observations, mainly in the MS from the turn-off down. The redder
cluster RGB most likely reflects a slightly higher metallicity than Z = 0.0004, close to
Z = 0.001. Despite the discrepancies seen in Figs. 6 and 7 (mainly in the RGB of M67
and NGC 6397) the agreement between the predicted isochrones and the loci in the CMD
Gustavo Bruzual A.: Population Synthesis at low and high z
11
Figure 5. Points: LMC cluster (U − B) vs. (B − V ) colors from Bica et al. (1996b),
discriminated by SWB class. The lines represent SSP models at various age ranges.
of these clusters may be regarded as satisfactory, and is excellent in some parts of the
diagram.
9. Observed Color-Magnitude Diagrams and Integrated Spectra
The high quality and depth of the HST VI-photometry NGC 6528 (Z <∼ Z⊙), as well
as the high signal-to-noise ratio of the integrated SED of NGC 6528 currently available,
provide an excellent framework for testing the range of validity of population synthesis
models (see Bruzual et al. 1997 for details).
The integrated spectrum of NGC 6528 in the wavelength range λ = 3500 - 9800 A was
obtained by combining the visible, near-infrared, and near-ultraviolet spectra of Bica
& Alloin (1986, 1987) and Bica et al.
(1994), respectively. We applied a reddening
correction of E(B-V) = 0.66, adopted in Bica et al. (1994 and references therein). The
SED of NGC 6528 is typical of old stellar populations of high metal content (Santos et
al. 1995). It may happen that NGC 6528, similarly to NGC 6553, has [Fe/H] < 0.0
(Barbuy et al. 1997a), whereas the [α-elements/Fe] are enhanced, resulting in [Z/Z⊙] ≈
0.0, or possibly slightly below solar. Since we have evolutionary tracks for Z = 0.02 and
Z = 0.008, we adopt Z = 0.02.
The data available for this cluster provide a unique opportunity to examine the 6
different options for the Z = Z⊙ models considered by BC2000. Thus, we can study
objectively which of the basic building blocks used by BC2000 for Z = Z⊙: P or G tracks;
Pickles, LCB97-O (original version), or LCB97-C (corrected version) stellar libraries, is
12
Gustavo Bruzual A.: Population Synthesis at low and high z
Figure 6. CMD of M67 and the Hyades compared with isochrones derived from the Padova
tracks for solar metallicity. See text for details.
Model
1
2
3
4
5
6
Spectral
Library
Pickles
"
LCB97-C
"
LCB97-O
"
Stellar Best-fitting
Tracks
age (Gyr) Σ2
min
P
G
P
G
P
G
11.75
10.25
10.00
8.50
11.50
9.00
2.22
1.34
1.73
1.38
3.35
1.95
Table 2. Z⊙ model fits to SED of NGC 6528
most successful at reproducing the data. In Table 2 I show the age at which Σ2, defined
as the sum of squared residuals [log Fλ(observed) − log Fλ(model)]2, is minimum for
various Z = Z⊙ models. The values of Σ2
min given in Table 2 indicate the goodness-of-
fit. According to this criterion, models 2 and 4 provide the best fit to the integrated
SED of NGC 6528 in the wavelength range λ 3500-9800 A. This fit is shown in Fig.
8. Except for the differences in the best-fitting age, model 3 provides a comparable,
although somewhat poorer, fit to the SED of this cluster. The residuals for models 1, 5
Gustavo Bruzual A.: Population Synthesis at low and high z
13
Figure 7. CMD of NGC 6397 compared with isochrones derived from the Padova-1994 tracks
for Z = 0.0004 (top), and the Padova-2000 tracks for Z = 0.001 (bottom). The observations
were assembled by D'Antona (1999) from the sources indicated in the figure.
14
Gustavo Bruzual A.: Population Synthesis at low and high z
Figure 8. Best fit to the integrated spectrum of NGC 6528 (heavy line) in the range λλ 3500
- 9800 A for model 2 in Table 2 (thin line extending over the full wavelength range). The best
fit occurs at 10.25 Gyr. The residuals of the fit, log Fλ(observed) − log Fλ(model), are shown
as a function of wavelength.
Figure 9.
Intrinsic MV vs. (V − I)0 CMD of NGC 6528 shown together with theoretical
isochrones for the Z = Z⊙ models listed in Table 2.
Gustavo Bruzual A.: Population Synthesis at low and high z
15
and 6 are considerably larger. Spectral evolution is slow at these ages, and the minimum
in the function Σ2 vs. age is quite broad. Reducing or increasing the model age by 1
or 2 Gyr produces fits of comparable quality to the one at which Σ2 is minimum. For
instance, for model 2, Σ2(12 Gyr) = 1.88, and Σ2(8 Gyr) = 2, which are still better fits
according to the Σ2 criterion than the best fits provided by some models in Table 2.
Fig. 9 shows the intrinsic HST VI CMD of NGC 6528 together with the isochrones
corresponding to the models in Table 2.
It is apparent from this figure that all the
isochrones shown provide a good representation of the cluster population in this CMD,
especially the position of the turn-off and the base of the asymptotic giant branch (AGB).
We note that NGC 6528 shows a double turnoff, the upper one being due to contamination
from the field star main sequence. The appropriate TO location would be around that
indicated by the 10, 11 and 12 Gyr isochrones. Despite the fact that models 2 and 4
provide better fits to the SED of this cluster than models 1 and 3, the isochrones from
model 3 reproduce more closely the CMD diagram than models 1 and 2. The noisy
nature of the isochrones computed with models 1 and 2 is due to the lack of some stellar
types in the Pickles stellar library.
From these results we conclude that:
(1) The λ = 3500 − 9800 A SED for E(B-V)=0.66 and the VI CMD of NGC 6528 are
well reproduced by Z = Z⊙ models at an age from 9 to 12 Gyr.
(2) The age derived from the fit to the observed SED of NGC 6528 is extremely sensitive
to the assumed E(B-V). Using E(B-V) = 0.59 instead of 0.66, increases the best-fitting
ages by 2 to 3 Gyr, since the observed SED is then intrinsically redder. On the other
hand, if E(B-V) = 0.69, the observed SED is intrinsically bluer and the derived ages are
2 to 3 Gyr younger than for E(B-V) = 0.66. However, inspecting the isochrones in the
CMD, the ages derived for E(B-V) = 0.66 listed in Table 2, seem appropriate.
(3) The SED and CMD of this cluster are consistent with those expected for a Z = Z⊙
population at an age of ≈ 9-12 Gyr, if overshooting occurs in the convective core of stars
down to 1 M⊙ (P tracks, models 1, 3, and 5 in Table 2). If overshooting stops at 1.5 M⊙,
as in the G tracks, this age is reduced to ≈ 8-10 Gyr (models 2, 4, and 6 in Table 2).
(4) For the same spectral library, the ages derived from the P tracks are older than
the ones derived from the G tracks. This is due to the fact that the P tracks include
overshooting in the convective core of stars more massive than 1 M⊙ whereas the G
tracks stop overshooting at 1.5 M⊙. Thus, stars in this mass range require more time in
the P tracks to leave the main sequence than in the G tracks.
(5) For the same set of evolutionary tracks, the corrected LCB97 library seems to pro-
vide a better fit to the CMD than the Pickles atlas. Interpolation in the finer LCB97 grid
of models produces smoother isochrones than in the coarser Pickles atlas. We attribute
this to the fact that M stars are very sparse in the Pickles atlas. Furthermore, the tem-
perature scale becomes problematic for these stars in the Pickles atlas. In the LCB97
library, the temperature scale for giants relies on measurements of angular diameters
and fluxes, which enter directly in the definition of effective temperature. For dwarfs,
the temperature scale is more difficult to define, as discussed in LCB98.
(6) Noticeable differences exist in the isochrones computed for both sets of LCB97
libraries. The differences are more pronounced for the M giants of (V − I)0 > 1.6, and
(J − K)0 > 1, corresponding to a temperature of Te ≤ 4000K, and for the cool dwarfs
of (V − I)0 > 1, corresponding to a temperature of Te ≤ 4700K. There are very few of
these stars in the CMD of NGC 6528 to favor a particular choice of library. However,
the corrected library produces better fits to the observed SED. We attribute this fact to
the relative importance of the luminosity of M giants.
(7) In general the LCB97 corrections redden the stellar SEDs in the optical range,
16
Gustavo Bruzual A.: Population Synthesis at low and high z
producing redder SSP models at an earlier age. As a consequence, the ages derived in
Table 1 for the LCB97-C library are younger than the ones derived from the LCB97-O
library for the same set of tracks.
(8) We have adopted Z = Z⊙ for this cluster. However, for a slightly lower value of
Z, the derived age would be older.
10. Comparison of model and observed spectra
10.1. Solar metallicity
Fig. 10 shows a model fit to the average spectrum of an E galaxy (kindly provided by
M. Rieke). The model SED is the line extending over the complete wavelength range
shown in the figure. The observed SED covers the range from 3300 A to 2.75 µm. The
residuals (observed - model) are shown at the bottom of the figure in the same vertical
scale. The model corresponds to a 10 Gyr Z = Z⊙ SSP computed for the Salpeter
(1955) IMF (mL = 0.1, mU = 125 M⊙) using the P tracks and the Pickles (1998) stellar
atlas. The fit is excellent over most of the spectral range. A minor discrepancy remains
in the region from 1.1 to 1.7 µm. The source of this discrepancy is not understood at
the moment. In Fig. 11 I show the same model and E galaxy SED as in Fig. 10 but in
different units. In addition, in Fig. 11 I include the broad band fluxes representing the
average of many E galaxies in the Coma cluster (solid squares) from A. Stanford (private
communication). The observed SED is the one with the lowest spectral resolution. Fig.
12 shows a closer look at the same data in an enlarged scale. Again the agreement is
excellent for the 3 data sets. The discrepant line in Fig. 12 corresponds to the same
model shown in Figs 10 and 11 but I used the LCB97 synthetic stellar atlas instead of
the empirical stellar SEDs. Fig. 12 shows clearly that the models based on empirical
stellar SEDs are to be preferred over the ones based on theoretical model atmospheres.
Unfortunately, complete libraries of empirical stellar SEDs are available only for solar
metallicity.
10.2. Non-solar metallicity
Figs 13 and 14 show the results of a comparison of SSP models built for various metallic-
ities using the LCB97 atlas, all for the Salpeter IMF, with several of the average spectra
compiled by Bica et al. (1996a). The name and the metal content of the observed spectra
indicated in each panel is as given by Bica et al. The quoted age is derived from the
best fit of our model spectra to the corresponding observations. The residuals (observed
- model) are shown in the same vertical scale. See the description to Fig. 10 above for
more details. Even though, in detail, the fits for non-solar metallicity stellar populations
are not as good as the ones for solar metallicity, over all the models reproduce the ob-
servations quite well over a wide range of [Z/Z⊙], and provide a reliable tool to study
these stellar systems. The discrepancy can be due both to uncertainties in the synthetic
stellar atlas or the evolutionary tracks at these [Z/Z⊙]. I have used SSPs in all the fits,
neglecting possibly composite stellar populations, as well as any interstellar reddening.
11. Different sources of uncertainties in population synthesis models
11.1. Uncertainties in the astrophysics of stellar evolution
There are significant differences in the fractional contribution to the integrated light by
red giant branch (RGB) and asymptotic giant branch (AGB) stars in SSPs computed for
different sets of evolutionary tracks. Fig. 15 shows the contribution of stars in various
Gustavo Bruzual A.: Population Synthesis at low and high z
17
Figure 10. Best fit to an average Elliptical galaxy SED (heavy line) in the range λλ 3300
- 27500 A. The model is the thin line extending over the full wavelength range. The best fit
occurs at 10 Gyr for this model SED. The residuals of the fit, log Fλ(observed) − log Fλ(model),
are shown as a function of wavelength.
Best fit to an average Elliptical galaxy SED (same model and E galaxy SED
Figure 11.
as in Fig. 10 but in different units). The broad band fluxes representing the average of many
E galaxies in the Coma cluster (solid squares) from A. Stanford (private communication) are
shown. The observed SED is the one with the lowest spectral resolution.
18
Gustavo Bruzual A.: Population Synthesis at low and high z
Figure 12. Fits to an average Elliptical galaxy SED. This figure shows a closer look at the
same data of Fig. 11 in an enlarged scale. The discrepant line in corresponds to the same model
shown in Figs 10 and 11 but using the LCB97 synthetic stellar atlas instead of the empirical
stellar SEDs.
evolutionary stages to the bolometric light, and to the broad-band U BV RIKL fluxes for
a Z = Z⊙ model SSP computed for the Salpeter IMF (mL = 0.1, mU = 125 M⊙) using
the P tracks and the Pickles (1998) stellar atlas. The meaning of each line is indicated
in the top central frame. Fig. 16 shows the corresponding plot for an equivalent model
computed according to the G tracks. The contribution of the RGB stars is higher in the
P track model than in the G track model. Correspondingly, the AGB stars contribute
less in the P track model than in the G track model. For instance, for t > 1 Gyr, RGB
and AGB stars contribute 40% and 10%, respectively, to the bolometric light in the P
track model (Fig. 15). These fractions change to 30% and 20% in the G track model
(Fig. 16). These differences are seeing more clearly in Fig. 17 which shows the ratio
of the fractional contribution by different stellar groups in the G track model to that in
the P track model. According to the fuel consumption theorem (Renzini 1981), these
numbers reflect relatively large differences in the amount of fuel used up in the RGB and
AGB phases by stars of the same mass and initial chemical composition depending on
the stellar evolutionary code.
Fig. 20 shows the difference in B magnitude and B − V and V − K color between a G
track model SSP and a P track model SSP as seen both in the rest frame of the galaxy
(vs. galaxy age in the left hand side panels) and the observer frame (vs. redshift in the
right hand side panels). These differences reach quite substantial values. The observer
frame quantities include both the k and the evolutionary corrections. Here and elsewhere
in this paper I assume H0 = 65 km s−1 Mpc−1, Ω = 0.10, and the age of galaxies to be
tg = 12 Gyr.
Gustavo Bruzual A.: Population Synthesis at low and high z
19
Figure 13. Best fit to average optical SED (heavy line) of star clusters of various metallicities
compiled by Bica et al.
(1996a). The SSP model is the thin line extending over the full
wavelength range. The residuals of the fit, log Fλ(observed) − log Fλ(model), are shown as a
function of wavelength. The name and the metal content of the observed spectra indicated in
each panel is as given by Bica et al. The quoted age is derived from the best fit of our model
spectra for the indicated metallicity to the corresponding observations.
Figure 14. Same as Fig. 13, but for four different star clusters.
20
Gustavo Bruzual A.: Population Synthesis at low and high z
Figure 15. Contribution of stars in various evolutionary stages to the bolometric light, and
to the broad-band U BV RIKL fluxes for a Z = Z⊙ model SSP computed for the Salpeter IMF
(mL = 0.1, mU = 125 M⊙) using the P tracks and the Pickles (1998) stellar atlas.
Figure 16. Contribution of stars in various evolutionary stages to the bolometric light, and
to the broad-band U BV RIKL fluxes for a Z = Z⊙ model SSP computed for the Salpeter IMF
(mL = 0.1, mU = 125 M⊙) using the G tracks and the Pickles (1998) stellar atlas.
Gustavo Bruzual A.: Population Synthesis at low and high z
21
11.2. On the energetics of model stellar populations
Buzzoni (1999) has argued that most population synthesis models violate basic prescrip-
tions from the fuel consumption theorem (FCT). Fig. 18 should be compared to Fig. 2
of Buzzoni (1999). The line with square dots along it is reproduced from Buzzoni's Fig.
2. The other lines show the dependence of the ratio of the Post-MS to MS contribution
to the bolometric flux for different models. The heavy lines correspond to the Salpeter
IMF models. The thin lines to the Scalo IMF models. The solid lines correspond to
P track models, whereas the dashed lines correspond to the G track models. The G
track model for the Salpeter IMF (heavy dashed line) is in quite good agreement with
Buzzoni's model for t > 5 Gyr.
Fig. 19 (after Buzzoni's Fig. 1) plots the M/LV ratio vs. the Post-MS to MS contribu-
tion in the V band. The open dots correspond to the models shown in Buzzoni's Fig. 1.
The solid dot is Buzzoni's model marked B in his Fig. 1. The solid triangles correspond
to our Z⊙ SSP models for various stellar atlas using the P tracks and the Salpeter IMF.
The open triangles are for the same models but for the Scalo IMF. The solid pentagons
represent the G track models for the Salpeter IMF and the open pentagons the same
models but for the Scalo IMF. The three solid squares joined by a line represent sub-
solar metallicity models for the P tracks and the Salpeter IMF. The three open squares
joined by a line are for identical models using the Scalo IMF. Fig. 19 shows clearly that
the position of points representing various models in this diagram is a strong function
of the stellar IMF, the set of evolutionary tracks, and the chemical composition of the
stellar population. It may be too simplistic to attribute the dispersion of the points to a
violation of the FCT (Buzzoni 1999).
11.3. Uncertainties in the stellar IMF
It is constructive to compare models computed for identical ingredients except for the
stellar IMF. Fig. 21 shows the results of such a comparison. Brightness and color
differences with respect to the Salpeter IMF model SSP are shown vs. galaxy age in the
galaxy rest frame (LHS panels) and vs. redshift in the observer frame (RHS panels) for
SSP P track models computed for the following IMFs: Scalo (1986, solid line), Miller &
Scalo (1979, short dashed line), and Kroupa et al. (1993, long dashed line).
Fig. 22 compares in the same format as before the results of using different solar
metallicity stellar libraries for a P track SSP model. Brightness and color differences
with respect to the SSP model computed with the Pickles (1998) stellar atlas are shown
vs. galaxy age in the galaxy rest frame (LHS panels) and vs. redshift in the observer
frame (RHS panels) for SSP P track models computed for the following stellar libraries:
Extended Gunn & Stryker atlas used by BC93 (solid line), LCB97 uncorrected atlas
(short dashed line), and LCB97 corrected atlas (long dashed line).
11.4. Different chemical composition
In this section we explore the differences between SSP models for non-solar composition
and the solar case. Fig. 23 shows the brightness and color differences with respect to
the SSP model for Z = Z⊙. All models shown in this figure are for the P tracks and use
the LCB97 stellar atlas. The lines in this figure have the following meaning: Z = 0.0001
(solid line), Z = 0.0004 (short dashed line), Z = 0.004 (long dashed line), Z = 0.008 (dot
- short dashed line), Z = 0.05 (short dash - long dashed line), and Z = 0.1 (dot - long
dashed line).
22
Gustavo Bruzual A.: Population Synthesis at low and high z
Figure 17. Ratio of the fractional contribution by different stellar groups in the G track
model to that in the P track model.
Ratio of the Post-MS to MS contribution to the bolometric flux vs. age for
Figure 18.
different models. This figure should be compared to Fig. 2 of Buzzoni (1999). The line with
square dots along it is reproduced from Buzzoni's Fig. 2. The heavy lines correspond to the
Salpeter IMF models. The thin lines to the Scalo IMF models. The solid lines correspond to P
track models, whereas the dashed lines correspond to the G track models. The G track model
for the Salpeter IMF (heavy dashed line) is in quite good agreement with Buzzoni's model for
t > 5 Gyr.
Gustavo Bruzual A.: Population Synthesis at low and high z
23
Figure 19. M/LV ratio vs. the Post-MS to MS contribution in the V band. This figure should
be compared to Fig. 1 of Buzzoni (1999). The open dots correspond to the models shown in
Buzzoni's Fig. 1. The solid dot is Buzzoni's model marked B in his Fig. 1. The solid triangles
correspond to our Z⊙ SSP models for various stellar atlas using the P tracks and the Salpeter
IMF. The open triangles are for the same models but for the Scalo IMF. The solid pentagons
represent the G track models for the Salpeter IMF and the open pentagons the same models
but for the Scalo IMF. The three solid squares joined by a line represent sub-solar metallicity
models for the P tracks and the Salpeter IMF. The three open squares joined by a line are for
identical models using the Scalo IMF.
11.5. Different history of chemical evolution
Fig. 24 shows three possible chemical evolutionary histories, Z(t), quick (ZQ, short
dashed line), linear (ZL, solid line), and slow (ZS, long dashed line), that reach Z =
0.1 = 5 × Z⊙ at 15 Gyr. The dotted lines indicate Z⊙ and tg = 12 Gyr. I have computed
models for a SFR Ψ(t) ∝ exp(−t/τ ), with τ = 5 Gyr, which evolve chemically accordingly
to the lines shown in Fig. 24. The difference in brightness and color of these three models
with respect to a Z = Z⊙ model for the same SFR are shown in Fig. 25. The meaning
of the lines is as follows: Z(t) = ZQ (short dashed line), Z(t) = ZL (solid line), and
Z(t) = ZS (long dashed line).
11.6. Evolution in the observer frame at various cosmological epochs
In Figs. 26 to 31, I summarize the range of values expected in the measured (V − R)
and (V − K) colors in the observer frame at various redshifts z as a function of galaxy
age. In these figures, the panel marked T RACKS shows the range of colors obtained for
solar metallicity SSP models computed using the Pickles empirical stellar atlas with the
Salpeter IMF for the P and the G tracks. In the panel marked IM F I show Z = Z⊙ SSP
24
Gustavo Bruzual A.: Population Synthesis at low and high z
Figure 20. Difference in B magnitude and B − V and V − K color between a G track model
SSP and a P track model SSP, vs. age in the galaxy rest frame (LHS panels) and vs. redshift
in the observer frame (RHS panels).
Figure 21. Difference in B magnitude and B − V and V − K color for various IMF, P track
model SSP's, with respect to the Salpeter IMF model, vs. galaxy age in the galaxy rest frame
(LHS panels) and vs. redshift in the observer frame (RHS panels).
Gustavo Bruzual A.: Population Synthesis at low and high z
25
Figure 22. Difference in B magnitude and B − V and V − K color for P track model SSP's
computed for various Z = Z⊙ stellar libraries, with respect to the SSP model computed with
the Pickles (1998) stellar atlas, vs. galaxy age in the galaxy rest frame (LHS panels) and vs.
redshift in the observer frame (RHS panels).
Figure 23. Differences between SSP models for non-solar composition and the solar case, vs.
galaxy age in the galaxy rest frame (LHS panels) and vs. redshift in the observer frame (RHS
panels).
26
Gustavo Bruzual A.: Population Synthesis at low and high z
Figure 24. Three possible chemical evolutionary histories, Z(t), quick (ZQ, short dashed line),
linear (ZL, solid line), and slow (ZS, long dashed line), that reach Z = 0.1 = 5 × Z⊙ at 15 Gyr.
Figure 25. Difference in brightness and color of three models that evolve chemically according
to Fig. 24 with respect to a Z = Z⊙ model for the same SFR, vs. galaxy age in the galaxy rest
frame (LHS panels) and vs. redshift in the observer frame (RHS panels).
Gustavo Bruzual A.: Population Synthesis at low and high z
27
models computed for the P tracks, the Pickles stellar atlas, and the Salpeter, the Scalo,
and the Miller-Scalo IMFs. The panel marked SEDs shows the evolution of Z = Z⊙
SSP models computed with the P tracks and the Salpeter IMF, using the empirical
Gunn-Stryker and Pickles stellar libraries, as well as the original and corrected versions
of the LCB atlas for Z = Z⊙. The panel marked SF R shows the evolution of an SSP
model together with a model in which stars form at a constant rate during the first Gyr
in the life of the galaxy (1 Gyr model), both computed with the P tracks, Salpeter IMF,
and the Pickles stellar library. The panel marked Z shows the range of colors covered
by SSP models of metallicity Z = 0.004, 0.008 and 0.02 (solar), computed with the P
tracks and the Salpeter IMF. In the solar case, I repeat the models shown in the panel
marked SEDs. The panel marked ALL summarizes the results of the previous panels.
The reddest color obtained at any age in the previous 5 panels is shown as the top solid
line. The bluest color is shown as a dotted line. The average color is indicated by the
solid line between these two extremes. The 1 Gyr model is shown as a dashed line to
show the dominant effects of star formation in galaxy colors.
Figs. 26 and 27 show the range of values expected in (V − R) and (V − K) in the
observer frame at z = 0 as a function of galaxy age. The maximum age allowed is the
age of the universe, tu = 13.5 Gyr at z = 0 using H0 = 65 km s−1 Mpc−1, Ω = 0.10.
Figs. 28 and 29 show the same quantities but for galaxies seen at z = 1.552. The age of
the universe for this z in this cosmology is tu = 4.6 Gyr. Figs. 30 and 31 correspond to
z = 3, in this case tu = 2.7 Gyr. From Figs. 26 to 31 I conclude that metallicity Z and
the SFR are the most dominant factors determining the range of allowed colors.
The horizontal lines shown across the panels in Figs. 28 and 29 indicate the color ±σ
of the galaxy LBDS 53W091 observed by Spinrad et al. (1997). Our models reproduce
the colors of this galaxy at an age close to 1.5 Gyr.
Figs. 32 to 35 are based on the panel marked ALL of Figs. 26 to 29, and similar figures
for (V − U ), (V − B), (V − I), and (V − J) not shown in this work. To build these figures
I have subtracted from each line in the previous figures, the color of the Z = Z⊙ SSP
model computed with the P tracks, the Salpeter IMF, and the Pickles stellar library. We
conclude that the evolution of (V − R) is less model dependent than for any other color
shown in these figures.
Figs. 36 to 39 are also based on the panel marked ALL of Figs. 26 to 29 and similar
figures for other values of z not shown in this work. Figs. 36 to 39 show the evolution in
time of (V − U ), (V − B), (V − R), and (V − K) in the observer frame for several values
of the redshift z. The color of the Z = Z⊙ SSP model computed with the P tracks, the
Salpeter IMF, and the Pickles stellar library has been subtracted from the lines in the
previous figures. Again, (V − R) shows less variations with model than the other colors.
12. Summary and Conclusions
Present population synthesis models show reasonable agreement with the observed
spectrum of stellar populations of various ages and metal content. Differences in results
from different codes can be understood in terms of the different ingredients used to
build the models and do not necessarily represent violations of physical principles by
some of these models. However, inspection of Fig. 20 shows that two different sets
of evolutionary tracks for stars of the same metallicity produce models that at early
ages differ in brightness and color from 0.5 to 1 mag, depending on the specific bands.
The differences decrease at present ages in the rest frame, but are large in the observer
frame at z > 2. Thus any attempt to date distant galaxies, for instance, based on fitting
observed colors to these lines will produce ages that depend critically on the set of models
28
Gustavo Bruzual A.: Population Synthesis at low and high z
Figure 26.
(V − R) vs. time in the observer frame at z = 0. See §11.6 for details.
Figure 27.
(V − K) vs. time in the observer frame at z = 0. See §11.6 for details.
Gustavo Bruzual A.: Population Synthesis at low and high z
29
Figure 28.
(V − R) vs. time in the observer frame at z = 1.552. See §11.6 for details.
Figure 29.
(V − K) vs. time in the observer frame at z = 1.552. See §11.6 for details.
30
Gustavo Bruzual A.: Population Synthesis at low and high z
Figure 30.
(V − R) vs. time in the observer frame at z = 3. See §11.6 for details.
Figure 31.
(V − K) vs. time in the observer frame at z = 3. See §11.6 for details.
Gustavo Bruzual A.: Population Synthesis at low and high z
31
Figure 32. Color vs. time in the observer frame at z = 0. The color of the Z = Z⊙ SSP
model computed with the P tracks, the Salpeter IMF, and the Pickles stellar library have been
subtracted from each line. See §11.6 for details.
Figure 33. Color vs. time in the observer frame at z = 1. The color of the Z = Z⊙ SSP
model computed with the P tracks, the Salpeter IMF, and the Pickles stellar library have been
subtracted from each line. See §11.6 for details.
32
Gustavo Bruzual A.: Population Synthesis at low and high z
Figure 34. Color vs. time in the observer frame at z = 2. The color of the Z = Z⊙ SSP
model computed with the P tracks, the Salpeter IMF, and the Pickles stellar library have been
subtracted from each line. See §11.6 for details.
Figure 35. Color vs. time in the observer frame at z = 3. The color of the Z = Z⊙ SSP
model computed with the P tracks, the Salpeter IMF, and the Pickles stellar library have been
subtracted from each line. See §11.6 for details.
Gustavo Bruzual A.: Population Synthesis at low and high z
33
(V − U ) vs. time in the observer frame for various values of z. The color of
Figure 36.
the Z = Z⊙ SSP model computed with the P tracks, the Salpeter IMF, and the Pickles stellar
library have been subtracted from each line. See §11.6 for details.
(V − B) vs. time in the observer frame for various values of z. The color of
Figure 37.
the Z = Z⊙ SSP model computed with the P tracks, the Salpeter IMF, and the Pickles stellar
library have been subtracted from each line. See §11.6 for details.
34
Gustavo Bruzual A.: Population Synthesis at low and high z
(V − R) vs. time in the observer frame for various values of z. The color of
Figure 38.
the Z = Z⊙ SSP model computed with the P tracks, the Salpeter IMF, and the Pickles stellar
library have been subtracted from each line. See §11.6 for details.
(V − K) vs. time in the observer frame for various values of z. The color of
Figure 39.
the Z = Z⊙ SSP model computed with the P tracks, the Salpeter IMF, and the Pickles stellar
library have been subtracted from each line. See §11.6 for details.
Gustavo Bruzual A.: Population Synthesis at low and high z
35
which is used. Note that from z of 3 to 3.5 (V − K) in the two models differs by more
than 1 mag. This difference is produced by the corresponding difference between the
models seen in the rest frame at 10 Myr. From Figs. 15 and 16 these differences can be
understood in terms of the different contribution of the same stellar groups to the total
V and K flux in the two models.
Even though at the present age models built with different IMFs show reasonably
similar colors and brightness, the early evolution of these models is quite different at
early ages (Fig. 21), resulting in larger color differences in the observer frame at z >
2. Thus, the more we know about the IMF, the better the model predictions can be
constrained. The small color differences seen in the rest frame when different stellar
libraries of the same metallicity are used, are magnified in the observer frame (Fig. 22).
When the k correction brings opposing flux differences into each filter, the difference in
the resulting color is enhanced. Fig. 23 shows the danger of interpreting data for one
stellar system with models of the wrong metallicity. The color differences between these
models, especially in the observer frame, are so large as to make any conclusion thus
derived very uncertain.
It is common practice to use solar metallicity models when no information is available
about the chemical abundance of a given stellar system. Galaxies evolving according the
Z(t) laws of Fig. 24 show color differences with respect to the Z = Z⊙ model which are
not larger than the differences introduced by the other sources of uncertainties discussed
so far. Hence, the solar metallicity approximation may be justified in some instances.
The color differences between the chemically inhomogeneous composite population and
the purely solar case (Fig. 25), are much smaller than the ones shown in Fig. 23 for
chemically homogeneous SSPs.
Figs. 26 to 39 indicate that some colors, especially (V − R), when measured in the
observer frame are less sensitive to model predictions than other colors. From Figs. 26
to 31, metallicity Z and the SFR are the most dominant factors determining the range
of allowed colors.
I expect that through these simple examples the reader can get a feeling of the kind of
uncertainties introduced by the many ingredients entering the stellar population synthesis
problem, and that he or she will be motivated to try his or her own error estimates when
using these models.
REFERENCES
Allard, F., & Hauschildt, P.H. 1995, ApJ, 445, 433
Alongi, M., Bertelli, G., Bressan, A., Chiosi, C., Fagotto, F., Greggio, L., & Nasi, E. 1993,
A&AS, 97, 851
Arag´on-Salamanca, A., Ellis, R.S.E., Couch, W. J., Carter, D. 1993, MNRAS, 262, 764A
Arimoto, N., & Yoshii, Y. 1987, A&A, 173, 23
Barbuy, B., Ortolani, S., Bica, E., Renzini, A., & Guarnieri, M.D. 1997, in IAU Symp. 189,
Fundamental Stellar Parameters: Confrontation Between Observation and Theory, eds. J.
Davis, A. Booth & T. Bedding, Kluwer Acad. Pub., p. 203
Bender, R., Ziegler, B., & Bruzual A., G. 1996, ApJ Letters, 463, L51
Bessell, M.S., Brett, J., Scholtz, M., & Wood, P. 1989, A&AS, 77, 1
-- -- -- . 1991, A&AS, 89, 335
Bica, E., & Alloin, D. 1986, A&A, 162, 21
-- -- -- . 1987, A&A, 186, 49
Bica, E., Alloin, D., & Schmitt, H. 1994, A&A, 283, 805
36
Gustavo Bruzual A.: Population Synthesis at low and high z
Bica, E., Alloin, D., Bonatto, C., Pastoriza, M.G., Jablonka, P., Schmidt, A., & Schmitt, H.R.
1996a, in A Data Base for Galaxy Evolution Modeling, eds. C. Leitherer et al., PASP, 108,
996
Bica, E., Clari´a, J.J., Dottori, H., Santos Jr., J.F.C., Piatti, A. E. 1996b, ApJS, 102, 57
Bressan, A., Chiosi, C., & Fagotto, F. 1994, ApJS, 94, 63
Bressan, A., Fagotto, F., Bertelli, G., & Chiosi, C. 1993, A&AS, 100, 647
Bruzual A., G. 1998,
in The Evolution of Galaxies on Cosmological Time scales, eds. J.E.
Beckman and T.J. Mahoney, ASP Conference Series, Vol. 187, p. 245
-- -- -- . 1999, in The Hy-Redshift Universe: Galaxy Formation and Evolution at High Redshift,
eds. A. J. Bunker and W. J. M. van Breugel, ASP Conference Series, Vol. 193, p. 121
-- -- -- . 2000, in Euroconference on The Evolution of Galaxies, I- Observational Clues, eds. J.M.
V´ılchez, G. Stasinska, and E. P´erez, Kluwer Academic Publisher, in press
Bruzual A., G., Barbuy, B., Ortolani, S., Bica, E., Cuisinier, F., Lejeune, T., & Schiavon, R.
1997, AJ, 114, 1531
Bruzual A., G. & Charlot, S. 1993, ApJ, 405, 538 (BC93)
-- -- -- . 2000, ApJ, in preparation (BC2000)
Burstein, D., Bertola, F., Buson, L.M., Faber, S.M., and Lauer, T.R. 1988, ApJ, 328, 440
Buzzoni, A. 1989, ApJS, 71, 817
-- -- -- . 1999, in IAU Symposium No. 183 Cosmological Parameters and the Evolution of the
Universe, ed. K. Sato, Dordrecht: Kluwer, p. 134
Charlot, S., and Bruzual A., G. 1991, ApJ, 367, 126 (CB91)
Cool, A.M. 1997, in Advances in Stellar Evolution, eds. R. T. Rood and A. Renzini, Cambridge
University Press, p. 191
D'Antona, F. 1999, in The Galactic Halo: from Globular Clusters to Field Stars, 35th Liege Int.
Astroph. Colloquium, astro-ph/9910312
Eggen, O.J., and Sandage, A.R. 1964, ApJ, 140, 130
Fagotto, F., Bressan, A., Bertelli, G., & Chiosi, C. 1994a, A&AS, 100, 647
-- -- -- . 1994b, A&AS, 104, 365
-- -- -- . 1994c, A&AS, 105, 29
Fluks, M. et al. 1994, A&AS, 105, 311
Fritze-v.Alvensleben, U. & Gerhard, O.E. 1994, A&A, 285, 751
Gilliland, R.L., Brown, T.M., Duncan, D.K., Suntzeff, N.B., Wesley Lockwood, G., Thompson,
D.T., Schild, R.E., Jeffrey, W.A., and Penprase, B.E., 1991, AJ, 101, 541
Girardi, L., Bressan, A., Chiosi, C., Bertelli, G., & Nasi, E. 1996, A&AS, 117, 113
Girardi, L., Bressan, A., Bertelli, G., & Chiosi, C. 2000, A&AS, 141, 371
Greggio, L., and Renzini, A., 1990, ApJ, 364, 35
Guarnieri, M.D., Ortolani, S., Montegriffo, P., Renzini, A.,
Guiderdoni, B. & Rocca-Volmerange, B. 1987, A&A, 186, 1
Gunn, J.E., and Stryker, L.L., 1983, ApJS, 52, 121
Iglesias, C.A., Rogers, F.J., & Wilson, B.G. 1992, ApJ, 397, 717
Janes, K.A., 1985, in Calibration of Fundamental Stellar Quantities, IAU Symposium No. 111,
D.S. Hayes, L.E. Pasinetti, and A.G. Davis Philip, (Dordrecht: Reidel), 361
Janes, K.A., and Smith, G.H., 1984, AJ, 89, 487
Kaluzny, J. 1997, A&AS, 121, 455
King, I.R., Anderson, J., Cool, A.M., Piotto, G. 1998, ApJ, 492, L37
Kroupa, P., Tout, C.A., & Gilmore, G. 1993, MNRAS, 262, 545
Kurucz, R. 1995, private communication
Lejeune, T., Cuisinier, F., & Buser, R. 1997, A&AS, 125, 229 (LCB97)
Gustavo Bruzual A.: Population Synthesis at low and high z
37
-- -- -- . 1998, A&AS, 130, 65 (LCB98)
Metcalfe, N., Shanks, T., Fong, R., Gardner, J., Roche, N. 1996, IAU Symp. 171, p. 225
Micela, G., Sciortino, S., Vaiana, G.S., Schmitt, J.H.M.M., Stern, R.A., Harnden, F.R., Jr., and
Rosner, R., 1988, ApJ, 325, 798
Miller, G.E. & Scalo, J.M. 1979, 41, 513
Peterson, D.M., & Solensky, R. 1988, ApJ, 333, 256
Pickles, A.J. 1998, PASP, 110, 863
Pozzetti, L., Bruzual A., G., Zamorani, G. 1996, MNRAS, 281, 953
Racine, R., 1971, ApJ, 168, 393
Renzini, A., 1981, Ann. Phys. Fr., 6, 87
Salpeter, E.E. 1955, ApJ, 121, 161
Santos, J.F.C.Jr., Bica, E., Dottori, H., Ortolani, S., & Barbuy, B. 1995, A&A, 303, 753
Scalo, J.M. 1986, Fund. Cosmic Phys, 11, 1
Spinrad, H., Dey, A., Stern, D., Dunlop, J., Peacock, J., Jim´enez, R., Windhorst, R. 1997, ApJ,
484, 581
Stanford, S.A., Eisenhardt, P.R., & Dickinson, M. 1995, ApJ, 450, 512
-- -- -- . 1998, ApJ, 492, 461
Upgren, A.R., 1974, ApJ, 193, 359
Upgren, A.R., and Weis, E.W., 1977, AJ, 82, 978
Worthey, G. 1994, ApJS, 95, 107
Worthey, G., Faber, S.M., Gonz´alez, J.J., & Burstein, D. 1994, ApJS, 94, 687
|
astro-ph/0405138 | 1 | 0405 | 2004-05-07T09:33:27 | A Survey of z>5.7 Quasars in the Sloan Digital Sky Survey III: Discovery of Five Additional Quasars | [
"astro-ph"
] | We present the discovery of five new quasars at z>5.7, selected from the multicolor imaging data of the Sloan Digital Sky Survey (SDSS). Three of them, at redshifts 5.93, 6.07, and 6.22, were selected from ~1700 deg^2 of new SDSS Main Survey imaging in the Northern Galactic Cap. An additional quasar, at redshift 5.85, was discovered by coadding the data obtained in the Fall Equatorial Stripe in the SDSS Southern Survey Region. The fifth object, at redshift 5.80, is selected from a non-standard SDSS scan in the Southern Galactic Cap outside the Main Survey area. The spectrum of SDSS J162331.81+311200.5 (z=6.22) shows a complete Gunn-Peterson trough at z_abs > 5.95, similar to the troughs detected in other three z>6.2 quasars known. We present a composite spectrum of the z>5.7 quasars discovered in the SDSS to date. The average emission line and continuum properties of z~6 quasars exhibit no significant evolution compared to those at low redshift. Using a complete sample of nine z>5.7 quasars, we find that the density of quasars with M_1450 < -26.7 at z~6 is (6+/-2) x 10^-10 per Mpc^3 consistent with our previous estimates. The luminosity distribution of the sample is fit with a power law luminosity function Psi(L) ~ L^(-3.2+/-0.7), somewhat steeper than but consistent with our previous estimates. | astro-ph | astro-ph |
AJ, in press (Aug, 2004)
A Survey of z > 5.7 Quasars in the Sloan Digital Sky Survey III: Discovery of
Five Additional Quasars1
Xiaohui Fan2,3 , Joseph F. Hennawi3,4, , Gordon T. Richards4 , Michael A. Strauss4 , Donald P.
Schneider5 , Jennifer L. Donley2 , Jason E. Young2 , James Annis6 , Huan Lin6 , Hubert Lampeitl6 ,
Robert H. Lupton4 , James E. Gunn4 , Gillan R. Knapp4 , W. N. Brandt5 , Scott Anderson6 , Neta
A. Bahcall4 , Jon Brinkmann7 , Robert J. Brunner8 , Masataka Fukugita7 , Alexander S. Szalay9 ,
Gyula P. Szokoly10 , Donald G. York11
ABSTRACT
We present the discovery of five new quasars at z > 5.7, selected from the multicolor
imaging data of the Sloan Digital Sky Survey (SDSS). Three of them, at redshifts 5.93,
1Based on observations obtained with the Sloan Digital Sky Survey, and with the Apache Point Observatory
3.5-meter telescope, which is owned and operated by the Astrophysical Research Consortium; and with the MMT
Observatory, a joint facility of the University of Arizona and the Smithsonian Institution, with the Univeristy of
Arizona 2.3-meter Bok Telescope, with the Kitt Peak National Observatory 4-meter Mayall Telescope, with the
6.5-meter Landon Clay Telescope at the Las Campanas Observatory, a collaboration between the Observatories
of the Carnegie Institution of Washington, University of Arizona, Harvard University, University of Michigan,
and Massachusetts Institute of Technology , and with the Hobby-Eberly Telescope, which is a joint pro ject of
the University of Texas at Austin, the Pennsylvania State University, Stanford University, Ludwig-Maximillians-
Universitat Munchen, and Georg-August-Universitat Gottingen.
2Steward Observatory, The University of Arizona, Tucson, AZ 85721
3Visiting Astronomer, Kitt Peak National Observatory, National Optical Astronomy Observatory, which is
operated by the Association of Universities for Research in Astronomy, Inc. (AURA) under cooperative agreement
with the National Science Foundation.
4Princeton University Observatory, Princeton, NJ 08544
5Department of Astronomy and Astrophysics, The Pennsylvania State University, University Park, PA 16802
6Fermi National Accelerator Laboratory, P. O. Box 500, Batavia, IL 60510
6University of Washington, Department of Astronomy, Box 351580, Seattle, WA 98195
7Apache Point Observatory, P. O. Box 59, Sunspot, NM 88349-0059
7 Institute for Cosmic Ray Research, University of Tokyo, Midori, Tanashi, Tokyo 188-8502, Japan
8Dept. of Astronomy & NCSA, University of Illinois, 1002 W. Green Street, Urbana, IL 61801
9Department of Physics and Astronomy, The Johns Hopkins University, Baltimore, MD 21218, USA
10Max-Planck-Institut fur extraterrestrische Physik, Postfach 1312, 85741 Garching, Germany
11University of Chicago, Astronomy & Astrophysics Center, 5640 S. Ellis Ave., Chicago, IL 60637
– 2 –
6.07, and 6.22, were selected from ∼ 1700 deg2 of new SDSS Main Survey imaging
in the Northern Galactic Cap. An additional quasar, at redshift 5.85, was discovered
by coadding the data obtained in the Fall Equatorial Stripe in the SDSS Southern
Survey Region. The fifth ob ject, at redshift 5.80, is selected from a non-standard
SDSS scan in the Southern Galactic Cap outside the Main Survey area. The spectrum
of SDSS J162331.81+311200.5 (z = 6.22) shows a complete Gunn-Peterson trough at
zabs > 5.95, similar to the troughs detected in other three z ∼> 6.2 quasars known.
We present a composite spectrum of the z > 5.7 quasars discovered in the SDSS to
date. The average emission line and continuum properties of z ∼ 6 quasars exhibit no
significant evolution compared to those at low redshift. Using a complete sample of
nine z > 5.7 quasars, we find that the density of quasars with M1450 < −26.7 at z ∼ 6
is (6 ± 2) × 10−10 Mpc−3 (H0 = 65 km s−1Mpc−1 , Ω = 0.35 and Λ = 0.65), consistent
with our previous estimates. The luminosity distribution of the sample is fit with a
power law luminosity function Ψ(L) ∝ L−3.2±0.7 , somewhat steeper than but consistent
with our previous estimates.
Subject headings: quasars: general; quasars: emission line; quasars: absorption lines
1.
Introduction
This paper is the third in a series presenting i-dropout z ∼> 5.7 quasars selected from the
multicolor imaging data of the Sloan Digital Sky Survey (SDSS; York et al. 2000, Stoughton et
al. 2002, Abaza jian et al. 2003, 2004). In Fan et al. (2000) and in the first two papers of this
series (Fan et al. 2001c, Paper I, Fan et al. 2003, Paper II), we presented the discovery of seven
luminous quasars at z = 5.74 − 6.42, selected from ∼ 2900 deg2 of SDSS imaging in the Northern
Galactic Cap. In this paper, we describe the discovery of five quasars at z = 5.80, 5.85, 5.93, 6.07
and 6.22, respectively. The scientific ob jectives, photometric data reduction, target selection and
follow-up observation procedures are described in detail in Paper I. Three of the new quasars
were selected using the same color selection procedures described in Paper II, as outlined briefly
in §2.1. One ob ject was selected using the co-added catalog from multi-epoch imaging of the Fall
Equatorial Stripe in the SDSS Southern Survey region. The co-addition procedures are described
in §2.2. The final ob ject was selected in a non-standard SDSS scan in the Southern Galactic
Cap outside the main survey area. We present the spectroscopic follow-up observations of the
i-dropout candidates and the photometric and spectroscopic properties of the newly-discovered
quasars in §3. Combining the quasars in this paper and those presented in Fan et al. (2002) and
in Papers I and II, we construct the composite spectrum of z ∼ 6 quasars, and compare it with
the average quasar spectrum at low redshift (§4). Finally, we update our estimate of the evolution
of the high-redshift quasar luminosity function in §5.
Following the previous papers in this series, we use two cosmologies to present our results: (1)
– 3 –
H0 = 50 km s−1 Mpc−1 , ΩΛ = 0 and ΩM = 1 (Ω-model); (2) H0 = 65 km s−1 Mpc−1 , ΩΛ = 0.65
and ΩM = 0.35 (Λ-model). Coincidentally, the luminosity and number densities measured
under our Λ-model are within 1% of those in the WMAP (Spergel et al. 2003) cosmology with
H0 = 71 km s−1 Mpc−1 , ΩΛ = 0.73 and ΩM = 0.27.
2. Candidate Selection and Identification
The Sloan Digital Sky Survey is using a dedicated 2.5m telescope and a large format CCD
camera (Gunn et al. 1998) at the Apache Point Observatory in New Mexico to obtain images in
five broad bands (u, g , r , i and z , centered at 3551, 4686, 6166, 7480 and 8932 A, respectively;
Fukugita et al. 1996) of high Galactic latitude sky in the Northern Galactic Cap. About 6000 deg2
of sky have been imaged at the time of this writing (Mar 2004). The imaging data are processed
by the astrometric pipeline (Pier et al. 2003) and photometric pipeline (Lupton et al. 2001), and
are photometrically calibrated to a standard star network (Smith et al. 2002, see also Hogg et al.
2001). In addition to the Main Survey, the SDSS also obtains multi-epoch imaging in ∼ 270 deg2
along the Celestial Equator in the Southern Galactic Cap (the SDSS Southern Survey, York et al.
2000). At the time of this writing, the Southern Survey region has been imaged between 5 and 15
times, depending on the RA of the field. The multi-epoch data are used to study variable ob jects
and are co-added to reach fainter limiting magnitudes. We will use imaging from both the Main
Survey and the Southern Survey to search for quasar candidates.
2.1. Selection in the Main Survey Area
The SDSS quasar target selection pipeline (Richards et al. 2002) selects only quasar
candidates at z ∼< 5.5. The highest-redshift ob ject discovered in the SDSS spectroscopic survey is
at z = 5.4 (Anderson et al. 2001). At higher redshift, quasars become i-dropout ob jects; their
selections require additional follow-up observations. Papers I and II present the results from a
survey of i-dropout candidates selected from ∼ 2900 deg2 of SDSS Main Survey imaging carried
out in the Springs of 2000, 2001 and 2002. In Spring 2003, we searched for i-dropout quasar
candidates in 37 new SDSS imaging runs. These imaging data were taken between 10 October
2002 (Run 3358) and 29 April 2003 (Run 3919). We used the same criterion to decide which
photometric runs to include in the i-dropout survey as in Papers I and II: the z band image
quality, measured by the psfWidth parameter (= 1.06 FWHM for a Gaussian profile) in the fourth
column of the SDSS camera, should be better than 1.8′′ . The median seeing in the i and z bands
is ∼ 1.4′′ for the entire survey area (Abaza jian et al. 2003). There is overlap between adjacent
SDSS strips and stripes (York et al. 2000), meaning that these 37 new runs overlap somewhat
with the area covered in Papers I and II. Taking these overlaps into account, we find that the
total new area of the sky covered by these runs is 1708 deg2 , bringing the combined sky coverage
of Papers I, II and this paper to 4578 deg2 .
– 4 –
We applied the same color selection criteria as in Paper II (see Figure 1 and 2 in Paper II)
to the new SDSS imaging data to selection z > 5.7 quasar candidates. A total of 80 i-dropout
candidates were selected in the main survey area. The photometric and spectroscopic follow-up
observations were carried out over a number of nights between December 2002 and June 2003.
Independent z photometry was carried out using the Seaver Prototype Imaging camera (SPICAM)
in the SDSS z filter on the ARC 3.5m telescope at the Apache Point Observatory. J photometry
was carried out using the 256 × 256 NICMOS imager on Steward Observatory’s 2.3m Bok Telescope
at Kitt Peak, and using the GRIM II instrument (the near infrared GRIsm spectrometer and
IMager), also on the ARC 3.5m. The spectroscopic follow-up observations were obtained using
the Red-Channel Spectrograph on the MMT 6.5m telescope on Mt. Hopkins, the Double Imaging
Spectrograph (DIS) on the ARC 3.5m, the Multi-Aperture Red Spectrograph (MARS) on the
4-m telescope on Kitt Peak, and the Low Resolution Spectrograph (LRS, Hill et al. 1998) on the
Hobby-Eberly Telescope.
2.2. Selection in the SDSS Southern Survey Region
The SDSS Main Survey imaging consists of single epoch observations with exposure times
of 54.1 seconds. For the average seeing conditions (FWHM ∼ 1.4′′ ), the 5-σ limiting magnitudes
in the i and z bands are 22.5 and 20.5, respectively (Stoughton et al. 2002, Abaza jian et al.
2003). At this depth, the SDSS Main Survey only allows selection of i-dropout candidates at
z < 20.2, and therefore is only sensitive to the most luminous quasars at z ∼ 6 (MB ∼< −27).
The multi-epoch imaging obtained in the SDSS Southern Survey area will eventually provide
photometry ∼ 1.5 magnitudes deeper than the main survey over ∼ 270 deg2 , enabling selection of
much fainter quasar candidates. By the end of 2003, the SDSS Southern Survey Region had been
scanned 5 – 15 times, resulting in a dataset that goes more than one magnitude deeper than the
Main Survey.
In Fall 2002, we used a preliminary co-added catalog of ∼ 5 epoch imaging to select z ∼ 6
quasar candidates. The co-added catalog was generated by matching sources detected in more
than one SDSS run within a radius of 1′′ . The average flux is calculated from the SDSS asinh
magnitudes (Lupton, Gunn & Szalay 1999), weighted by the inverse of the flux variances. Note
that when co-adding at the catalog level, an ob ject is required to be detected in the individual
runs. Therefore, while the co-addition improves the S/N of faint sources, allowing us to select
candidates all the way down to the 6-σ detection limit, it does not allow measurement of ob jects
fainter than the detection limit of single SDSS exposures. Therefore, we can only select candidates
at z < 20.5 − 20.7. We searched for candidates from ∼ 100 deg2 of co-added catalogs. The
color selection criteria in the Southern Survey region are the same as those used in the Main
survey area. A number of faint i-dropout candidates were observed using Magellan II Clay
Telescope and the Boller and Chivens (B&C) spectrograph on 16 Oct 2002. One quasar, SDSS
– 5 –
J000552.34–000655.814 (hereafter SDSS J0005–0006, zAB = 20.5) was discovered at a redshift of
5.85.
2.3. Selection in the Constant-Longitude Scans
The SDSS has obtained some imaging data outside the survey boundary (York et al. 2000;
Finkbeiner et al. 2004), including a few strips in the Southern Galactic Cap that extend to low
Galactic latitude. We have also selected i-dropout candidates in these runs. One new quasar,
SDSS J000239.39+255034.8 (z = 5.80), was selected from Run 4152, a constant-longitude scan (at
l = 110◦ ), observed on 29 Sep 2003. The sample in these scans which go outside the main survey
area is not yet complete, and so we do not include this ob ject in our complete sample.
3. Discovery of Five New Quasars at z > 5.7
Among the 80 i-dropout candidates candidates in the Main Survey area, 15 are false z
band only detections which are most likely cosmic rays; 55 are M or L dwarfs (mostly classified
photometrically based on their red z ∗ − J colors); and 7 are likely T dwarfs. Several ob jects
still lack proper infrared spectroscopy, so the T dwarf classification is still preliminary. Three
of the candidates are identified as quasars at z > 5.7: SDSS J141111.29+121737.4 (z = 5.93,
SDSS J0002+2550 hereafter), SDSS J160254.18+422822.9 (z = 6.07, SDSS J1602+4228), and
SDSS J162331.81+311200.5 (z = 6.22, SDSS J1623+3112). The discovery spectra of the first two
ob jects were obtained using DIS on the ARC 3.5m in April 2003, and the discovery spectrum of
the last ob ject was obtained in June 2003 at MMT using the Red Channel spectrograph. We have
subsequently obtained longer exposures of these three quasars using the Red Channel on MMT
and MARS on the Kitt Peak 4-meter, with total exposure times of 2 – 4 hours each. These spectra
are shown in Figure 2. As discussed above, SDSS J000552.34–000655.8 (z = 5.85) was discovered
in the deep imaging of the SDSS Southern Survey region. Figure 2 shows the discovery spectrum,
a 40-min exposure obtained with the Clay telescope and B&C spectrograph in Oct 2002. The
discovery spectrum of SDSS J000239.39+255034.8 (z = 5.80, §2.3) was obtained at the MMT
using the Red Channel Spectrograph in Nov 2003. This one hour exposure is shown in Figure 2.
The finding charts of the five new quasars are presented in Figure 1. The spectra are
flux-calibrated to match the observed z band photometry. Table 1 presents the photometric
properties of the new quasars, and Table 2 presents the measurements of their continuum
properties. Following Papers I and II, the quantity AB1280 is defined as the AB magnitude of
the continuum at rest-frame 1280A, after correcting for interstellar extinction using the map
14The naming convention for SDSS sources is SDSS JHHMMSS.SS±DDMMSS.S, and the positions are expressed
in J2000.0 coordinates. The astrometry is accurate to better than 0.1′′ in each coordinate.
– 6 –
of Schlegel, Finkbeiner & Davis (1998). We extrapolate the continuum to rest-frame 1450A,
assuming a continuum shape fν ∝ ν −0.5 , to calculate AB1450 . None of the five quasars is detected
in the FIRST (Becker, White & Helfand 1995) or NVSS (Condon et al. 1998) radio surveys. The
discovery of these five new quasars brings the total of z > 5.7 quasars known to twelve, all selected
from SDSS imaging.
3.1. Notes on Individual Ob jects
SDSS J000239.39+255034.8 (z = 5.80). This ob ject is selected in a non-standard SDSS
scan in the Southern Galactic Cap. It is outside the main survey region, and the spectroscopic
follow-up of this region for i-dropout candidates is not yet complete. We therefore do not include
it in the luminosity function calculations below.
SDSS J0002+2550 has i = 21.47 and z = 18.99, but is undetected in 2MASS (Skrutskie
et al. 1997), indicating that J > 16.5 and z − J < 2.5. If it were an L dwarf, it would have a
z − J color of 2.5 or redder, so we targeted this ob ject as a high-redshift quasar candidate. SDSS
J1044–0125 (z = 5.74, Fan et al. 2000) was selected in a similar manner.
At zAB ∼ 19 and M1450 = −27.55 (Λ-model), SDSS J0002+2550 is very luminous. It is
the second brightest quasar known at z > 5.7 so far and provides an excellent target for high
resolution spectroscopic follow-up observations. The redshift of SDSS J0002+2550 is determined
by the locations of the peak of the Lyα emission line and of the OI 1300A emission line, and is
accurate to 0.02. The quasar has a Lyα+NV emission line rest-frame equivalent width of ∼ 60A,
comparable to that of most high-redshift quasars (Fan et al. 2001b).
SDSS J000552.34–000655.8 (z = 5.85). This ob ject is selected in the SDSS Southern Survey
region. At zAB = 20.54 ± 0.10, it is the faintest z > 5.7 quasar in our sample. Although the
discovery spectrum (Figure 2) has low S/N, the strong Lyα and NV emission lines and the strong
Lyman break are clearly visible. The redshift is determined by the locations of the Lyα and NV
emission line peaks, and is accurate to 0.02. The emission lines in this quasar appear to be quite
narrow: the Lyα and NV emission are clearly separated. We estimate a FWHM of about 1500 –
2500 km s−1 for individual lines, with large uncertainties due to the low S/N.
SDSS J141111.29+121737.4 (z = 5.93). The redshift is determined by the locations of the OI
1300A and NV 1240A lines and is accurate to 0.02. The ob ject has a moderately strong Lyα+NV
line (rest-frame equivalent width of ∼ 100A). This ob ject is detected in two overlapping SDSS
runs with consistent photometry (Table 1).
SDSS J160254.18+422822.9 (z = 6.07). The redshift of this ob ject is determined by the peaks
of the Lyα and NV 1240A lines and is accurate to 0.02. A weak OI 1300A line is also detected in
the spectrum. The quasar spectrum shows a number of dark patches in the Lyα and Lyβ forest,
although the S/N of the current spectrum is not sufficient to determine whether there is a short
– 7 –
Gunn-Peterson trough in the spectrum.
Due to its proximity on the sky to the high-redshift cluster candidate CL1603,
SDSS J1602+4228 serendipitously lies in a 28 ks pointed observation (sequence rp800239)
made with the ROSAT Position Sensitive Proportional Counter (PSPC). We do not find any
significant X-ray detection of SDSS J1602+4228 in these data. The 3σ upper limit on its
observed-frame, Galactic absorption-corrected, 0.5–2 keV flux is 5.9 × 10−14 erg cm−2 s−1 ,
adopting a power-law model with a photon index of Γ = 2 and the Galactic column density of
NH = 1.3 × 1020 cm−2 . Given the AB1450 magnitude of SDSS J1602+4228, comparison with
Figure 2 of Vignali et al. (2003) shows that the X-ray upper limit is consistent with X-ray
observations of other z > 4 quasars. The slope of a nominal power law between rest-frame 2500 A
and 2 keV is constrained to be αox > 1.2.
SDSS J162331.81+311200.5 (z = 6.22). This is the highest redshift quasar presented in this
paper, and is the third highest redshift quasar yet known. 15 This quasar has two striking features.
First, it has an extremely strong Lyα emission line. The total equivalent width of Lyα+NV is
∼> 150A in the rest-frame, without taking into account the strong absorption due to the Lyα forest
on the blue side of the Lyα emission. For comparison, Fan et al. (2001b) measured the mean and
standard deviation of the rest-frame Lyα+NV equivalent width of 69.3 ± 18.0A, based on a sample
of ∼ 40 quasars at z ∼ 4 selected from the SDSS (see also Schneider, Schmidt & Gunn 1991 and
§4). The line strength of this quasar is more than a factor of two larger than the average at high
redshift. It is one of the strongest-lined quasars at z > 4 yet known, and the strongest-lined quasar
at z > 5.
Second, SDSS J1623+3112 has a complete Gunn-Peterson (1965) trough. Following Becker et
al. (2001), Fan et al. (2002) and White et al. (2003), we define the transmitted flux ratio as:
T (zabs ) ≡ Df obs
ν E ,
ν /f con
(1 + zabs − 0.1) × 1216A < λ < (1 + zabs + 0.1) × 1216A,
(1)
where f con
is the continuum level extrapolated from the red side of Lyα emission. Using
ν
the spectrum presented in Figure 2, we find that at z = 6.05, the transmitted flux ratio
T (zabs = 5.95 − 6.15) = 0.004 ± 0.008, consistent with zero flux in the Gunn-Peterson trough
region. We also detect a complete Lyβ Gunn-Peterson trough in this quasar. This is the fourth
quasar with a complete Gunn-Peterson trough, after SDSS J1306+0524 (z = 6.28, Becker et al.
2002), SDSS J1148+5251 (z = 6.42, White et al. 2003) and SDSS J1048+4637 (z = 6.18, Fan et
al. 2003).
In al l quasars at z > 6.1, complete Gunn-Peterson troughs are detected, starting from
zabs = 5.95 ± 0.1, and extending to the highest redshift not affected by the quasar proximity effect.
15 SDSS J104845.05+463718.3 was reported to have a redshift of 6.23 in Fan et al. (2003). Subsequent observations
show that it is a Broad Absorption Line (BAL) quasar (Maiolino et al. 2004, Fan et al.
in preparation) and the
original redshift determination is biased; the best redshift estimate is z = 6.18 based on new observations.
– 8 –
The detection of a complete Gunn-Peterson trough in SDSS J1623+3112 further confirms the
rapid transition of the ionization state of the IGM at z ∼ 6 (e.g. Becker et al. 2001, Djorgovski et
al. 2001, Fan et al. 2002, White et al. 2003; see also Songaila 2004 for a different interpretation).
The S/N of the spectrum of SDSS J1623+3112 is considerably lower than that of the other
three quasars, therefore the optical depth limit we are able to place is not yet very strong. In a
subsequent paper, we will present a detailed analysis of the constraints on the evolution of IGM
properties using the five new quasars presented in this paper.
4. Quasar Composite Spectrum at z ∼ 6
The spectral energy distributions of luminous quasars show little evolution out to high
redshift. There is growing evidence from emission line ratio measurements that quasar broad
emission line regions have roughly solar or even higher metallicities at z > 4 (e.g., Hamann &
Ferland 1993, Dietrich et al. 2003a), similar to that in low redshift quasars. Dietrich et al. (2003b)
found the FeII/MgII ratio to have roughly the same value in a sample of z ∼ 5 quasars as at lower
redshift, suggesting that the metallicity of quasar emission line region remains high to even earlier
epochs.
The sample of twelve quasars at z > 5.7 from the SDSS provides the first opportunity to
study the evolution of quasar spectral properties at z ∼ 6, less than 1 Gyr after the Big Bang and
only 700 million years from the first star formation in the Universe (Kogut et al. 2003, Spergel et
al. 2003). Optical and infrared spectroscopy of some z ∼ 6 quasars already indicates a lack of
evolution in the spectral properties of these luminous quasars: Pentericci et al. (2002) show that
the CIV/NV ratio in two z ∼ 6 quasars are indicative of supersolar metallicity in these systems.
Freudling et al. (2003) and Barth et al. (2003) detected strong FeII emission in the spectra of four
z ∼ 6 SDSS quasars. In addition, the optical-to-X-ray flux ratios and X-ray continuum shapes
show at most mild evolution from low redshift (e.g. Brandt et al. 2002, Vignali et al. 2003). These
results, if confirmed with a larger sample, suggest that the accretion disk and photoionization
structure of quasars reached maturity very early on and are probably insensitive to the host galaxy
environment.
Figure 3 shows the composite of eleven of our twelve z > 5.7 quasar spectra. We omit SDSS
J0005-0006 due to its low S/N. To produce the composite, we simply redshift all the spectra to
zero, scale the continuum level of each quasar based on its m1450 magnitude, and average all the
scaled spectra with equal weighting. The composite covers rest-frame wavelengths from 1100A
to 1450A. Also plotted in Figure 3 is the low-redshift SDSS quasar composite of Vanden Berk et
al. (2001). The effective redshift of the low-redshift composite in this redshift range is z ∼ 2.
Blueward of Lyα emission, the strong IGM absorption at z ∼ 6 almost completely removes the
quasar flux. Redward of Lyα emission, there is no detectable difference in the intrinsic UV spectral
properties. The continuum shape is consistent with the power law fν ∝ ν −0.4 measured by Vanden
Berk et al. (2001). Clearly, spectral coverage in the near-IR is needed to put stronger constraints
– 9 –
on the continuum shape. Pentericci et al. (2003) use IR photometry of a sample of quasars at
z = 3.5 − 6 to measure a continuum shape of fν ∝ ν −0.5 , independent of redshift. The strengths
of the emission lines at z ∼ 6, including NV 1240A, OI 1300A, CII 1335A, and SiIV+OIV 1400A,
are also comparable to those at low redshift. This composite does not go red enough to cover the
CIV 1549A line, so we cannot test whether the Baldwin (1977) effect exists at these redshift. The
strength of the red wing of the Lyα emission line shows no evolution from that at low redshift.
The weaker Lyα emission in the blue wing is due to the strong IGM absorption. The average
FWHM of emission lines is ∼ 6000 km s−1 , also consistent with the low-redshift average.
5. Luminosity Function of z ∼ 6 Quasars
In Papers I and II, we estimated the comoving density of quasars at z ∼ 6 using a sample
of six quasars, covering a total area of 2870 deg2 . We repeat the calculation here, including
the additional three quasars in the complete sample; as explained above, SDSS 0002+2550 and
SDSS J0005-0006 were not selected as part of the flux-limited complete sample and will not be
included in the quasar luminosity function calculation. These nine quasars form a complete sample
satisfying the selection criteria in Eq. (1) over a total area of 4578 deg2 . Following Papers I and
II, we calculate the selection function of z ∼ 6 quasars using a Monte-Carlo simulation of quasar
colors, taking into account the distribution of quasar emission line and continuum properties, Lyα
absorption, the SDSS photometric errors and Galactic extinction. The selection function as a
function of redshift z and absolute magnitude M1450 is illustrated in Figure 7 of Paper II for the
Λ-model. The total spatial density of quasars at z ∼ 6 is derived using the 1/Va method. We
find that at the average redshift of hz i = 6.07, ρ(M1450 < −26.4) = (10.5 ± 3.9) × 10−10 Mpc−3
for the Ω = 1 model, and ρ(M1450 < −26.7) = (6.4 ± 2.4) × 10−10 Mpc−3 for the Λ-model. The
results, which are consistent with those in Paper II with smaller error bars, are plotted in Figure 4,
together with the measurements at lower redshifts from the 2dF survey (Boyle et al. 2000, Croom
et al. 2004) at z < 2.2, from Schmidt et al. (1995) at 2.7 < z < 4.8 and from Fan et al. (2001a) at
3.6 < z < 5.0. The comoving density of luminous quasars at z ∼ 6 is 30 times smaller than that
at z ∼ 3.
Following Paper II, we derive the bright-end slope from the luminosity distributions of the
sample using a maximum likelihood estimate. Assuming a single-power law luminosity function:
Ψ(M1450 ) = Ψ∗10−0.4[M1450 +26](β+1) ,
(2)
we find that for the Λ-model, Ψ∗ = (3.3+3.8
−1.6 ) × 10−9Mpc−3 . The best-fit bright-end slope is
β = −3.2, with a 68% confidence range of [–2.5, –4.0] and a 95% confidence range of [–2.2, –4.2].
This slope is steeper than that measured in Paper II, where the best-fit value was –2.3 with a
68% range of [–1.6, –3.1]. The reason for this 1-σ change is that all three new quasars included in
the sample have zAB ∼ 20, close to the detection limit. Their inclusion, when correcting for the
lower completeness at the faint end, drives the best-fit luminosity function to steeper slopes. For
– 10 –
example, the average selection probability of quasars with the redshift and luminosity of SDSS
J1623+3112 is of the order 10%; the strong Lyα emission boosts the z -band magnitude and makes
the z − J color bluer, making the selection of this ob ject easier than an average z = 6.2 quasar.
The best-fit slope omitting SDSS J1623+3112 is –2.8. The slopes derived here and in Paper II
differ only at the 1-σ level. The errors are large because of the small number of ob jects in the
sample, especially at the faint end. This underlines the need for a large sample to put strong
constraints on the quasar luminosity function at the highest redshift.
The slope of the quasar luminosity function has important implications. The total quantity
of ionizing photons emitted by the high-redshift quasar population is determined by the quasar
luminosity function. Fan et al. (2001) found that z ∼ 6 quasars are not likely to be the sources
responsible for reionizing the universe, or keeping the universe ionized at high redshift, assuming
a slope of β > −3.5. Interestingly, we are approaching this limit with the current analysis.
In a flux-limited sample, the lensing probability of the brightest observed ob jects is boosted
by magnification bias (e.g. Turner et al. 1984). The theoretical prediction of the fraction
of strongly-lensed quasars at high redshift could be be of order unity for a sufficiently steep
luminosity function (e.g. Wyithe & Loeb 2002a, Comerford, Haiman & Schaye 2002). Fan et al.
(2003) and Richards et al. (2004) used the lack of multiply-imaged quasars among the SDSS z ∼ 6
quasar sample to constrain the shape of the quasar luminosity function to be β > −4.6 at the 3-σ
level. The lensing fraction increases by a factor of ∼ 5 by assuming a slope of β = −3.3 rather
than β = −2.2 (Wyithe & Loeb 2002b). The quasar emissivity and lensing fraction also depends
strongly on the faint-end quasar slope, which is currently completely unknown.
In this paper, we present the first quasar discovered in the faint quasar survey using co-added
catalogs from the SDSS Southern Survey region. In Fall 2003, we also started to use co-added
images of the SDSS Southern Survey from ∼ 8 SDSS observations to select candidates. This
overcomes the limitation of the co-added catalog which requires the ob ject to be detected in
a single exposure. We have generated a preliminary photometric catalog using the SExtractor
software; the integration of the co-added imaging into the SDSS photometric pipeline is currently
underway. Using the catalogs generated by co-added imaging, we were able to recover SDSS
J000552.34–000655.8. The remaining candidate identification is still in progress and will be
reported in the future. In the next few years, the faint quasars selected from the Southern Survey
will be combined with the bright quasars in the Main Survey region to study the evolution of
quasar population at z ∼ 6 down to much fainter luminosities to probe the evolution of faint
quasars.
Funding for the creation and distribution of the SDSS Archive has been provided by the
Alfred P. Sloan Foundation, the Participating Institutions, the National Aeronautics and Space
Administration, the National Science Foundation, the U.S. Department of Energy, the Japanese
Monbukagakusho, and the Max Planck Society. The SDSS Web site is http://www.sdss.org/.
The SDSS is managed by the Astrophysical Research Consortium (ARC) for the Participating
– 11 –
Institutions. The Participating Institutions are The University of Chicago, Fermilab, the Institute
for Advanced Study, the Japan Participation Group, The Johns Hopkins University, Los Alamos
National Laboratory, the Max-Planck-Institute for Astronomy (MPIA), the Max-Planck-Institute
for Astrophysics (MPA), New Mexico State University, University of Pittsburgh, Princeton
University, the United States Naval Observatory, and the University of Washington. We thank the
staffs at Apache Point Observatory, the MMT, the Bok Telescope, Kitt Peak, the Hobby-Eberly
Telescope, and Magellan for their expert help. We acknowledge support from NSF grant AST
03-07384, a Sloan Research Fellowship and the University of Arizona (X.F.), NSF grants AST
00-71091 and AST 03-07409 (M.A.S.) and NSF grants AST 99-00703 and AST 03-07582 (D. P.
S.).
REFERENCES
Abaza jian, K., et al. 2003, AJ, 126, 2081
Abaza jian, K., et al. 2004, AJ, in press (astro-ph/0403325)
Anderson, S. F., et al. 2001, AJ, 122, 503
Baldwin, J. A. 1977, ApJ, 214, 679
Barth, A. J., Martini, P., Nelson, C. H., & Ho, L. C. 2003, ApJ, 594, L95
Becker, R. H., White, R. L., & Helfand, D. J. 1995, ApJ, 450, 559
Becker, R. H. et al. 2001, AJ, 122, 2850
Boyle, B.J., Shanks, T., Croom, S. M., Smith, R . J., Miller, L., Loaring, B., & Heymans, C. 2000,
MNRAS, 317, 1014
Brandt, W.N., et al. 2002, ApJ, 569, L5
Comerford, J., Haiman, Z., & Schaye, J. 2002, ApJ, 580, 63
Condon, J. J., Cotton, W. D., Greisen, E. W., Yin, Q. F., Perley, R. A., Taylor, G. B., &
Broderick, J. J., 1998, AJ, 115, 1693
Croom, S. M., Smith, R. J., Boyle, B. J., Shanks, T., Miller, L., Outram, P. J., & Loaring, N. S.,
2004, MNRAS, in press (astro-ph/0403040)
Dietrich, M., Appenzeller, I., Hamann, F., Heidt, J. Jaeger, J., Vestergaard, M., & Wagner, S. J.,
2003a, A&A, 398, 891
Dietrich, M., Hamann, F. Shields, J. C., Constantin, A., Heidt, J., Jaeger, M., Vestergaard, M., &
Wagner, S. J., 2003b, ApJ, 722, 732
Djorgovski, S. G., Castro, S., Stern, D., & Mahabal, A. A. 2001, ApJ, 560, L5
– 12 –
Fan, X. et al., 2000, AJ, 120, 1167
——, 2001a, AJ, 121, 31
——, 2001b, AJ, 121, 54
——, 2001c, AJ, 122, 2833
——, 2002, AJ, 123, 1247
——, 2003, AJ, 125, 1649
Finkbeiner, D. et al. 2004, AJ, submitted
Freudling, W., Corbin, M. R., & Korista, K. T. 2003, ApJ, 587, L67
Fukugita, M., Ichikawa, T., Gunn, J.E., Doi, M., Shimasaku, K., & Schneider, D.P. 1996, AJ, 111,
1748
Gunn, J. E., & Peterson, B. A. 1965, ApJ, 142, 1633
Gunn, J.E., et al. 1998, AJ, 116, 3040
Hamann, F., & Ferland, G., 1993, ApJ, 418, 11
Hill, G. J., et al. 1998, Proc. SPIE, 3355, 433
Hogg, D., et al. 2001, AJ, 122, 2129
Kogut, A. et al. 2003, ApJS, 148, 161
Lupton, R.H., Gunn, J.E., & Szalay, A. 1999, AJ, 118, 1406
Lupton, R. H., Gunn, J. E., Ivezic, Z., Knapp, G. R., Kent, S. M., & Yasuda, N., 2001, ASP
Conf. Ser. 238: Astronomical Data Analysis Software and Systems X, 10, 269
Maiolino, R., Oliva, E., Ghinassi, F., Pedani, M. Mannucci, F., Mujica, R., & Juarez, Y., 2004,
A&A, submitted (astro-ph/0312402)
Pentericci, L., et al. 2002, AJ, 123, 2151
Pentericci, L., et al. 2003, A&A, 409, 47
Pier, J. et al. 2003, AJ, 125, 1559
Richards, G. T., et al. 2002, AJ, 123, 2925
Richards, G. T., et al. 2004, AJ, 127, 1305
– 13 –
Schlegel, D.J, Finkbeiner, D.P., & Davis, M. 1998, ApJ, 500, 525
Spergel, D. N. et al. 2003, ApJS, 148, 175
Schmidt, M., Schneider, D. P., & Gunn, J.E., 1995, AJ, 110, 68 (SSG)
Schneider, D. P., Schmidt, M., & Gunn, J.E. 1991, AJ, 102, 837
Skrutskie, M. F., et al., 1997, The Impact of Large-Scale Near-IR Sky Surveys, ed. F. Garz´on, N.
Epchtein, A. Omont, B,. Burton, & P. Persei (Dordrecht: Kluwer), 25
Smith, J., et al. 2002, AJ, 123, 2121
Songaila, A., 2004, AJ, in press (astro-ph/0402347)
Stoughton, C. et al. 2002, AJ, 123, 485
Vignali, C., Brandt, W.N., Schneider, D.P., Garmire, G.P., & Kaspi, S. 2003, ApJ, 125, 2876
Turner, E. L., Ostriker, J. P., & Gott, R. 1984, ApJ, 284, 1
Vanden Berk, D. E., et al. 2000, AJ, 122, 549
White, R. L., Becker, R. H., Fan, X., & Strauss, M. A., 2003, AJ, 126, 1
Wyithe, J. S. B., & Loeb, A., 2002a, Nature, 417, 923
Wyithe, J. S. B., & Loeb, A., ApJ, 577, 57
York, D. G., et al. 2000, AJ, 120, 1579
This preprint was prepared with the AAS LATEX macros v4.0.
– 14 –
Table 1. Photometric Properties of Five New z > 5.7 Quasars
ob ject
redshift
i
z
J
SDSS run
J000239.39+255034.8
J000552.34−000655.8
J141111.29+121737.4
5.80 ± 0.02
5.85 ± 0.02
5.93 ± 0.02
J160254.18+422822.9
J162331.81+311200.5
6.07 ± 0.02
6.22 ± 0.02
21.47 ± 0.11
23.40 ± 0.34
23.43 ± 0.37
22.85 ± 0.30
22.78 ± 0.38
24.52 ± 0.62
18.99 ± 0.05
20.54 ± 0.10
19.63 ± 0.07
19.65 ± 0.08
19.89 ± 0.10
20.09 ± 0.10
> 16.5
19.87 ± 0.10
18.95 ± 0.05
18.46 ± 0.05
19.15 ± 0.10
4152
multiple
3836
3996
3705
3918
The SDSS photometry (i, z ) is reported in terms of asinh magnitudes on the AB system. The
asinh magnitude system is defined by Lupton, Gunn & Szalay (1999); it becomes a linear scale in
flux when the absolute value of the signal-to-noise ratio is less than about 5. In this system, zero
flux corresponds to 24.4 and 22.8, in i, and z , respectively; larger magnitudes refer to negative flux
values. The J magnitude is on a Vega-based system.
Table 2. Continuum Properties of new z > 6 Quasars
ob ject
redshift
AB1280
AB1450
M1450
(Ω-model)
M1450
(Λ-model)
E (B − V )
(Galactic)
J000239.39+255034.8
J000552.34−000655.8
J141111.29+121737.4
J160254.18+422822.9
J162331.81+311200.5
5.80
5.83
5.93
6.07
6.22
19.09
20.30
20.04
19.93
20.20
19.02
20.23
19.97
19.86
20.13
–27.40
–26.21
–26.49
–26.63
–26.40
–27.66
–26.46
–26.75
–26.82
–26.67
0.037
0.033
0.025
0.014
0.022
– 15 –
Fig. 1.— SDSS z -band images of the five new z > 5.7 quasars. Each side of the finding chart is
160′′ . The arrow at the lower left indicates the direction of North on the finding chart; East is 90◦
counterclockwise from North.
– 16 –
Fig. 2.— Spectra of the five new quasars at z > 5.7. The spectrum of SDSS J0002+2550 is a
60-min exposure taken with the MMT-Red Channel; the spectrum of SDSS J0005-0006 is a 40-min
exposure taken with Magellan II and B&C spectrograph; the spectrum of SDSS J1411+1217 is
a 120-min exposure taken with KPNO-4m and MARS; the spectrum of SDSS J1602+4228 is a
120-min exposure taken with the MMT-Red Channel; and the spectrum of SDSS J1623+3112 is a
co-addition of two 120-min exposures using MARS and the Red Channel. The fluxes are scaled to
reproduce the z -band magnitude as measured by the SDSS. All spectra are binned to a dispersion
of 5A per pixel; the spectral resolutions are between 500 and 1000, depending on the spectrograph
used.
– 17 –
Fig. 3.— The composite spectrum of eleven z ∼ 6 quasars (solid line). The spectrum of SDSS
J0005-0002 was not included because of its low S/N. The spectrum of each quasar is redshifted,
scaled according to its m1450 magnitude, and averaged with equal weighting. For comparison, we
also plot the low-redshift quasar spectral composite from Vanden Berk et al. (2001). The effective
redshift in the 1000 – 1500A range in the Vanden Berk et al. composite is about 2. The quasar
intrinsic spectrum redward of Lyα emission shows no detectable evolution up to z ∼ 6, in terms
of both the continuum shape and emission line strengths. On the blue side of Lyα emission, the
strong IGM absorption at z ∼ 6 removes most of the quasar flux.
– 18 –
Fig. 4.— The evolution of the quasar comoving spatial density at M1450 < −26.7 in the Λ-model.
The filled circle represents the result from this survey. The error-bar in redshift indicates the
redshift range covered by the i-dropout survey. The dashed and dotted lines are the best-fit models
from Fan et al. (2001b) and Schmidt et al. (1995, SSG), respectively. The solid line is the best-fit
model from the 2dF survey (Croom et al. 2004) at z < 2.3.
|
0707.0468 | 1 | 0707 | 2007-07-03T18:10:34 | A Spitzer Study of the Mass Loss Histories of Three Bipolar Pre-Planetary Nebulae | [
"astro-ph"
] | We present the results of far-infrared imaging of extended regions around three bipolar pre-planetary nebulae, AFGL 2688, OH 231.8+4.2, and IRAS 16342$-$3814, at 70 and 160 $\mu$m with the MIPS instrument on the Spitzer Space Telescope. After a careful subtraction of the point spread function of the central star from these images, we place constraints on the existence of extended shells and thus on the mass outflow rates as a function of radial distance from these stars. We find no apparent extended emission in AFGL 2688 and OH 231.8+4.2 beyond 100 arcseconds from the central source. In the case of AFGL 2688, this result is inconsistent with a previous report of two extended dust shells made on the basis of ISO observations. We derive an upper limit of $2.1\times10^{-7}$ M$_\odot$ yr$^{-1}$ and $1.0\times10^{-7}$ M$_\odot$ yr$^{-1}$ for the dust mass loss rate of AFGL 2688 and OH 231.8, respectively, at 200 arcseconds from each source. In contrast to these two sources, IRAS 16342$-$3814 does show extended emission at both wavelengths, which can be interpreted as a very large dust shell with a radius of $\sim$ 400 arcseconds and a thickness of $\sim$ 100 arcseconds, corresponding to 4 pc and 1 pc, respectively, at a distance of 2 kpc. However, this enhanced emission may also be galactic cirrus; better azimuthal coverage is necessary for confirmation of a shell. If the extended emission is a shell, it can be modeled as enhanced mass outflow at a dust mass outflow rate of $1.5\times10^{-6}$ M$_\odot$ yr$^{-1}$ superimposed on a steady outflow with a dust mass outflow rate of $1.5\times10^{-7}$ M$_\odot$ yr$^{-1}$. It is likely that this shell has swept up a substantial mass of interstellar gas during its expansion, so these estimates are upper limits to the stellar mass loss rate. | astro-ph | astro-ph |
A Spitzer Study of the Mass Loss Histories of Three Bipolar
Pre-Planetary Nebulae
Tuan Do, Mark Morris
Physics and Astronomy Department, University of California, Los Angeles, CA 90095-1547
Raghvendra Sahai and Karl Stapelfeldt
Jet Propulsion Laboratory, California Institute of Technology, Pasadena, CA 91109
ABSTRACT
We present the results of far-infrared imaging of extended regions around
three bipolar pre-planetary nebulae, AFGL 2688, OH 231.8+4.2, and IRAS
16342−3814, at 70 and 160 µm with the MIPS instrument on the Spitzer Space
Telescope. After a careful subtraction of the point spread function of the central
star from these images, we place constraints on the existence of extended shells
and thus on the mass outflow rates as a function of radial distance from these
stars. We find no apparent extended emission in AFGL 2688 and OH 231.8+4.2
beyond 100 arcseconds from the central source. In the case of AFGL 2688, this
result is inconsistent with a previous report of two extended dust shells made on
the basis of ISO observations. We derive an upper limit of 2.1 × 10−7 M⊙ yr−1
and 1.0 × 10−7 M⊙ yr−1 for the dust mass loss rate of AFGL 2688 and OH 231.8,
respectively, at 200 arcseconds from each source. In contrast to these two sources,
IRAS 16342−3814 does show extended emission at both wavelengths, which can
be interpreted as a very large dust shell with a radius of ∼ 400 arcseconds and
a thickness of ∼ 100 arcseconds, corresponding to 4 pc and 1 pc, respectively,
at a distance of 2 kpc. However, this enhanced emission may also be galactic
cirrus; better azimuthal coverage is necessary for confirmation of a shell. If the
extended emission is a shell, it can be modeled, with some assumptions about
its dust properties, as enhanced mass outflow at a dust mass outflow rate of
1.5 × 10−6 M⊙ yr−1 superimposed on a steady outflow with a dust mass outflow
rate of 1.5 × 10−7 M⊙ yr−1. Because of the size of the possible shell, it is likely
that this shell has swept up a substantial mass of interstellar gas during its ex-
pansion, so these estimates are upper limits to the stellar mass loss rate. We find
a constant color temperature of 32 K throughout the circumstellar envelope of
IRAS 16342−3814, which is consistent with heating by the interstellar radiation
field.
– 2 –
Subject headings: (ISM:) planetary nebulae: individual (AFGL 2688, OH 231.8+4.2,
IRAS 16342−3814) - stars: AGB and post-AGB - stars: mass loss
1.
Introduction
Pre-planetary nebulae (PPNe) represent a fleeting stage of stellar evolution. These
transitional objects arise between the rapid mass loss phase at the end of the asymptotic
giant branch (AGB) and the ionized planetary nebula stage. AGB stars are important
because they are the Galaxy's main mechanism for the replenishment of dust and gas into
the interstellar medium (ISM), ejecting most of their main sequence mass within a few times
105 years because of their high mass loss rates. By looking at the circumstellar material
around PPNe, we can see an imprint of the activity of the earlier AGB. PPNe are also in
a stage where the geometry of the mass loss from the central star usually changes from
spherically symmetric to an axially symmetric bipolar outflow, as illustrated in the HST
images of AFGL 2688 (Sahai et al. 1998). In these images, a spectacular bipolar structure is
superimposed on a set of concentric shells of gas and dust ejected near the end of the AGB.
In the study reported here, we have attempted to characterize the mass loss behavior of the
progenitor AGB stages of three systems that are now bipolar.
Using the direct detection of infrared emission from dust lost over as much as 105 years
during the AGB, it is possible to study the mass loss histories of these objects. There has
been a dearth of observations of large extended shells of evolved stars out to the distances
sampled in this project primarily because of the difficulty in observing in the far infrared
where the cool dust has its peak emission. With the Spitzer Space Telescope (Werner et al.
2004), the far infrared is now accessible at a resolution and sensitivity sufficient to potentially
resolve the structures of these dust shells, and thereby to test theories of mass loss in AGB
stars.
Several tantalizing cases were observed by IRAS, showing that large dust shells exist
around some evolved stars. For example, Gillett et al. (1986) found that R CrB shows a
very large shell with a radius of 4 parsecs, and Hawkins (1990) reported a dust shell with
a diameter of 30 - 40 arcminutes (∼ 1 pc) around W Hya. However, these observations
could not resolve the structure that may be present in the shells, or address the possibility
of multiple shells. If the mass loss during the AGB is constant, we should expect to see a
smooth envelope with column density declining as 1/b, where b is the displacement from the
star in the sky. However, if the mass loss rate fluctuates as the star goes through thermal
pulses during the AGB, as the models suggest (e.g., Vassiliadis & Wood 1993), then we
should see enhanced emission in the form of multiple dust shells around our objects. For
– 3 –
example, Speck et al. (2000) reported on the basis of ISO data that there are large concentric
shells in AFGL 2688 and AFGL 618. This result is among a small set of observations
showing evidence for periodic mass loss caused by thermal pulsation as well as the time scale
between thermal pulses predicted by AGB evolutionary models (e.g., Vassiliadis & Wood
1993). Confirmation of these shells at higher resolution and greater sensitivities motivated
our selection of AFGL 2688 as a target for this study. There have been no previous claims
of large-scale dust emission from OH 231.8+4.2 (hereafter OH 231.8) or IRAS 16342−3814.
However, they are presently very young PPNe with high mass loss rates of approximately
10−4M⊙ yr−1 (Alcolea et al. 2001; Sahai et al. 1999), which suggests that the emission from
dust produced during the AGB may be readily visible.
Here we report on the lack of very extended emission, spherically symmetric or otherwise,
from AFGL 2688 and OH 231.8. However, we do see possible evidence for a very large diffuse
dust shell around IRAS 16342−3814 with a radius of 400 arcseconds, though this extended
emission might also be from galactic cirrus.
2. Observations and Data Reduction
AFGL 2688 (the Egg Nebula), OH 231.8, and IRAS 16342−3814 were observed with the
MIPS instrument (Rieke et al. 2004) at 70 and 160 µm on the Spitzer Space Telescope. These
PPNe were observed along two mutually perpendicular scan paths in order to sample dust
emission out to about 800 arcseconds from the central source, and to determine background
levels. At 70 µm, the scan paths are 15′ ×3.0′ in one direction and 11′ ×7.7′ in the other, while
at 160 µm, the scan paths are 15′ × 2.6′ and 10.5′ × 6.5′. The pixel scale is 9.2 arcseconds
at 70 µm and 16 arcseconds at 160 µm. However, the central sources were not directly
observed in order to avoid persistence artifacts from these bright sources. This observing
strategy limits the azimuthal coverage of the sources, but should give an adequate estimate
of the presense of well-defined shells or asymmetries such as large-scale bipolarity.
The basic science calibrated data (BCD) was reduced using the software package MOPEX1
from the Spitzer Science Center (SSC). MOPEX was used for rejecting outliers as well as
for co-adding and mosaicking the individual BCD frames to create the final image. Further
reduction was done manually to remove striping due to bright latents as described in the
MIPS data handbook2. The correction involved finding the median value measured for every
1http://ssc.spitzer.caltech.edu/postbcd/
2http://ssc.spitzer.caltech.edu/mips/dh/
– 4 –
pixel from a series of BCD frames far from the central source, where we assume there to be
no extended emission, then subtracting this median from each data frame before co-adding
and mosaicking. The median subtraction removes both the bright latents and the uniform
component of background emission. Figure 1 shows the final mosaicked images.
The most prominent features of the 70 µm images of the bipolar targets are the diffrac-
tion spikes from the point spread function (PSF) of the central source. Although we did
not image the central star, the wings of the PSF are still present and bright out to about
200 arcseconds from the source. At 160 µm, the PSF is less pronounced but we see much
more background than at 70 µm from galactic cirrus. The galactic cirrus emission presents
a problem in determining the true sky background since it covers much of our fields. For
both the 70 and 160 µm images, we use the median surface brightness of regions of low
and uniform brightness far from the central source as an estimate of the sky background.
This procedure removes a uniform background from the images, retaining possible enhanced
emission from the source and galactic cirrus. The uncertainty in the background (see below)
at 160 µm is higher than at 70 µm because it is more difficult to find a large patch of uniform
surface brightness to estimate the sky background.
2.1. Sensitivities
The sensitivity of our data is estimated by examining the standard deviation of a patch
of uniform background far from the source both before and after the median background
subtraction method described above. For AFGL 2688, before background subtraction, the
mean of a patch in the eastern scan path at 70 µm is 14.1 MJy Sr−1 with a standard deviation
of 1.11 MJy Sr−1. Some of the variance in the background is likely from the effects of bright
latents from the detector pixels, which are removed by the median background subtraction
described above. After median subtraction, the mean background is at 0.6 MJy Sr−1 with
a standard deviation of 0.92 MJy Sr−1. The mean surface brightness at 160 µm before
background subtraction is 31.45 MJy Sr−1 with a standard deviation of 1.67 MJy Sr−1.
After background subtraction, the mean is at 0.33 MJy Sr−1 with a standard deviation of
1.51 MJy Sr−1. The sensitivity for OH 231.8 is better than for AFGL 2688, probably because
of the lower sky background levels. The sensitivity for IRAS 16342−3814 is comparable to
that of AFGL 2688 at 70 µm, but worse at 160 µm due to a higher sky background and
greater presence of galactic cirrus. The mean background and sensitivities for all three
sources at 70 and 160 µm are given in Tables 1 and 2, respectively.
– 5 –
2.2. PSF Subtraction
Since the PSF pattern from the central source is so prominent at 70 µm, we attempted
to fit a model PSF to our images in order to subtract the bright central source in search
of extended dust emission near the source. The stinytim software provided by the SSC was
used to generate model PSFs. We chose to use the default MIPS throughput curve along
with a blackbody spectrum at several temperatures ranging from 20 to 100 K. See Figure
2 for examples of the model PSFs. Multidimensional fitting was carried out for our images
in order to minimize the residuals of the PSF subtraction using various x-y shifts as well as
the scaling of the PSF flux values. The best fit was found by minimizing the square root of
the sum of the squares of the differences in flux. We found several limitations in the model
PSFs when attempting to subtract the PSF from the images. Even for the best-fit PSFs, the
subtraction still leaves some obvious PSF residuals at a level of 10−4 relative to the peak as
extrapolated from the model PSF. These residuals show that the wings of the model PSF are
not well characterized at that flux level. The PSF subtraction also tends to leave portions
of the region within the Airy ring with negative flux values. The most likely reason is a
non-linear pixel response as the pixels approach saturation so that although the model has a
higher brightness value closer to the core of the PSF, the actual pixel values are leveling off
for our bright sources. Unfortunately, there are presently no empirical data on the MIPS 70
or 160 µm PSF to compare to the model PSF at distances greater than 100 arseconds from
the peak. Beyond about 240 arseconds, PSF features are no longer present in the MIPS
images so our ability to detect extended emission there is only limited by background and
integration time.
The results of PSF subtraction for AFGL 2688 are shown in Figure 3. Note the remaining
features near the north and at a position angle of ∼ 50◦ east of north, about 150 arcseconds
from location of the source in the PSF subtraction in Figure 3a. This residual can also been
seen in the roughly the same region of the 70 µm PSF subtraction of OH 231.8 in Figure 4.
It seems unlikely that these features are physical, but the northern feature in AFGL 2688
does align with its bipolar outflows seen in the optical and infrared, but which are observed
within the inner 10 arcseconds of the PPN. The blob in AFGL 2688 at a position angle
of ∼ 50◦ also roughly aligns with the outflow direction seen in previous CO observations
(Skinner et al. 1997), also at around 10 arcseconds from the central star. The fact that the
features we see are at nearly 200 arcseconds from the source and lie almost on top of the
diffraction rays of the original PSF casts doubt upon their reality. This dilemma cannot
be resolved until a well sampled empirical PSF out to several arcminutes is available. We
may be seeing possible features in the wings of the actual PSF that the model has failed to
reproduce.
– 6 –
As a substitute for an empirical PSF, we used OH 231.8 to subtract the PSF from AFGL
2688. Figure 3b shows the result. The residuals from using OH 231.8 as an empirical PSF
appear to be much smaller than those obtained using the model PSF, reducing the residual
of 30 MJy Sr−1 to less than 10 MJy Sr−1 at 100 arcseconds. Note that there are almost
no PSF features, such as diffraction spikes, remaining in Figure 3b compared to Figure 3a.
However, it may be problematic to use OH 231.8 +4.2 as an empirical PSF because it may
have extended emission as well, although it would be remarkably fortuitous if both sources
had extended emission having precisely the same morphology. It is also problematic to scale
OH 231.8 to the same surface brightness as AFGL 2688 because in order to subtract the PSF,
the whole image needs to be scaled, thus scaling the background as well. Because AFGL
2688 is over twice as bright as OH 231.8, but the background in AFGL 2688 is not, scaling to
the level of the AFGL 2688 PSF would overemphasize the background in OH 231.8 so that
the residuals after the PSF subtraction might be dominated by the background in OH 231.8.
We can nevertheless see that the PSF subtraction using OH 231.8 results in residuals at a
level ∼ 10 times below that from using the model PSF at 150 arcseconds, which suggests
that OH 231.8 is either a point source at 70 µm or that its extended structure has the same
orientation, scale, and shape as in AFGL 2688.
The two-dimensional PSF subtraction at 160 µm was done in a similar way using the
model stinytim PSF. There are PSF features remaining after this subtraction as well (see
Figure 3); the ends of the diffraction spikes at 200 arcseconds from the source in the original
image are not completely removed. We did not attempt a PSF subtraction of the 160 µm
image of IRAS 16342−3814, because that the PSF of the central source is weak enough at
this wavelength that its features are below the background and the extended structures.
2.3.
1-D PSF Subtraction
In order to investigate the hypothesis of spherical shells around a central source, we
have determined an azimuthal average of the intensity around the central source. Such an
average would show enhancements for spherically concentric shells projected as circularly
symmetric features from episodic mass loss during the AGB phase. An azimuthal average
also has the advantage that it greatly reduces the parameter space necessary to fit the model
PSF. The source position and intensity are found by fitting to the brightest Airy ring visible
in our images, which is the second brightest Airy ring of the PSF (see Figure 2), since the
source itself does not appear in our images. The PSF Airy ring used for scaling the PSF
is located 76 arcseconds from the source at 70 µm and 170 arcseconds from the source at
160 µm. We also find that the azimuthal averages provide a subtraction with less residuals
– 7 –
than using the two dimensional subtraction because localized irregularities of the PSF are
averaged away.
The one dimensional PSF subtractions for AFGL 2688 and IRAS 16342−3814 at 70
µm were done using both the model PSF and OH 231.8, and are shown in Figure 5. The
profiles of AFGL 2688 and OH 231.8 follow each other almost exactly out to beyond 200
arcseconds from the source. This further indicates that OH 231.8 and AFGL 2688 are both
strongly centrally concentrated at 70 µm, unless they both have exactly the same azimuthal
excesses. They both also match the model PSF very well out to 100 arcseconds beyond
which both appear to have slight excess (< 1% of the extrapolated peak of the PSF), above
the model PSF. Once again, this excess in the wings is more likely due to the model PSF
not reproducing the wings of the actual PSFs rather than resulting from physical emission
associated with the sources. At 160 µm, we only use the model PSF for subtraction because
of possible contamination due to galactic cirrus in OH 231.8.
3. Results and Analysis
3.1. AFGL 2688
3.1.1. 70 µm
At 70 µm, the PSF-subtracted images of both AFGL 2688 and OH 231.8 are similar. The
PSF residuals extend to about 150 arcseconds from the source, which may be obscuring some
physical extended features. However, there are no obvious circularly symmetric residuals
in the PSF subtraction even with the confusion near the central source, which would be
expected if there had been spherically symmetric mass loss above a dust mass loss rate of
∼ 2.1 × 10−7M⊙ yr−1 (see discussion section). Previous studies with ISO by Speck et al.
(2000) led to the claim of large shells around AFGL 2688 having radii of 150 and 300
arcseconds on the basis of a one-dimensional scan of the Egg Nebula with ISOPHOT at
120 and 180 µm. We find, with better spatial resolution and azimuthal coverage using
Spizer, that at 70 µm, there are no signs of shell-like extended emission in our field. At 150
arcseconds from the source, the PSF subtraction residuals have a surface brightness of about
16 MJy Sr−1 using the model PSF and about 1 MJy Sr−1 using OH 231.8 as the empirical
point source(Figure 5). Note also that there is no discernible excess emission above the
fluctuations in the background (σ ∼ 0.92 MJy Sr−1) at 300 arcseconds from the source.
– 8 –
3.1.2. 160 µm
The 160 µm image (see Figure 3) shows roughly uniform enhanced emission to the edge
of our field of view at 1000 arcseconds from the source in the eastern scan path; the northern
scan show enhanced emission out to about 300 arcseconds, then dropping off to background
levels. The enhanced emission in the two paths is asymmetric as it does not fall off with
distance in the eastern scan path. There does not appear to be any structure to the emission
in either scan, other than that which may be attributable to the PSF diffraction spike in
the north. Because the residual extended emission shows no symmetry about the central
star, and because it shows no intensity falloff in the eastern scan path, this emission most
likely arises from irregularly distributed galactic cirrus in this direction; this is supported by
the relatively lower quality IRAS 100 µm image, which shows substantial extended cirrus
beyound our survey region. The ISOPHOT observation of a possible shell at 300 arcseconds
was based on a 53′ × 3′ linear scan with 30′′ × 92′′ pixels, at a position angle of 8 degrees east
of north centered on the source, which in our case would be sampling the emission in the
northern scan. We see with better azimuthal coverage with MIPS that the emission in both of
our scans is likely strongly contaminated with galactic cirrus. The presence of the inner shell
reported by Speck et al. at 150 arcseconds cannot be directly confirmed by our observations
because the 160 µm PSF Airy ring at 170 arcseconds from the source overlaps that region.
The surface brightness at the Airy ring is about 17 MJy Sr−1 after background subtraction
(but before the PSF subtraction). This is comparable to the background subtracted surface
brightness of 20 (30) MJy Sr−1MJy Sr−1 at 120 (180) µm reported by Speck et al. (2000).
In the azimuthally averaged surface brightness of our 160 µm image (Figure 6), we see
a similar result as the 70 µm data. Speck et al. reported a brightness value of the putative
shell at 300 arcseconds from the source to be 20 MJy Sr−1 at 120 µm and 60 MJy Sr−1
at 180 µm, without background subtraction. The 120 µm emission is about 10 MJy Sr−1
above their background as extrapolated from the surface brightness far from the star in their
plot. At 180 µm, the excess emission is about 15 MJy Sr−1 above the background. In the
MIPS 160 µm image, we find that the surface brightness at 300 arcseconds from the source
is about 2 MJy Sr−1 after background subtraction. Although the comparison is not at the
same wavelength, the excess emission we detect is 7 times below that measured by ISO.
When comparing the two scans in our observations separately, we see that the eastern scan
has roughly uniform brightness of 4 MJy Sr−1 out to 1000 arcseconds, while the emission in
the northern scan drops off at 400 arcseconds from the source from about 2 MJy Sr−1 to the
level we use as the background in the data reduction. Both the asymmetry in the enhanced
emission in both scans and the low surface brightness in the Spitzer data at 300 arcseconds
from the source lead us to conclude that the shell at this distance reported on the basis of
ISO data is probably not associated with the source.
– 9 –
3.2. OH 231.8+4.2
Neither the 70 nor 160 µm images of OH 231.8 show any signs of spherically symmetric
extended emission. The PSF subtraction at 70 µm leaves similar residuals as for AFGL
2688, though the brightnesses of these residuals are much lower because the central source is
intrinsically less bright. The strongest PSF residuals are along the diffraction spikes, similar
in shape to those from AFGL 2688. This further suggests that these residuals are not physical
features, but are rather PSF artifacts not accounted for by the stinytim model. In the 160
µm azimuthally averaged plots, both scans appear to have rather uniform emission, with the
northern scan having slightly higher surface brightness at approximately 2 MJy Sr−1, while
the eastern scan has an average surface brightness of 1 MJy Sr−1 (Figure 7).
The 160 µm image shows some slightly clumpy emission along the northern scan path,
slightly beyond the PSF diffraction spike (see Figure 1) at 320 arcseconds. The clumps, with
a width of about 80 arcseconds, have a surface brightness of about 3.6 MJy Sr−1 on a 2
MJy Sr−1 diffuse background. The IRAS 100 µm image shows diffuse galactic cirrus in this
region, which may be the origin of these clumps. They are aligned with the bipolar axis, but
unfortunately also with the diffraction ray so it may also be a PSF feature. Since this study
is concerned with investigating spherically symmetric emission, we will defer the question
of their nature. The azimuthally averaged surface brightness shows no enhanced emission
attributable to the source (Figure 8).
3.3.
IRAS 16342−3814
In contrast to the previous two sources, IRAS 16342−3814 does show some evidence for
what may be a large patchy shell at both 70 and 160 µm. The enhanced emission out to
a radial distance of ∼ 400 arcseconds is consistent with rough circular symmetry about the
star in the coverage area available (Figure 1). This radius corresponds to a physical size of
about 4 parsecs, assuming the distance to IRAS 16342−3814 to be 2 kpc (Sahai et al. 1999).
This possible shell structure is rather patchy, with a large arc in the western scan at 70 µm.
At 160 µm, the same arc is present, but is more diffuse. Nevertheless, the azimuthal average
shows an excess above the background at ∼ 400 arcseconds from the central source. The
brightest portion of this feature occurs at a position angle of 250 degrees east of north, which
is the same direction as the axis of the bipolar nebula seen in the optical. The ratio of the
70 to 160 µm average surface brightness, is roughly constant (∼ 0.3) as a function of radial
distance from the source. This ratio yields a color temperature of ∼ 32 ± 2 K, and, if a ν 1
emissivity law is assumed, a dust temperature of 26 ± 2 K. The constant color temperature
suggests that the dust is heated primarily by the interstellar radiation field.
– 10 –
Because there are a number of interstellar cirrus features in the general direction of
IRAS16342-3814, we cannot definitively argue that the patchy arc of emission in the scan
paths around this source is caused by mass loss from the star. The IRAS 100 µm image of
this region with an overlay of the coverage of our data (figure 9) shows that there is indeed
a substantial likelihood that this emission is cirrus. With this caveat in mind, we proceed in
the analysis below on the assumption that the extended emission has resulted from mass loss
from the star to obtain some indicative numbers for the mass loss rate necessary to produce
such a shell. In the following analysis, we will use the term "mass outflow rate" to indicate
the total mass flowing outward per unit time across a spherical shell at a particular radius;
this includes both the stellar component and the interstellar medium being swept up; we
will use the term mass loss rate to refer to mass loss from only the star.
Making a few assumptions about the mass outflow and the dust, we can set some limits
on the mass outflow rate of IRAS 16342−3814. Assuming that the mass loss is spherically
symmetric, we can fit a model of the expected brightness profile to the observed radial profile
of the source. Figures 10 and 11 show the azimuthally averaged radial profiles at 70 and 160
µm, respectively. We fit only the 70 µm radial profile because there is less contamination by
galactic cirrus compared to 160 µm. Following the derivation by Gillett et al. (1986), if we
assume a 1/r2 density profile, appropriate for constant mass outflow, where r is the distance
from the star, and a constant expansion velocity, ve, then the surface brightness Iν(b) has
the form:
.
M κνBν(T )
Iν(b) =
cos−1(
b
Rmax
)
for Rmin ≤ b ≤ Rmax
Iν(b) =
.
M κνBν(T )
2πveb
2πveb
(cid:20)cos−1(
b
Rmax
) − cos−1(
b
Rmin
)(cid:21)
for b < Rmin
(1)
where b is the projected physical distance from the source in the sky, κν is the opacity
coefficient in units of cm2 g−1, Bν(T ) is the blackbody function, and Rmax and Rmin are the
outer and inner radii of the shell. We further assume that the outflow velocity is constant
at 15 km s−1, and that the dust is composed of small astronomical silicate grains with κν
= 104 cm2 g−1 at 70 µm (using a = 0.1 µm, ρ = 2.3 g cm−3, and Qν
abs = 2.99 × 10−3 from
Draine & Lee (1984)). We also adopt a constant temperature of 26 K throughout the shell
and envelope. Figure 12 gives a comparison between different models having various mass
outflow rates. The best-fit model is a shell with a 4.2 pc radius and a thickness of 1 pc, with
a dust outflow rate of 1.5×10−6M⊙ yr−1 superimposed on a smooth envelope with a constant
dust outflow rate of 1.5 × 10−7M⊙ yr−1. This implies a rather high gas outflow rate on the
order of 3 ×10−4M⊙ yr−1 in the shell and 3 ×10−5M⊙ yr−1 in the smooth envelope, assuming
– 11 –
a gas to dust mass ratio of 200. Using OH 231.8 as the empirical PSF, the fit only requires a
shell component with a mass outflow rate of 3 × 10−4M⊙ yr−1 without an underlying smooth
envelope. These fits depend on the choice of PSF, but once the PSF is chosen, the mass
outflow rates are insensitive to residuals from PSF subtraction, since we only fit for the
region beyond 150 arcseconds from the central source (see Figure 10). However, our value
for the mass outflow rate is highly dependent on the assumed temperature of the dust. For
typical ISM dust temperatures between 22 and 35 K, the inferred dust outflow rate is related
.
M ∝ T −7.6 at 70 µm. If T > 26 K, we would
empirically to the adopted temperature by:
infer a substantially lower mass outflow rate.
Given the model thickness of the shell and assuming a typical expansion velocity of 15
km s−1, we estimate the maximum duration of the enhanced mass loss event which produced
the shell to be about 65,000 years. This is an upper limit because the internal velocity
dispersion in the shell broadens the shell as it expands. A 1 km s−1 dispersion, for example,
would reduce the duration of the mass loss that produced the shell by about a third. If we
use an age of 40,000 years for the duration of the enhanced mass loss and approximate the
shell as spherical, the total dust mass would be about 0.04 M⊙, giving a total shell mass of ∼
8 M⊙, assuming a gas-to-dust mass ratio equal to 200. The velocity of the shell is probably
slower than the molecular outflow velocity of a typical AGB star and the total amount of
mass lost by the AGB star is likely smaller than this amount because of the interaction with
the ISM out to the distances we are observing (see below).
The mass of the shell can be estimated directly from the infrared emission using the
equation M = FνD2/(Bν(T )κν), where Fν is the total flux density in the shell and D is the
distance to the source. The integrated flux density from the limited azimuthal coverage that
we have between radii of 300 and 400 arcseconds is 6.4 Jy at 70 µm, corresponding to a dust
mass of 0.015 M⊙ in our observed portion of the shell, assuming a temperature of 26 K for
the dust. We have about one fifth of the full azimuthal coverage with data at this distance
from the star, and can estimate the total dust mass of the shell if we approximate the shell
as isotropic with an averaged surface brightness of 2.5 MJy Sr−1 at 70 µm from the radial
profile, with a width of 100 arcseconds, and a temperature of 26 K. The total dust mass of
the extrapolated shell is then about 0.03 M⊙, comparable to the above estimate.
Because the shell is so large, the amount of interstellar matter that may have been swept
up and accumulated in the shell during the AGB could be significant. For an ISM hydrogen
density of 1 cm−3, a 4 parsec radius shell would have swept up about 7 M⊙, which would
account for almost all the mass that we may be measuring. The accumulation of interstellar
material could potentially lower the stellar mass loss rate estimated from our model by a
large factor. We therefore emphasize that the mass outflow rate estimate is only an upper
– 12 –
limit to the stellar mass loss rate.
One of the difficulties in addressing the mass loss history of IRAS 16342−3814 with the
current data is its location in the Galaxy; with a scale height of only 150 pc, there is possible
confusion with diffuse galactic cirrus. An unfortunate alignment with background emission
is an alternative to the existence of a shell produced by mass loss. Further observations with
better azimuthal coverage are needed to test the circumstellar shell hypothesis.
4. Discussion
The extended dust emission in AFGL 2688 and OH 231.8 seems to show that these two
objects do not have as long a mass loss history as one might have anticipated. Our results
for these two sources probe the region beyond ∼ 100 arcseconds from the central star, which
corresponds to 1.0 × 104 yrs ago for AFGL 2688 and 2.8 × 104 yrs ago for OH 231.8, given
a distance of 420 pc (Ueta et al. 2006) and 1.2 kpc (Jura & Morris 1985), respectively, and
an expansion velocity of 20 km s−1. We can can set a limit to the dust mass loss rate by
inverting equation 1:
M =
Iν(b, T )2πveb
κνBν(T ) cos−1(b/Rmax)
(2)
The dust mass loss rate can be simplified as a function of the observed surface brightness I
and the projected distance from the source in the sky b with the assumption that b ≪ Rmax:
M (70µm) ∼ 3.8 × 10−8M⊙ yr−1(cid:18) I70 µm
M (160µm) ∼ 2.9 × 10−8M⊙ yr−1(cid:18) I160 µm
1 MJy Sr−1(cid:19)(cid:18)100 cm2g−1
1 MJy Sr−1(cid:19)(cid:18)20 cm2g−1
κ70 µm (cid:19)(cid:18) b
κ160 µm (cid:19)(cid:18) b
1 pc(cid:19)(cid:16)
1 pc(cid:19)(cid:16)
15 km s−1(cid:17)(cid:18) T
15 km s−1(cid:17)(cid:18) T
30 K(cid:19)−7.6
30 K(cid:19)−3.6
ve
ve
If b ∼ Rmax, then the full form of equation 2 must be used because the cos−1(b/Rmax)
term in the denominator of equation 2 becomes important and will cause the dust mass
loss rate inferred from a given surface brightness value to increase drastically. For the cases
such as AFGL 2688 and OH 231.8 where we do not see a clear envelope associated with the
source, we cannot be sure that the condition b ≪ Rmax holds since Rmax is indeterminate.
The equations above also show the temperature dependence in the form of a power law to
approximate the blackbody function between 22 and 35 K. From these equations we see that
the temperature dependence at 160 µm is less steep than at 70 µm, but the contamination
by galactic cirrus near the sources makes estimating upper limits problematic at 160 µm.
We can establish an upper limit to the mass loss rates for AFGL 2688 and OH 231.8 at 70
µm based up on the surface brightness of the residual left from the PSF subtraction at 200
– 13 –
arcseconds from the source, which is a compromise between a location far enough from the
source that PSF subtraction errors are small and being close enough to the source that there
could plausibly be a circumstellar envelope. For AFGL 2688, 200 arcseconds corresponds to
a radial distance from the source of 0.4 parsecs, using a distance to the source of 420 parsecs
(Ueta et al. 2006). The surface brightness after PSF subtraction with the model PSF is
about 14 MJy Sr−1 at 70 µm. Assuming an expansion velocity of 20 km s−1 as observed in
CO by Skinner et al. (1997), a dust temperature of 30 K, κν = 104 cm2 g−1 appropriate for
carbon dust (Draine & Lee 1984), and that the possible envelope has a radius significantly
greater than 0.4 pc, then the dust mass loss rate upper limit is 2.1 × 10−7 M⊙ yr−1. Similarly
for OH 231.8, the residual emission from the PSF subtraction is 3 MJy Sr−1 at 200 arcseconds
from the central star, which corresponds to 1.3 parsecs at a distance of 1.3 kpc. This surface
brightness implies a dust mass loss rate of 1.0 × 10−7 M⊙ yr−1 assuming an outflow velocity
of 15 km s−1, κν = 98 cm2 g−1 (appropriate for silicate dust), and a temperature of 30 K.
There are several potential explanations for the lack of extended emission seen in this
study. One is that, even with Spitzer's increased sensitivity, the dust emission is below the
threshold for detection. Since the distances we are studying are far from the central star,
we would expect the temperature of the dust to be determined by the ambient interstellar
radiation field. If the dust temperature is about 20 K, typical for the ISM (Mathis et al.
1983), and an emissivity ∝ ν 1, the surface brightness would be 12 times stronger at 160
µm, than at 70 µm where the resolution and sensitivity of MIPS is better. We would be
more likely to detect cool dust emission at 160 µm than at 70 µm. However, confusion with
the galactic cirrus is also likely at 160 µm because it is about the same temperature as any
hypothetical extended dust emission associated with the source that is heated primarily by
the interstellar radiation field. For comparison, our detection of the possible shell from IRAS
16342-3184 has a ratio of I160µm/I70µm ∼ 3.3, corresponding to a color temperature of 32 K,
which requires a higher than average interstellar radiation field.
The upper limits to the mass loss rates derived in this study can be compared with
theoretical AGB evolutionary models, particularly those that predict enhanced mass loss
rates due to thermal pulsation near the end of the AGB. The spatial coverage in this study
between ∼ 200 to 1000 arcseconds from the central star translates into a probe of the history
of mass loss between 2 × 104 yrs to 1 × 105 years ago for AFGL 2688 and 6 × 104 yrs to
3 × 105 yrs ago for OH 231.8, using an expansion velocity of 20 km s−1 for both sources. The
models by Vassiliadis & Wood (1993) show that for a 2.0 M⊙ progenitor, during the last few
×105 yrs of AGB evolution, there are several thermal pulses which result in enhanced mass
loss rates peaking at about 1.3 × 10−5 M⊙ yr−1 during the end of each pulse. Using a gas to
dust mass ratio of 200, we find that for AFGL 2688 and OH 231.8, the upper limit to the
total mass loss rate during the above time intervals is about 4 × 10−5 M⊙ yr−1 and 2 × 10−5
– 14 –
M⊙ yr−1, respectively. These limits are close to the sensitivity necessary to see shells that
may be the result of thermal pulsation on the AGB. Although we do not detect the shells
reported by Speck et al. (2000), which they attribute to thermal pulses, the signatures of
thermal pulses may still be present, but below our current sensitivity.
We also consider the possibility that OH 231.8 does not show extended emission in
the far infrared because of interactions with a binary companion, which caused the central
star to lose mass more rapidly and more recently than during the evolution of a lone AGB
star. A companion to QX Pup (the central star of OH 231.8), is evidenced by optical
spectra consistent with a companion of stellar type A0 V (S´anchez Contreras et al. 2004).
QX Pup also shows a peculiar paradox of being an M9III (Cohen 1981) AGB star, while its
bipolar activity and morphology display all the signs of post-AGB activity of typical PPNe.
Having a close companion would enhance the mass loss rate and provide a mechanism for
generating the bipolar outflows (Morris 1987), thus shortening its mass loss history enough
that we should not be surprised to see no emission far from the source.
In contrast to
the collimated outflow from OH 231.8 (Alcolea et al. 2001), AFGL 2688 has a spherically
symmetric envelope seen in 13CO (Yamamura et al. 1996) and evidenced by the partial,
concentric, circular arcs present in HST images (Sahai et al. 1998). Since a binary interaction
would not cause the past spherically symmetric mass loss, a possible companion is probably
not a good explanation for initiating mass loss, although the present bipolar morphology of
AFGL 2688 is consistent the possibility that a binary interaction has altered the mass loss
in more recent times.
Tracing the mass loss history during the AGB phase should be relatively straightforward
via mapping the emission from the circumtellar envelope. But in practice, it has been
difficult because molecular-line observations are ultimately limited by the photodissociation
of molecules in the outer envelope regions, and far-infrared observations of dust emission have
been limited by the generally low angular resolution of the space-based telescopes which have
been available for this purpose in the past (IRAS and ISO). Hence the reported detections of
very extended emission in a few dying stars with IRAS and ISO has generally been recognized
as an important milestone in the study of mass loss. However, such detections have also raised
the very important question of why only a few select objects like Y CVn (Izumiura et al.
1996) or RY Dra (Young et al. 1993) – not particularly known for having high mass loss rates
– show extended envelopes, whereas large numbers of stars having high CO-determined mass
loss rates do not reveal the presence of such envelopes. Is this because the radial density
law for most of these high mass-loss stars is significantly steeper than in objects like Y CVn
or RY Dra, and in particular steeper than r−2 (an issue of profound importance for the
evolutionary times of stars through the AGB phase and theories of mass-loss), or are the
claimed detections of extended envelopes really a result of poorly characterised instrumental
– 15 –
artifacts? The Spitzer data presented in this paper clearly show that the presence of these
shells cannot be confirmed at a level well below the intensities expected from Speck et al.'s
results. Our non-detections call into question not only the ISO results on AFGL 2688, but
all results on the detection of extended envelopes in other objects using the same linear scan
technique described by Speck et al.. More detailed mapping of many high mass-loss rate
objects like AFGL 2688, to search for such shells, is crucially needed.
5. Summary
Spitzer observations of extended envelopes of expanding, dusty outflows from bipolar
AGB and post-AGB stars reveal that there may be a very large dust shell having a radius of
400 arcseconds around IRAS 16342−3814. The combination of the presence of nearby cirrus
emission and the limited azimuthal coverage in the images makes the conclusion for a shell
uncertain, but if the shell is indeed the result of an episodic mass loss event, then it would
represent one of the largest dust shells found so far.
Our observations of AFGL 2688 at 70 µm do not show the dust shells at 150 and
300 arcseconds from the source reported by Speck et al. (2000). Since the dust shell may
be very cool, the non-detection at 70 µm does not alone rule out dust shells. However,
we find that there is only a slight excess at 160 µm of 2 MJy Sr−1 above the background
at the reported location of the outer shell (∼ 300 arcseconds). With greater azimuthal
coverage than was previously obtained with ISOPHOT we find that there is substantial
contamination by galactic cirrus in the region at 160 µm, with galactic cirrus emission above
the sky background present throughout the entire eastern scan path of our observations.
Unfortunately, at 160 µm we can only set an upper limit to the emission from a dust shell
at 150 arcseconds from the source because one of the Airy rings of the PSF overlaps that
region. We also see no extended emission from OH 231.8 at either 70 or 160 µm other than
that attributable to galactic cirrus. In fact, using OH 231.8 as an empirical point source for
PSF subtraction from AFGL 2688 at 70 µm appears to substantially reduce PSF subtraction
residuals compared to using the model PSF.
The limitation of our method of observation is that we have only two radial directions to
probe possible extended emission. For cases like IRAS 16342−3814 where the shell is patchy,
better azimuthal information would help to resolve whether the origin of the emission is from
the star or from galactic cirrus. The MIPS data show that observing very extended emission
in the far infrared is possible, but difficult because of the prominence of galactic cirrus
emission at these wavelengths.
– 16 –
This work is based on observations made with the Spitzer Space Telescope, which is
operated by the Jet Propulsion Laboratory, California Institute of Technology under a con-
tract with NASA. Support for this work was provided by NASA through an award issued
by JPL/Caltech. MM and RS was partially funded for this work from a GO Spitzer award
and an LTSA award (no. 399-20-40-06) from NASA.
Alcolea, J., Bujarrabal, V., S´anchez Contreras, C., Neri, R., & Zweigle, J. 2001, A&A, 373,
REFERENCES
932
Cohen, M. 1981, PASP, 93, 288
Draine, B. T., & Lee, H. M. 1984, ApJ, 285, 89
Gillett, F. C., Backman, D. E., Beichman, C., & Neugebauer, G. 1986, ApJ, 310, 842
Hawkins, G. W. 1990, A&A, 229, L5
Izumiura, H., Hashimoto, O., Kawara, K., Yamamura, I., & Waters, L. B. F. M. 1996, A&A,
315, L221
Jura, M., & Morris, M. 1985, ApJ, 292, 487
Mathis, J. S., Mezger, P. G., & Panagia, N. 1983, A&A, 128, 212
Morris, M. 1987, PASP, 99, 1115
Rieke, G. H., et al. 2004, ApJS, 154, 25
Sahai, R., Hines, D. C., Kastner, J. H., Weintraub, D. A., Trauger, J. T., Rieke, M. J.,
Thompson, R. I., & Schneider, G. 1998, ApJ, 492, L163+
Sahai, R., Te Lintel Hekkert, P., Morris, M., Zijlstra, A., & Likkel, L. 1999, ApJ, 514, L115
S´anchez Contreras, C., Gil de Paz, A., & Sahai, R. 2004, ApJ, 616, 519
Skinner, C. J., Meixner, M., Barlow, M. J., Collison, A. J., Justtanont, K., Blanco, P., Pina,
R., Ball, J. R., Keto, E., Arens, J. F., & Jernigan, J. G. 1997, A&A, 328, 290
Speck, A. K., Meixner, M., & Knapp, G. R. 2000, ApJ, 545, L145
Ueta, T., Murakawa, K., & Meixner, M. 2006, ApJ, 641, 1113
– 17 –
Vassiliadis, E., & Wood, P. R. 1993, ApJ, 413, 641
Yamamura, I., Onaka, T., Kamijo, F., Deguchi, S., & Ukita, N. 1996, ApJ, 465, 926
Young, K., Phillips, T. G., & Knapp, G. R. 1993, ApJ, 409, 725
Werner, M. W., et al. 2004, ApJS, 154, 1
This preprint was prepared with the AAS LATEX macros v5.2.
– 18 –
Object
AFGL 2688
OH 231.8
IRAS 16342−3814
Mean Before
Subtraction
(MJy Sr−1)
σ
(MJy Sr−1)
Mean After
Subtraction
(MJy Sr−1)
σ
(MJy Sr−1)
14.1
10.2
30.3
1.1
0.9
1.3
0.6
0.1
0.3
0.9
0.6
0.9
Table 1: The 70 µm mean surface brightness and sensitivity for a region of uniform back-
ground far from the source, before and after the median background removal described in
the text.
– 19 –
Object
AFGL 2688
OH 231.8
IRAS 16342−3814
Mean Before
Subtraction
(MJy Sr−1)
σ
(MJy Sr−1)
Mean After
Subtraction
(MJy Sr−1)
σ
(MJy Sr−1)
31.4
23.1
90.1
1.7
0.7
2.9
0.3
-0.2
1.0
1.5
0.7
2.3
Table 2: The 160 µm mean surface brightness and sensitivity for a region of uniform back-
ground far from the source, before and after the median background removal described in
the text.
– 20 –
70 µm
160 µm
140
120
100
80
60
40
20
0
64
56
48
40
32
24
16
8
0
14
12
10
8
6
4
2
0
−2
8
8
6
2
L
G
F
A
8
.
1
3
2
H
O
4
1
8
3
-
2
4
3
6
1
S
A
R
I
20.0
17.5
15.0
12.5
10.0
7.5
5.0
2.5
0.0
14
12
10
8
6
4
2
0
40
35
30
25
20
15
10
5
0
Fig. 1.- The three PPNe in our sample after mosaicking and background subtraction. The
units of the colorbars are in MJy Sr−1. The bars in the images are 200′′ in length for scale
and the × marks the location of the point source. Top: AFGL 2688. Center: OH 231.8.
Bottom: IRAS 16342−3814. Note the prominence of the PSF features at 70 µm in all the
images. At 160 µm, the PSF is not apparent in IRAS 16342−3814. North is up and east is
to the left.
– 21 –
Fig. 2.- Model stinytim PSFs in a log stretch to emphasize the Airy rings and diffraction
spikes. Left: The 70 µm PSF. Right: The 160 µm PSF. The lines in the figures are 200
arcseconds in length. For scale, the surface brightness of the diffraction spikes ∼ 200 arc-
seconds from the center is ∼ 10−4 times lower than at the center. See Figures 3 and 5 for
azimuthally averaged plots of the model PSF.
– 22 –
Fig. 3.- PSF subtracted images of AFGL 2688. The color stretches are linear, with the
units in MJy Sr−1. The x marks the location of the point source while the line represents
200′′ for scale. a) PSF subtraction using the stinytim model PSF at 70 µm. The three PSF
diffraction spikes visible in Figure 1 appear to subtract differently, leaving different levels
of residuals. b) PSF subtraction at 70µm using OH 231.8 as a PSF, showing almost no
residuals except for the region at about 60 degrees east of north. c) PSF subtraction using
the stinytim PSF at 160 µm. The feature in the northern scan path coincides with the
position of the diffraction spike in the model PSF. The images have the same orientation as
Figure 1, with north being up.
– 23 –
Fig. 4.- Left: PSF subtraction of OH 231.8 using the model PSF at 70 µm. The residuals
are very similar to those resulting from the PSF subtraction of AFGL 2688 in Figure 3.
Right: PSF subtraction at 160 µm. The northern scan path has more structure than the
eastern scan, probably from galactic cirrus. The units of the colorbars are in MJy Sr−1. The
x marks the location of QX Pup, the central star of OH 231.8. North is up as in Figure 1.
– 24 –
AFGL 2688 70 microns Azimuthal Plots
10000
AFGL 2688 70 microns PSF Subtraction
30
AFGL 2688 70
OH 231.8
Model PSF
Subtraction using Model PSF
Subtraction using OH 231.8
)
r
S
/
y
J
M
(
s
s
e
n
h
g
i
r
t
B
1000
100
10
0
100
200
300
Distance from Center (arcseconds)
)
r
S
/
y
J
M
(
s
s
e
n
h
g
i
r
t
B
25
20
15
10
5
0
400
−5
0
100
200
300
400
Distance from Center (Arcseconds)
Fig. 5.- Left: The azimuthally averaged surface brightness of AFGL 2688, OH 231.8, and
the model PSF at 70 µm. The model PSF and OH 231.8 are scaled to the Airy ring from the
AFGL 2688 PSF at 76 arcseconds. Right: The result of the PSF subtractions. The surface
brightness drops to background levels beyond about 250 arcseconds. The excess emission
left from the PSF subtraction using the stinytim model PSF is very similar to the residuals
left after the same model PSF subtraction of OH 231.8 in Figure 7. The PSF subtraction
using OH 231.8 as a template PSF instead of the model PSF is shown with a dashed line; the
excess emission from the subtraction using the model PSF is almost completely eliminated.
– 25 –
AFGL 2688 160 microns Azimuthal Plots
50
AFGL 2688 160 microns PSF Subtraction
20
Average of both Scans
Eastern Scan Path
Northern Scan Path
Model PSF
ISOPHOT 180 microns
ISOPHOT 120 microns
200
400
600
800
Distance from Center (arcseconds)
)
r
S
/
y
J
M
(
s
s
e
n
h
g
i
r
t
B
40
30
20
10
0
0
)
r
S
/
y
J
M
(
s
s
e
n
h
g
i
r
t
B
15
10
5
0
1000
−5
0
Average of both Scans
Eastern Scan Path
Northern Scan Path
ISOPHOT 180 microns
ISOPHOT 120 microns
200
400
600
800
1000
Distance from Center (Arcseconds)
Fig. 6.- Left: Azimuthally averaged surface brightness of AFGL 2688 at 160 µm. The
two scan paths of the observations are plotted separately to show the differences in their
background levels. The solid line represents the average of the two scans. Beyond about
750 arcseconds from the source, the northern scan path stops so only the eastern path is
represented beyond this distance. The PSF is scaled to the brightness of the Airy ring at
170 arcseconds from the source. Right: The result of PSF subtraction. The region less
than 200 arcseconds from the source may be affected by the PSF subtraction so we cannot
make strong quantitative statements about the emission there at this time, except that the
surface brightness of this region is likely to be about 2 MJy Sr−1.
ISOPHOT data from
Speck et al. (2000) are also plotted for comparison after subtracting the background offset
determined from the median of the values beyond about 400 arcseconds from the source.
The reported ISOPHOT intensity at 300 arcseconds is almost eight times higher than the
surface brightness seen by MIPS at 160 µm after PSF subtraction.
– 26 –
Fig. 7.- Left: the azimuthally averaged surface brightness of OH 231.8 and the model PSF
at 70 µm. The model PSF has been scaled to match the Airy ring emission of OH 231.8 at a
radius of 76 arcseconds. Right: the result of subtraction of the model PSF. The shape of the
residuals is almost the same as those left from the PSF subtraction of AFGL 2688 (Figure
5), falling off to background levels beyond 250 arcseconds from the source.
– 27 –
OH 231.8 160 microns Azimuthal Plots
10000.0
OH 231.8 160 microns PSF Subtracted
10
Source
Eastern Scan Path
Northern Scan Path
Model PSF
100
Distance from Center (arcseconds)
500
400
200
300
)
r
S
/
y
J
M
(
s
s
e
n
h
g
i
r
t
B
8
6
4
2
0
600
−2
0
1000.0
)
r
S
/
y
J
M
(
s
s
e
n
h
g
i
r
t
B
100.0
10.0
1.0
0.1
0
Average Azimuthal Brightness
Eastern Scan Path
Northern Scan Path
100
Distance from Center (Arcseconds)
200
400
300
500
600
Fig. 8.- Left: the azimuthally averaged surface brightness of OH 231.8 along with the
model PSF at 160 µm. Right: the 1-D PSF subtraction. There is no apparent enhanced
emission other than attributable to fluctuations from the galactic cirrus.
– 28 –
100
96
92
88
84
80
76
72
68
)
1
−
r
S
y
J
M
(
s
s
e
n
t
h
g
i
r
B
e
c
a
f
r
u
S
Fig. 9.- The MIPS 70 µm field of view of IRAS 16342−3814 overlaid on the IRAS 100 µm
image of the surrounding regions. Note the abundance of galactic cirrus. The bar above the
frame is 400 arcseconds in length, representing the radius of the potential shell seen with
MIPS around IRAS 16342−3814.
– 29 –
Fig. 10.- Azimuthal average and PSF subtraction of IRAS 16342−3814 at 70 µm. Left:
As with Figure 5, this plot shows the source plotted along with the radial profiles of OH
231.8 and the model PSF. Unlike AFGL 2688, IRAS 16342−3814 shows a definite excess
in its radial profile above the background and PSF profile. Right: the difference between
using the model and OH 231.8 for the PSF subtraction. Both give about the same level of
excess emission of ∼ 2.5 MJy Sr−1 between 300 to 400 arcseconds from the source, where we
see a patchy shell in the MIPS 70 and 160 µm images.
– 30 –
Fig. 11.- Left: The azimuthally averaged surface brightness of IRAS 16342−3814 and the
model PSF scaled to the brightness of the source at the location of the 170 arcsecond radius
Airy ring of the 160 µm PSF. Right: the result of PSF subtraction. The excess emission
seen in the MIPS images is clearly present out to 400 arcseconds.
– 31 –
Fig. 12.- Plot of the expected surface brightness for various mass loss rates. The solid line
shows the expected surface brightness for a smooth constant mass loss envelope with a radius
of 4.2 pc while the dashed lines show the radial profile of a single shell of 1 pc thickness at
the same mass loss rate. The diamonds show the radial profile of IRAS 16342−3814 after
the subtraction of the model PSF. We find that the best fit to the radial brightness profile
obtained from subtracting the model PSF is a two-component model with a shell of radius
4.2 pc and thickness 1 pc requiring a dust mass loss rate of 10−6 M⊙ yr−1 superimposed
upon an envelope from a constant dust mass loss of 10−7 M⊙ yr−1 (dashed blue line). The
asterisks show the radial profile found from using OH 231.8 as a substitute for the model
PSF in the PSF subtraction will fit using only the shell component with a dust mass loss
rate of 10−6 M⊙ yr−1 (dashed orange line). Points closer than a radial distance of ∼ 150
arcseconds are unreliable because they are most affected by PSF subtraction residuals and
are not used in the fit.
|
astro-ph/9510120 | 1 | 9510 | 1995-10-23T18:50:31 | The velocity gradient in the pseudo-photosphere of the peculiar supergiant HD101584 | [
"astro-ph"
] | In this paper preliminary results are presented based on a study of the low and high resolution ultraviolet spectrum of the peculiar supergiant (post-AGB star) HD101584. By a comparison of the low resolution spectrum (1200-3200ang) with standard stars, the star is classified as an A7I, indicating an effective temperature of 8150 K, where literature quotes spectral type F0I. The Doppler shift of the FeII absorption lines in the high resolution spectrum (2500-3000ang) show a relation with the line optical depth. This suggests an expanding accelerating wind, c.q. pseudo-photosphere. The relation is extended by a factor 10^5 in optical depth by using available data from optical HeI and NI lines. The relation suggests that the radial heliocentric velocity of the star is at least 54.5km/s. From the Halpha line a velocity of 96km/s is measured for the terminal velocity of the wind. | astro-ph | astro-ph |
1
THE VELOCITY GRADIENT IN THE PSEUDO-PHOTOSPHERE
OF THE PECULIAR SUPERGIANT HD 101584
ERIC J. BAKKER
SRON Laboratory for Space Research Utrecht
Sorbonnelaan 2
3584 CA UTRECHT
The Netherlands
Abstract
In this paper preliminary results are presented based on a study of the low and
high resolution ultraviolet spectrum of the peculiar supergiant (post-AGB star)
HD 101584. By a comparison of the low resolution spectrum (1200 − 3200 A) with
standard stars, the star is classified as an A7I, indicating an effective temperature
of 8150 K, where literature quotes spectral type F0I. The Doppler shift of the FeII
absorption lines in the high resolution spectrum (2500 − 3000 A) show a relation
with the line optical depth. This suggests an expanding accelerating wind, c.q.
pseudo-photosphere. The relation is extended by a factor 105 in optical depth by
using available data from optical HeI and NI lines. The relation suggests that the
radial heliocentric velocity of the star is at least 54.5 km s−1. From the Hα line a
velocity of 96 km s−1 is measured for the terminal velocity of the wind.
1
Introduction
The star HD 101584 (b = 6o) is classified as a 7.01 visual magnitude F0Iape with
(B − V) = +0.39 (Hoffleit 1983), indicating an effective temperature of 7700 K , log g =
1.7 , and thus (B − V)0 = 0.17 (Landolt-Bornstein 1982). Far- and near-infrared pho-
tometry reveals a strong infrared source at the position of the star (Humphreys & Ney
1974; Parthasarathy & Pottasch 1986). Molecular line observations show bipolar out-
flow for the OH maser (te Lintel Hekkert et al. 1992) and a very complex structure for
the CO(J = 1 → 0) transition (Trams et al. 1990; Loup et al. 1990; van der Veen et
al. 1992). The molecular line emission in OH and CO is normally discussed in terms
of evolved stars, and fits the idea that HD 101584 is a post-AGB star.
The first to classify the star HD 101584 as a post-AGB star were Parthasarathy
& Pottasch (1986). Their conclusion was based on the strong infrared excess of the star
which seems to be due to a large amount of dust around the star. An extensive study
by Trams et al. (1991) shows the resemblance of the infrared excess of HD 101584 with
other known post-AGB stars. It is now reasonably well established that HD 101584 is
a post-AGB star.
2 The ultraviolet spectrum
2.1 The low resolution IUE spectrum
The low resolution IUE spectra (9AA) of HD 101584, a A7I and a F0I standard star are
shown in figure 1. Before fitting the ultraviolet energy distribution to a standard star
the spectrum was smoothed over 12.6 A and corrected for interstellar and circumstellar
Typeset for LHLS workshop
2
Figure 1: The low resolution IUE spectrum of HD 101584 (upper), a standard A7I
star (middle), and a standard F0I star (lower). The spectra are smoothed over 12.6 A,
normalized, and dereddened
extinction (Mathis 1990) using a colour excess, E(B − V) = 0.27, derived for spectral
type A7I.
The low resolution ultraviolet spectrum was best fitted to the standard A7I star,
HD14873. An even better fit is possible if also an A6I star would be available in the
reference atlas (Heck et al. 1984). The spectrum of HD 101584 shows no flux lower
then 1400 A and can therefore not be fitted with a spectrum of a star of spectral type
A5I or hotter.
By comparing the spectrum of HD 101584 with the standard F0I star, α Lep,
an excess of flux for the program star between 1400 and 1700 A indicates a higher
temperature of the star. The slope of the continuum of the spectrum confirms the
supergiant nature of the star. The data on the spectrum of HD 101584 and on the two
reference stars is in table 1.
2.2
the high resolution IUE spectrum
In an extensive study of the high resolution (0.3 A) IUE spectrum of HD 101584 a large
number of absorption features between 2500 and 3000 A has been identified (Bakker
1994). The main conclusions from this work are that the spectrum of HD 101584 has
in principle the same absorption features as the F0 supergiant, α Lep, but the lines are
intrinsically broader and the lines are asymmetric in shape. This study limits itself to
the measured radial velocities of the absorption features as derived from the measured
Doppler shift of the core of the absorption profile.
Typeset for LHLS workshop
3
Table 1: Data on the low resolution spectra
HD number Name
Spectral E(B-V) Normalized on
[erg cm−2 s Hz]
type
4.1 10−23
A7I
6.0 10−22
α Lep F0I
2.8 10−23
A7I
0.27
0.04
0.25
HD 101584
HD 36673
HD 148743
Table 2: Stellar parameters for HD 101584 based on literature and on the low resolution
ultraviolet IUE spectra studied in this work
Work
Hoffleit 1983 F0I
This study
A7I
Spectral Effective
Type
Temperature
7700
8150
log g
(B − V)0 E(B − V)
1.7
1.9
+0.17
+0.12
0.22
0.27
Teff , log g, and (B − V)o from Landolt-Bornstein (1982) and are based on the given
spectral type.
In probing the photosphere of a star the line optical depth is a measure for the
depth in the photosphere seen. The stronger the optical depth τ the more outside layers
of the photosphere will be probed. The range in optical depth from the ultraviolet
spectrum is limited to about a factor 104. By incorporating some of the available
optical data, which are from much weaker lines, the deeper layers of the photosphere
can be probed as well and the relation can be extended over a much wider range of τ
by a factor of 105 to 109. There are however two assumptions in making the relation
valid for a larger range of τ . The first is solar abundance ratios, and that NI, HeI,
and FeII are the dominant ionization stages of these elements in the stellar wind (or
pseudo-photosphere). The second is that the effective temperature derived from the
energy distribution in the ultraviolet represents the real temperature of the gas. If the
first assumption is violated, the separate elements will shift horizontally in the fig. 2. If
the second assumption is violated, the relation within FeII will change, and there will
occur a small horizontal shift between the different elements. The data on FeII, NI, HeI
and Hα will be published in a separate paper which is in preparation (Bakker 1995).
Figure 2 shows the relation between the logarithm of the strength of an absorption
line and the heliocentric radial velocity measured for that line. The crosses are the FeII
lines from the ultraviolet spectrum, the triangle the optical nitrogen line (8680.24 A),
and the asterisk the optical helium line (5875.618 A).
Vhelio(FeII) = 49.1 − 5.1 ×(cid:20)log N + log gf −
5040χ
Teff (cid:21)(cid:2)km s−1(cid:3)
(1)
A first order approximation of the relation for the FeII line is given by eq. 1. The
reason that the HeI and NI line are not used for this first order approximation is that is
seems logical to assume a constant velocity for weak lines. These lines are formed in the
deeper layers of the photosphere and are therefore least affected by the unknown force
which accelerates the pseudo-photosphere outwards. The dashed line in fig. 2 shows the
first order approximation based on only the FeII lines. It is surprising to see that the
HeI and NI line fit this relation almost perfectly.
Typeset for LHLS workshop
4
Table 3: Main parameters of the chemical elements used in probing the pseudo-
photosphere
Element
He
N
Fe
Solar Abundance
log N
10.93
7.96
7.60
Ion
HeI
NI
FeII
Excitation
Ionization
Energy (eV) Energy (eV)
24.587
14.534
16.16
20.87
10.29
7.870 → 4.48
Solar Abundance, and ionization energies are from Allen (1985).
Figure 2: Relation between the strength of an absorption line and the radial velocity
measured for that line, indicating that we are looking at an accelerating expanding
wind. The dashed line is a first order fit based on the FeII lines. The HeI and NI line
seem to fit this relation exceptionally well. Where θ is 5040/Teff, χ is the excitation
potential of the lower level of the transition, and A is the abundance relative to H by
number
Typeset for LHLS workshop
5
3 Terminal velocity of the stellar wind
An upper limit on the maximum out streaming velocity (a lower limit on the blue shift
of a line) is given by the absorption part of the Hα line profile. The edge of the Hα
line profile is pretty steep, implying that hydrogen column density at that velocity does
not slowly decrease due to expansion (and thus dilution) of the gas, but rather that the
edge in the profile is due to hydrogen at the terminal velocity of the stellar wind. The
maximum out streaming velocity is 120 km s−1, and the out-streaming velocity derived
from the Doppler shift of the core of the absorption profile is 96 km s−1. The latter
velocity is the terminal velocity of the wind. The first is the net velocity maximum due
to turbulent motion in the wind and its expansion. If we assume that the velocity of the
star is best represented by the velocity of the HeI line, a minimum heliocentric velocity
of −41 km s−1 is expected. The dotted line in fig. 2 represents the terminal velocity of
the wind as determined from the Hα profile.
4 Discussion
The data in the literature concerning the radial velocity of HD 101584 is very confusing.
It's variations are not well understood and is a fruitful base for wild speculations. From
this study a little light is shed on these variations. The relation as shown in figure 2
reveals that variations in radial velocity within one spectrum (FeII lines) can be un-
derstood in terms of an expanding accelerating photosphere or wind. In the following
discussion this line absorbing region will be called pseudo-photosphere. This relation
can be extended to lower optical depth by incorporating the helium (5875.618 A) and
nitrogen (8680.24 A) optical absorption lines. That this relation holds for helium sug-
gests that this line is from the same star and not from a yet unseen hot companion
star. To produce a helium absorption line the star has to be much hotter than spectral
type A7I, probably even a B-type star. This contradicts the fact that the low resolution
ultraviolet spectrum is best fitted with a A7I reference star
Although the relation does not seem to go to a constant velocity for weaker lines
(HeI and NI) is seems reasonable to assume that the velocity for the HeI line is the
stellar velocity. By monitoring the velocity of the HeI absorption line it should be
possible to make a statement about the binary nature of the star. Radial velocity
measurements of stronger lines do not only have a contribution of the radial velocity of
the star due to binarity (if this would be the case), but will also have a contribution
from the pseudo-photosphere. These two contributions will be very hard to disentangle.
A maximum velocity of the out streaming wind as determined from the Hα profile is
96 km s−1. This means that velocities of absorption lines are to be expected in the
range between −41 km s−1 and 54.5 km s−1.
Table 4 summarizes the stellar parameters of HD 101584 as determined in this
study from ultraviolet spectra.
Acknowledgements The author would like to thank Dr. Norman Trams for making
the HIRES spectrum of HD 101584 available for this work. Dr. Christoffel Waelkens
was so kind to make the optical data of HeI, NI, and Hα lines available. Dr. Rens
Waters and Prof. Dr. Henny Lamers made a large contribution to this work by having
many discussion with the author. The author was supported by grant no. 782-371-040
by ASTRON, which receives funds from the Netherlands Organization for the Advance-
ment of Pure Research (NWO). This research has made use of the Simbad database,
operated at CDS, Strasbourg, France.
Typeset for LHLS workshop
6
Table 4: Improved data on HD 101584
Ultraviolet spectral type
Effective temperature
E(B − V)
Heliocentric velocity of star
Terminal velocity of wind
A7I
8150 K
0.27
54.5 km s−1
96 km s−1
54.5 km s−1
Maximum helio. velocity expected
Minimum helio. velocity expected −41 km s−1
References
[1985] Allen C.W.: 1985, Astrophysical Quantities, third edition
[1994] Bakker E.J.: 1994, A&AS 103, 189
[1995] Bakker E.J.: 1995, A&A in accepted
[1984] Heck A., Egret D., Jaschek M., Jaschek C.: 1984, IUE Low-Dispersion Spectra
Reference Atlas, Part 1. Normal Stars
[1983] Hoffleit D, Saladyga M., Wlasuk P.: 1983, Supplement to the Bright Star Cata-
logue, Yale University Observatory, USA
[1974] Humphreys R.M., Ney E.P.: 1974, ApJ 190, 339
[1982] Landolt-Bornstein 1982, Numerical Data and Functional Relationships in Science
and Technology, New Series, Group VI, Vol. 2b.
[1992] te Lintel Hekkert P., Chapman J.M., Zijlstra A.A.: 1992, ApJ 390, L23 prepa-
ration
[1990] Loup C., Forveille T., Nyman L. A, Omont A.: 1990, A&A 227, L29
[1990] Mathis J.S.: 1990, ARA&A 28. 37
[1986] Parthasarathy M, Pottasch S.R.: 1986, A&A 154, L16
[1990] Trams N.R., van der Veen W.E.C.J., Waelkens C., Waters L.B.F.M., Lamers
H.J.G.L.M.: 1990, A&A 233, 153
[1992] Van der Veen W.E.C.J., Trams N.R., Waters L.B.F.M.: 1992, A&A, submitted
[1991] Trams N.R., Waters L.B.F.M., Lamers H.J.G.L.M., Waelkens C., Geballe T.R.,
Th´e P.S.: 1991, A&AS 87, 361
|
astro-ph/9511091 | 1 | 9511 | 1995-11-20T14:08:43 | The Ionization Fraction in Dense Clouds | [
"astro-ph"
] | We present submillimeter observations of various molecular ions toward two dense clouds, NGC 2264 IRS1 and W 3 IRS5, in order to investigate their ionization fraction. Analysis of the line intensity ratios by the way of statistical equilibrium calculations allows determination of the physical parameters: n(H2)~(1-2)e6 cm-3 and T(kin)~50-100 K. Column densities and abundances are also derived. Together, the abundances of the observed ions provide a lower limit to the ionization fraction, which is (2-3)e-9 in both clouds. In order to better constrain the electron abundance, a simple chemical model is built which calculates the steady state abundances of the major positive ions, using the observed abundances wherever available. With reasonable assumptions, good agreement within a factor of two with the observations can be achieved. The calculated electron fraction is x(e)= (1.0-3.3)e-8 in the case of NGC 2264 and x(e)=(0.5-1.1)e-8 for W 3 IRS5. In the first case, the high abundance of N2H+ requires a rather high cosmic ray ionization rate >1e-16s-1, even if all nitrogen is assumed to be in gas phase N2. For W 3 IRS5, ionized metals such as Fe+ and Mg+ could provide 60% of the electrons. | astro-ph | astro-ph | THE IONIZATION FRACTION IN DENSE CLOUDS
C. de Boisanger
, F.P. Helmich
, and E.F. van Dishoeck
;
Commissariat (cid:18)a l'Energie Atomique, Centre d'Etudes de Bruy(cid:18)eres-le-Ch^atel,
Service PTN, F- Bruy(cid:18)eres-le-Ch^atel, France.
Leiden Observatory, P.O.-Box , RA Leiden, The Netherlands
Received August ; Accepted: October
Astronomy and Astrophysics, in press.
5
9
9
1
v
o
N
0
2
1
9
0
1
1
5
9
/
h
p
-
o
r
t
s
a
Abstract
We present submillimeter observations of various molecular ions toward two dense
clouds, NGC IRS and W IRS, in order to investigate their ionization frac-
tion. Analysis of the line intensity ratios by the way of statistical equilibrium cal-
culations allows determination of the physical parameters: n(H
) (cid:24) ( (cid:0) )
cm
(cid:0)
and T
(cid:24) (cid:0) K. Column densities and abundances are also derived. Together,
kin
the abundances of the observed ions provide a lower limit to the ionization fraction,
which is ( (cid:0) )
in both clouds. In order to better constrain the electron abun-
(cid:0)
dance, a simple chemical model is built which calculates the steady state abundances
of the ma jor positive ions, using the observed abundances wherever available. With
reasonable assumptions, good agreement within a factor of two with the observations
can be achieved. The calculated electron fraction is x
= (: (cid:0) :)
in the case
e
(cid:0)
of NGC and x
= ( : (cid:0) :)
for W IRS. In the (cid:12)rst case, the high
e
(cid:0)
abundance of N
H
requires a rather high cosmic ray ionization rate >
s
,
+
(cid:0)
(cid:0)
even if all nitrogen is assumed to be in gas phase N
. For W IRS, ionized metals
such as Fe
and Mg
could provide % of the electrons.
+
+
Key words: ISM: molecules { ISM: clouds { ISM: individual: W IRS, NGC IRS
Introduction
The ionization fraction of dense clouds is a basic parameter of interstellar physics,
since it determines the e(cid:14)ciency with which magnetic (cid:12)elds couple with the gas and
ultimately controls the rate of collapse and star formation. Moreover, electrons and
ions play a ma jor role in interstellar chemistry, and their abundances provide infor-
mation on the nature and importance of sources which ionize the gas. It is well known
that in the di(cid:11)use gas, interstellar photons with (cid:21) >
A provide most of the ioniza-
(cid:23)
tion so that the electron fraction is of order
, equal to the abundance of the ma jor
(cid:0)
supplier C
. This fraction is thought to drop by several orders of magnitude inside
+
dense clouds, where the ionization mostly results from cosmic ray interactions. The
ma jor ions here are thought to be H
, H
, He
, HCO
, H
O
and possibly metal
+
+
+
+
+
ions (e.g. Fe
, Mg
). Unfortunately, except for HCO
, all of these species are very
+
+
+
di(cid:14)cult to observe directly for spectroscopic reasons.
The (cid:12)rst attempts to derive the electron abundance in dense clouds date back more
than years (Watson ; Gu(cid:19)elin et al. , ; Wootten et al. ). The basic
idea was to use the observed DCO
/HCO
ratio as a measure of the H
D
/H
ratio,
+
+
+
+
which, together with a model for the deuterium fractionation and the H
recombina-
+
tion, gives a limit on the electron abundance. Typical electron fractions derived in
this way were
(cid:0)
, with an order of magnitude variation from cloud to cloud
(cid:0)
(cid:0)
(Langer ). However, this method was abandoned in , when measurements
by Adams & Smith ( ) found the dissociative recombination rate of H
with e
+
to be negligibly small, thereby invalidating the earlier determinations and raising the
upper limits to
. The H
rate has subsequently been a heated topic of discussion
(cid:0)
+
in astrochemistry for more than a decade, preventing further studies of the ionization
fraction.
There are several reasons why a new attack on this problem is now both timely and
warranted. First, the latest set of laboratory measurements on the H
dissociative
+
recombination rate seem to converge to the same (relative rapid) value under inter-
stellar conditions (Amano ; Canosa et al. ; Larsson et al. ), although
some lingering uncertainties remain (Smith &
Spanel ). Second, many more ions
(cid:21)
can now be observed with the new sensitive, high frequency receivers: apart from
+
+
+
+
HCO
and its isotopes, these include HCS
, N
H
and H
O
(Phillips et al. ),
whereas good upper limits on H
D
exist for a number of regions (van Dishoeck et al.
+
). Third, the dissociative recombination rates of HCO
, HCS
, and N
H
have
+
+
+
also recently been measured in the laboratory (Rowe et al. ). Fourth, the exci-
tation and optical depth of the lines in dense clouds is now much better constrained
by observations of several submillimeter lines, thereby minimizing the uncertainties in
abundance determinations and physical conditions.
We present here observations of various ions (HCO
, H
CO
, DCO
, HCS
and
+
+
+
+
+
N
H
) in two di(cid:11)erent dense clouds, NGC and W IRS, in order to investi-
gate their ionization fraction. The clouds are chosen to have a range in density and
temperature (as determined from our earlier H
CO data) and to have complementary
+
+
H
O
and/or H
D
emission observations available (Phillips et al. ; van Dishoeck
et al. ), as well as infrared absorption line data on H
(Black et al. ) and
+
CO/H
(Lacy et al. ). Previous studies in cold dark clouds such as TMC-, with
n(H
) =
cm
, have found :
(cid:20) x
(cid:20) :
, depending on the amount of
e
(cid:0)
(cid:0)
(cid:0)
PAHs and metals included in the modeling (Schilke et al. ). The sources studied
in this work have higher temperatures and densities up to two orders of magnitude,
and it will be interesting to investigate whether the electron fraction in them is lower,
as predicted by theory. The observations and results are presented in Sections {,
and their analysis in Section .
Together, the abundances of these ions provide a (cid:12)rm lower limit to the ionization
fraction of the cloud.
In order to constrain the electron abundance, more detailed
modeling is needed. The basic idea is that the molecular ions are formed by protona-
tion of abundant neutral species in reactions with H
:
+
+
+
X + H
! XH
+ H
where X=CO, H
O, N
, CS: : : Their destruction is governed by reactions with CO,
H
O and O and by dissociative recombination with an electron in the gas phase:
+
XH
+ e ! X + H
Thus by measuring XH
and the abundances of the parents X for a su(cid:14)cient number
+
of species, it should be possible to constrain both the electron fraction and the H
+
abundance (or equivalently the cosmic ray ionization rate (cid:16)
) in dense clouds. This
H
+
technique was pioneered by Wootten et al. ( ) based on the HCO
/CO ratio, but
is extended here to include many more ion/neutral pairs. The actual modeling, which
involves a more detailed network of reactions, is presented in Section .
Observations
The sources studied in this work are both associated with active regions of star for-
mation. NGC IR is a well studied molecular cloud core (d (cid:24) pc) (Schwartz
et al. ; Kr(cid:127)ugel et al ). Its shape on an optical plate resembles a dark cone
(the \Cone Nebula"). Presumably an intermediate mass star is forming here. The
infrared source is bright (Allen ) and has been used as background light source
for infrared absorption line studies of gas and solid phase species. In particular, an
upper limit of :
exists on the H
abundance in the gas (Black et al. ), and
(cid:0)
+
a lower limit of .
on the CO/H
ratio (Lacy et al. ).
(cid:0)
W IRS is the brightest infrared source (L = :
L
(Ladd et al. )) in the
(cid:12)
W Giant Molecular Cloud core (d (cid:24) : kpc). A large amount of dense and warm
material surrounds the massive star forming inside (Dickel ; Dickel et al. ;
Hasegawa et al. ; Hayashi et al. ; Tieftrunk et al. ; Helmich et al. ).
Because of the enormous visual extinction (A
> mag) the ultraviolet radiation
V
from the young stars is not expected to in(cid:13)uence the ionization balance deep inside
the cloud.
The observations of NGC were performed primarily at the James Clerk Maxwell
Telescope (JCMT
) in February , with additional data taken in April, June and
November . The facility receivers at GHz (A), GHz (Bi) and GHz
(C) were employed. These are double side-band receivers with upper and lower side-
bands . GHz (A, Bi) or . GHz (C) apart. No systematic side-band checks
were performed, since the spectra are simple and often there is only a single line in the
spectrum. The image frequencies were compared against spectral-line catalogues and
no obvious blending of lines was found. As the backend, the Digital Autocorrelation
Spectrometer (DAS) was used in di(cid:11)erent spectral resolutions (mostly . , . and
. MHz/channel). To improve the signal to noise, the spectra were smoothed to
. MHz/channel and in some cases to . MHz/channel. The former corresponds
to (cid:24) . km s
at GHz. Integration times were generally minutes on+o(cid:11) for
(cid:0)
and GHz receivers but up to an hour for lines in the GHz window, because
of the higher system temperatures. This resulted in typical (cid:27) noise in T
of (A){
A
(cid:3)
(Bi) mK and { (C) mK per resolution element ( . MHz). The pointing
was checked regularly on OMC- and was found to be within
.
The calibration at the JCMT was performed with the chopper-wheel method.
In
February , the beam e(cid:14)ciency was somewhat uncertain, because of inaccuracies
in the shape of the telescope surface. However, careful comparison with earlier runs
showed a reduction by only (cid:24) %. This e(cid:11)ect is corrected for by using an e(cid:11)ective
main beam-e(cid:14)ciency of ., . and . for receiver A, Bi and C, respectively.
During other runs, values of ., . and ., respectively, were used. In general
we expect the absolute calibration to be accurate to about %; relative uncertainties
can be signi(cid:12)cantly smaller for all runs. The beam size of the JCMT at , and
GHz is (cid:24)
,
and
, respectively. For most spectra, position-switching by
was found to be appropriate, except for the
CO lines for which a larger switch
was used. Additional data on other species toward NGC IRS will be presented
by Schreyer et al. ( ).
A limited number of spectra of NGC were taken at the Caltech Submillimeter
Observatory (CSO
) in January with the GHz (H
CO at GHz) and
GHz (C
O {) receivers. The low ( MHz) and high ( MHz) resolution
acousto-optical spectrometers were used as backends. Pointing is reproducible within
on the CSO. Main beam e(cid:14)ciencies of . and . were adopted, respectively.
The CSO data at GHz refer to a beam of (cid:24)
; those at GHz to a
beam.
The James Clerk Maxwell Telescope is operated by the Observatories on behalf of the Particle
Physics and Astronomy Research Council of the United Kingdom, the Netherlands Organisation for
Scienti(cid:12)c Research, and the National Research Council of Canada.
The Caltech Submillimeter Observatory is operated by the Californian Institute of Technology
under funding from the U.S. National Science Foundation (AST - ).
Earlier observations by van Dishoeck et al. ( ) with the CSO have resulted in
upper limits on the H
D
abundance. These same spectra also reveal surprisingly
+
strong N
H
{ emission in this source in a
beam.
+
The observations presented here for W IRS form part of the large JCMT GHz
survey of three star-forming cores in the W cloud (IRS, IRS and W (H
O))
Figure : Observed spectra toward: a and b NGC IRS and c W IRS. Note
that the DCO
{ line toward NGC consists of two components. Only the
+
broad component was used in the analysis. The N
H
{ line represents the (cid:12)rst
+
+
detection of this line in the interstellar medium. The N
D
{ line may be present at
the (cid:27) level. Since W IRS is a well-known out(cid:13)ow source, only the central Gaussian
was used in the analysis. The line next to the HCS
{ feature at V
= (cid:0) km s
LSR
+
(cid:0)
is the
SO
(cid:0)
line from the image side-band.
;
;
between January and November . A complete log and details of the obser-
vations will be given in a separate publication concerning the spectral survey (Helmich
& van Dishoeck ; see also Helmich et al. for (cid:12)rst results). For all runs, the
facility receivers (A, A, Bi and C) were used. Prior to July , the acousto-
optical spectrometer AOSC ( MHz) was employed as the backend; subsequently
the DAS was used in di(cid:11)erent spectral resolutions, mostly in its MHz bandwidth
mode. Binning of two channels resulted in a typical resolution of . MHz/channel.
As for NGC , no systematic side band identi(cid:12)cation was performed, because the
line crowding in IRS, compared with the two other W sources in the spectral-line
survey, is such that there is little ambiguity and possible blends are easily identi(cid:12)ed.
The pointing was checked regularly on the nearby continuum source W (OH) and is
within
. The observations of W IRS were performed in the
beam-switch
mode, except for the
CO lines where a larger position switch was used.
Additional observations of the H
O
ion toward W IRS were obtained by Phillips
+
et al. ( ) using the CSO, whereas an upper limit on H
D
was obtained by van
+
Dishoeck et al. ( ). The latter study also reported detection of the N
H
{ line.
+
Results
Tables and summarize the resulting Gaussian (cid:12)t parameters for the various lines
for the two sources. Examples of spectra are given in Fig. a, b and c. Although the
pro(cid:12)les of the lower lines of DCO
and other species toward NGC suggest the
+
presence of two components (Gu(cid:19)elin et al. ; Schreyer et al. ), it was found
that the lines observed in this work are well represented by single Gaussians with little
spread in the line widths, the optically thick
CO { and HCO
{ lines being
+
exceptions. Most other lines are optically thin, or at most moderately optically thick.
For example, the C
S { line is weak, and the intensity ratio CS/C
S={ is only
slightly smaller than the cosmic [
S]/[
S] abundance ratio of . . As mentioned
above, N
H
is very strong in this source; the { line at GHz presented in Fig. b
+
is the (cid:12)rst detection of this line in any astrophysical source. N
D
{ appears present
+
as well, but only at the (cid:27) level.
The lines toward W IRS are also well represented by single Gaussians but there is a
much larger spread in the line widths. The molecules HCS
, C
S and C
S have much
+
narrower lines than e.g., H
CO
, HCO
and CO-isotopomers. It is easy to explain
+
+
the larger line widths for CO and HCO
by optical depth e(cid:11)ects, but this becomes
+
more di(cid:14)cult for the rarer species. Part of the explanation may lie in inaccurate
Table : Parameters of Gaussian (cid:12)ts for NGC
Line
T
(cid:1)V
T
dV
MB
MB
R
(K)
km s
K km s
(cid:0)
(cid:0)
CO (cid:0)
.
.
C
O (cid:0)
.
.
.
C
O (cid:0)
.
.
.
CS (cid:0)
.
.
CS (cid:0)
.
.
CS (cid:0)
(cid:20) .
C
S (cid:0)
.
.
.
HDO
(cid:0)
(cid:20) .
HCO
(cid:0)
.
.
.
+
+
H
CO
(cid:0)
.
.
.
+
H
CO
(cid:0)
.
.
.
DCO
(cid:0)
.
.
.
+
+
HCS
(cid:0)
.
.
.
+
N
H
(cid:0)
.
.
.
+
N
D
(cid:0)
.
.
.
[C i]
P
(cid:0)
P
.
.
.
Coordinates (B . ): (cid:11) = . , (cid:14) = .
All data were taken with the JCMT except for C
O (cid:0) .
The upper limits are (cid:27) in a . MHz channel.
determinations of the C
S and HCS
line widths, because these lines are so weak.
+
However, source structure can play a r^ole as well. A complete investigation is beyond
the scope of this paper; therefore, for each species its observed line width is adopted
in the analysis. The detection of C
S implies that even the C
S lines are slightly
optically thick. From the
CS line, however, it can be seen that the optical depth
must be small.
Table : Parameters of Gaussian (cid:12)ts for W IRS
Line
T
(cid:1)V
T
dV
MB
MB
R
(K)
km s
K km s
(cid:0)
(cid:0)
CO (cid:0)
.
.
CO (cid:0)
.
C
O (cid:0)
.
.
.
C
O (cid:0)
.
.
.
CS (cid:0)
.
.
.
C
S (cid:0)
.
.
.
C
S (cid:0)
.
.
.
C
S (cid:0)
.
.
.
C
S (cid:0)
.
.
.
CS (cid:0)
.
.
.
HCO
(cid:0)
.
.
+
+
H
CO
(cid:0)
.
.
.
+
H
CO
(cid:0)
.
.
.
+
HC
O
(cid:0)
.
.
.
DCO
(cid:0)
.
.
.
+
+
HCS
(cid:0)
.
.
.
+
HCS
(cid:0)
.
.
.
+
N
H
(cid:0)
< |